Journal Pre-proof Sintering Mechanism and Microstructure of TaC/SiC Composites Consolidated by plasma-activated sintering Jianian Hu, Guoqiang Luo, Jian Zhang, Yi Sun, Qiang Shen, Lianmeng Zhang PII:
S0272-8842(19)32766-X
DOI:
https://doi.org/10.1016/j.ceramint.2019.09.229
Reference:
CERI 22996
To appear in:
Ceramics International
Received Date: 31 August 2019 Revised Date:
19 September 2019
Accepted Date: 23 September 2019
Please cite this article as: J. Hu, G. Luo, J. Zhang, Y. Sun, Q. Shen, L. Zhang, Sintering Mechanism and Microstructure of TaC/SiC Composites Consolidated by plasma-activated sintering, Ceramics International (2019), doi: https://doi.org/10.1016/j.ceramint.2019.09.229. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier Ltd.
1
Sintering Mechanism and Microstructure of TaC/SiC
2
Composites Consolidated by Plasma-Activated Sintering
3 4
Jianian Hu , Guoqiang Luoa,*, Jian Zhanga, Yi Suna,
5
Qiang Shena and Lianmeng Zhanga
6
a
a
State Key Laboratory of Advanced Technology for Materials Synthesis and
7
Processing, Wuhan University of Technology, Wuhan 430070, China
8
* Corresponding author: Tel.: +86- 87168606; E-mail:
[email protected]
9 10 11 12
Abstract TaC/SiC
composites
with
5
wt.%
SiC
addition
were
densified
by
13
plasma-activated sintering (PAS) at 1500–1800 °C for 5 min under 30 MPa. The
14
effects of plasma-activated sintering on microstructures, densification and
15
mechanical properties of the composites were investigated. The results showed that
16
TaC/SiC composites achieved a relative density more than 99% theoretical density
17
at 1600 °C. A low eutectic liquid phase generated by the oxide on the particle
18
surface was observed in the composite to realize a relatively low temperature
19
sintering densification. While the TaC particle size decreased insignificantly with
20
increasing sintering temperature,the transformation of morphology of SiC particles
21
changing from equiaxed to elongated grain was activated, accompanying with a
22
slight particle size decreasing of the SiC phase, thus promoting a relatively high 1
23
flexural strength of 550 MPa under 1800 °C. Besides, some ultra-fine 2 nm Ta2Si
24
was observed in the glassy pockets, strengthening the amorphous phase and thus
25
increasing the flexural strength.
26
Keywords: TaC matrix; SiC additive; surface oxide; PAS sintering; sintering
27
mechanism; low eutectic glassy phase
28
1. Introduction
29
Ultra-high temperature ceramic (UHTC) was the potential candidates for a new
30
generation of promising materials for applications in the aerospace, which have a
31
relatively high melting point more than 3000 °C and possess a relatively high
32
strength under high temperature [1-4], Among UHTC, TaC has the highest melting
33
point at about 4000 °C, besides, TaC-based ceramics possess better mechanical
34
properties than other UHTC. TaC ceramics have a room-temperature flexural
35
strength in the range of 400–700 MPa [5-8], compared to ZrC and other UHTC of
36
220–400 MPa [9-10]. TaC shows better chemical corrosion resistance and excellent
37
thermoelectric performance and electrical conductivity. TaC is an ideal material for
38
structures operating at very high temperatures, such as parts of space shuttles and
39
rocket nozzles.
40
However, because of its high covalent bonding characteristics and low
41
self-diffusion coefficient, TaC is hard to densify [11-13]. In the case of TaC,
42
hot-pressed micro TaC powder at 30 MPa only achieved 94% of the theoretical
43
density at 30 MPa, 2300 °C [14]. It has been found that oxide impurities of Ta2O5 in
44
the starting powders are an important cause of excessive growth of TaC grains 2
45
[15-16].
46
Relative fine TaC powders with a particle size less than 2 µm were
47
consolidated at 2300 °C without pressure, reaching a relative density of 97.5%
48
without the use of additives [31], the small particle size of TaC powder turned out
49
to be an important factor for preparing a dense TaC ceramic. Reducing additives
50
such as C and B4C were added into a TaC ceramic to promote sintering [17-18].
51
While TaC ceramic was introduced by 0.36 wt.% B4C as a sintering agent, the
52
relative density was increased from 89.1% to 98% at 2200 °C [19]. In the case of
53
B4C, the reaction between TaC and B4C particles generated TaB2 phase and graphite
54
C, physically immobilizing the TaC particles, thus avoiding the residual pores from
55
being trapped. SiC was also added to UHTC to enhance sintering and improve
56
mechanical properties. In addition to conventionally pinning grain boundaries and
57
refining the matrix grains, SiC also provides an oxide layer (SiO2) to react with the
58
matrix, generating a low eutectic mixture to enhance the consolidation [20-21,30].
59
Conventional sintering methods, such as pressureless sintering and hot-pressing
60
sintering, performed on UHTC need a very high sintering temperature with a
61
relatively high sintering time, which enhances grain growth but diminishes the
62
mechanical properties [31,32]. The literature has shown that spark plasma sintering
63
(SPS) can facilitate the densifying of UHTC. In the SPS process, the graphite mold
64
with the powder significantly increases heating rate via a strong current. Under the
65
heating rate of 100°C per min, the rapid heating speed prevents grain growth before
66
reaching the holding temperature, which is beneficial to ultrahigh-temperature 3
67
technology [22-23]. In addition to the advantages of SPS, plasma activated
68
sintering (PAS) is very helpful in reducing the negative effects of oxide layers on
69
sintering densification by providing a 30 seconds with supplied pulsed current before
70
sintering. In the PAS sintering process, by generating the discharge between
71
different particles, particle surface adsorbent was eliminated and the oxide layer was
72
broken down, thus contributing to sintering; in addition, the process produced the
73
composite at a lower temperature in a relatively short period of time [24-25].
74
The comparison of fracture strength of TaC-SiC system composites by several
75
methods was listed in table 1, hot pressing and cold isostatically pressing/hot pressing
76
method was difficult to density the TaC/SiC composites, thus TaSi2 and MoSi2 was
77
introduced into the system as sintering aid. The fracture strength of TaC/15 vol.%SiC
78
with 10 vol. % TaSi2 reached as high as 506 MPa by conventional hot pressing
79
sintering [33, 34]. However, the fracture strength of the TaC/ 20 vol.%SiC composites
80
without sintering aid sintered by SPS reached 682 MPa [21]. Fast current assisted
81
sintering is beneficial to the densification and fracture strength of TaC/SiC composites.
82
However, the use of PAS for the densification of TaC has not received much
83
attention. In our study, PAS was used to produce a TaC/5 wt.% SiC composite.
84
Sintering was carried out at 30 MPa while keeping the temperature lower than
85
1800 °C. It was found that using 5 wt.% SiC as an agent could produce the composite
86
with a relative density of 99.3% by PAS at 1600 °C for 5 min, by generating a low
87
eutectic mixture to enhance the densification. The flexural strength of the composite
88
of 550 MPa was obtained under 1800 °C due to the transformation of the equiaxed 4
89
SiC phase to the plate-like SiC phase and the enhancement of the low eutectic
90
mixture with the ultrafine Ta2Si phase. The effects of PAS on microstructures,
91
densification and mechanical properties of the composites were investigated.
92
2. Experiment and procedure
93
We used commercial-SiC powder (mainly -6H polymorph SiC, Aladdin,
94
Shanghai, China) and TaC (cubic TaC Aladdin, Shanghai, China) as raw materials
95
with diameters of 0.5-1 µm and 3-5 µm, respectively. The purities of TaC raw
96
powder were more than 99% and the main impurities were Ca and F. The purity of
97
the SiC was 99.5 wt%, and the main impurities were SiO2 and free carbon. TaC
98
and 5wt.% SiC powder mixture was dry-blended in a ZrO2 cylinder with ZrO2
99
balls as grinding media for 24 h. A batch of the mixed powder was put into 20 mm
100
inner diameter graphite mold. A power supply offered a DC pulse current
101
discharge (20 V, 100 A) to activate the particle surface in 30 s, and then raise the
102
temperature to the target temperature at a rate of 120 °C/min. A 30 MPa uniaxial
103
pressure was performed on the diameter graphite mold at the target temperature for
104
300 s, finally reduce the temperature to the target temperature at a rate of
105
60 °C/min, approximately.
106
For the microstructure, polished surfaces and the fracture morphology of the
107
materials were characterized by field-emission scanning electron microscopy
108
(FESEM, Quanta-250). For the detailed microstructure, JEM-2100F STEM was
109
performed on the samples to get the high-resolution transmission electron
110
microscopy (HRTEM) images of the materials. X-ray diffraction patterns were 5
111
obtained on the machine (XRD, Rigaku Ultima3) using Cu Kα radiation. The
112
flexural strength of the composite was characterized by the bending method on a
113
mechanical test system (Instron-5966).
114
3. Results and discussion
115
3.1 powder characterization
116
The TEM images of the TaC and SiC mixture after ball milling are shown in
117
Figure 1. The particle surface of the TaC and SiC particles were both surrounded
118
by a 2-3 nm coating, the coating materials was determined as Ta2O5 and SiO2 in
119
the related literature [15,16,30,35], respectively.
120
3.2 Phase identification
121
The XRD pattern of the TaC/SiC composite sintered at different
122
temperatures from 1500 °C to 1800 °C is shown in Figure 2. The main phase
123
composition is cubic TaC, corresponding to the starting ingredients. Because the
124
content of the SiC phase was small (5 wt.% SiC), 6H-SiC could hardly be
125
discerned in the XRD pattern of the composite (Figure 2). No other crystal phase
126
was shown in the X-ray diffraction pattern. As the sintering temperature
127
increased, the XRD peak of the TaC phase in different materials moved to a
128
lower position, indicating a systematic increase in the TaC lattice [26]. Therefore,
129
TaC/SiC composites can be consolidated by PAS without changing main
130
chemical composition.
131
3.3 Density
132
The relative density of the composites is listed in Table 2. The relative density 6
133
of the composites sintered under different temperatures was relatively high. The
134
composites have relative density more than 99% of the theoretical density when the
135
temperature was above1600 °C. The high densification of TaC/SiC composites was
136
partly explained PAS sintering system. TaC grains growing were prevented by the
137
fast heating rate of 120°C/min and sintering time of 5min, consequently alleviated
138
the pore entrapment. The high density can also be realized by the formation of a
139
liquid phase during sintering, which was observed in the microstructure in Figure 4
140
3.4 Microstructure
141
SEM images of the TaC/SiC composite microstructure at different sintering
142
temperatures are shown in Figure 3. The comparison of grey and black is TaC and SiC
143
phases, respectively. The secondary phase SiC was uniformly distributed in the TaC
144
matrix. As the sintering temperature increases, the grain size of the TaC phase
145
decreases. Given the temperatures of 1500 °C and 1800 °C, the average particle size
146
of the TaC and SiC were 2.8 µm and 1.7 µm, separately, compared to that of 4.7 µm
147
in pure TaC bulk without adding SiC produced by PAS sintered at 1800 °C for 5
148
minutes, which demonstrates the significant effect of small SiC particles on the
149
physical pinning of TaC particles. Due to the SiC particles pinning the grain
150
boundaries between TaC particles. Therefore, preventing excessive growth of TaC
151
particles and eliminating residual porosity [21]. The SiC particle size also changed
152
with increasing sintering temperature. The phenomenon was completely unanticipated,
153
however, it is demonstrated by comparing the results in Figures 3a and 3d. From the
154
images in Figures 3a and 3d for the composites sintered at 1500 °C and 1800 °C, the 7
155
particle size of SiC phase decreases from approximately 1.0 µm to 0.5 µm. In addition
156
to the SiC grain size, the morphology also changed. For the composites sintered under
157
a temperature below 1700 °C, the morphology SiC grains are mostly equiaxed,
158
nevertheless, anisotropic morphology of SiC was demonstrated in the composites
159
sintered under 1800 °C, thus promoting a better flexural strength. The main reason
160
for the equiaxed SiC transformation into a plate-like SiC was that the SiC phase
161
partially dissolved in a glassy layer and subsequently reprecipitated with an elongated
162
direction to lower the surface energy, which was observed in the HRTEM in Figure 5.
163
A transmission electron micrograph of the composite consolidated under 1600 °C
164
and a phase chemical analysis is shown in Figure 4. SiC and TaC particles were
165
distinguished by an electron diffraction pattern, which was shown in Figures 4b and
166
4c, separately. The electron diffraction of SiC and TaC suggested that the crystal types
167
of 6H-SiC and SiC were hexagonal and cubical, respectively. A stacking fault was
168
observed in intragranular SiC. Figure 4a also shows a glassy pocket with three particle
169
junctions, connecting the surface oxide of SiC and TaC in Figure 1. The glassy pocket
170
is thought to be generated by a reaction between SiO2 and Ta2O5 forming a low
171
eutectic glass, and the phenomenon of two oxides forming a low eutectic glass was
172
observed in a ZnO and SiO2 system [30]. In Figure 4f, the HRTEM image of this area
173
demonstrates amorphous feature of the phase. Such a smooth glassy pocket exists
174
merely around SiC particles boundary. As shown in Figure 4e, there was a clean
175
interface between SiC and TaC, and no oxide on the surface of TaC and SiC was
176
observed, suggesting the effect of PAS sintering on cleaning the oxide impurities on 8
177
the starting powder. Additionally, a parallel crystal plane of TaC (1 1 1) and SiC (1
178
01 2) was observed, and the interplanar space of the TaC (1 1 1) plane and SiC (1
179
01 2) plane was 0.257 nm and 0.251 nm, respectively, suggesting a relatively low
180
lattice mismatch: 0.02=(0.257-0.251)/0.257, thus creating a strong bonding between
181
TaC and SiC. As shown in Figure 4d, the glass capsules were revealed by EDS
182
analysis. In addition to Si, C and Ta, elements of F, Ca and O, which were introduced
183
by the impurities in the starting powder of the mixture of TaC and SiC. The main
184
reason for forming the glassy phase during the sintering process was that the oxide
185
layer of the surface SiO2 and Ta2O5 have a low eutectic, thus forming a small amount
186
of liquid under PAS high-temperature sintering. The liquid (Ta-Si-C-O) flows into the
187
grain boundary and collects most of the impurities (Ca, F). After sintering, rapid PAS
188
cooling quenches the sample, with a 100 °C/min cooling rate, preventing the liquid
189
phase from crystallizing and thereby forming an amorphous layer consisting of
190
Ta-Si-C-O and the impurity on the powder surface; thus, no significant glass layer
191
between the TaC/SiC or TaC/TaC particles was observed by HRTEM, as shown in
192
Figure 4a. Similar amorphous layers are typically formed around SiC particles
193
because the primary component of the amorphous layer is SiO2 from the EDS analysis,
194
which preferentially wets SiC. The formation of a small amount of liquid glassy phase
195
during the sintering process can effectively improve the densification, and therefore,
196
in addition to the physical pinning effect on the grain boundaries between TaC/TaC
197
particles of the SiC phase, the high relative density of the obtained TaC/SiC composite
198
material is partially explained by the low eutectic glassy phase. With introducing a 9
199
liquid phase during the sintering process, the densification of the composite was
200
effectively enhanced, thus partially accounting for the high densified TaC/SiC
201
composite.
202
When the temperature is further increased, as seen from the TEM microstructure
203
images of the TaC/SiC composite sintered under 1800 °C in Figure 5, we can observe
204
that there are very small nanoparticles with a diameter of 2 nm generated in the
205
amorphous layer with a crystal plane space of 0.257 nm, corresponding with the plane
206
space of Ta2Si (0 0 2); combined with the EDS analysis of this area, this suggests that
207
the nanoparticle was Ta2Si. The main reason is that some Ta and Si diffuses into the
208
glassy liquid during sintering and subsequently reprecipitates into nanoparticles with a
209
diameter of 2 nm in the cooling process. In addition, from the morphology change of
210
the SiC, we hypothesize that the elongation of SiC grains also depends on the liquid
211
generated in the sintering process with increasing sintering temperature. 6H-SiC grain
212
partially dissolved in the liquid and then reprecipitated on the direction to elongation,
213
thus reducing the surface energy of SiC and forming a plate-like SiC. This
214
transformation from equiaxed morphology to elongated morphology was found in the
215
SiC/ZrB2 ceramic composite prepared at elevated temperatures, but the reason for the
216
conversion is not mentioned in reference [27]. In PAS sintering, a high energy current
217
encouraged reverse vacancy jumping from the grain boundaries to the exterior of the
218
grains, thus reducing the net vacancy jump frequency [28,29], therefore, enhancing
219
the diffusion of TaC and SiC and contributing to generating a plate-like SiC phase and
220
Ta2Si phase. In any case, pore elimination benefits from the spread of species. 10
221
To examine the microstructure, we can use a schematic diagram to explain the
222
sintering mechanism of the PAS consolidating the TaC/SiC composite in Figure 6. In
223
the initial configuration, TaC particles contact with the SiC additive. Then, the
224
reaction between the oxide layer on the surface of the TaC and SiC was activated,
225
forming the liquid Ta-Si–O–C phase and removing the surface impurities such as F
226
and Ca, the reaction was illustrated as equation (1), this reaction leveled the
227
densification mechanism. Further increasing the sintering temperature, a certain
228
amount of Ta and Si dissolved in the liquid and reprecipitated into nanoparticles with
229
a diameter of 2 nm Ta2Si, as illustrated in Figure 6c and Figure 6d, which was
230
described by equation (2), thus strengthening the glassy area and promoting a better
231
mechanical property.
232
Eutectic Ta2O5 + SiO2 → Ta − Si − C − O(liqiud ) ( SiC / TaC )
233
dissolving (Ta − Si −C − O ) TaC + SiC → Ta2 Si reprecipitating
234
(2)
235
3.5 Fracture strength
236
Table 2 summarizes the flexural strength of the composite at room temperature. with
237
the sintering temperature increasing, the flexural strength of the material increases.
238
The minimum and maximum flexural strength of the composite sintered at 1500 °C
239
and 1800 °C are 465 MPa and 550 MPa, respectively. The promoted flexural strength
240
demonstrates the enhanced effect of the SiC additive on the TaC ceramic matrix. The
241
fracture morphology is shown in Figure 7. No significant defects were observed at the
242
origin of the crack, and the high uniformity of the microstructure and the dispersion of
(1)
11
243
SiC in the microstructure were verified, thereby increasing the strength of the material.
244
The fracture surface was dominated by interparticle breaks. A small amount of
245
intragranular fracture also occurred.
246
4. Conclusions
247
(1) A high densified TaC/SiC composite with 5 wt.% SiC as an addition agent was
248
prepared by PAS at temperatures from 1500 °C to 1800 °C under 30 MPa for 5
249
minutes, and the density reached 99.3% at 1600 °C.
250
(2) The sintering mechanism of the PAS sintered composites activated the oxide
251
layer on the surface of the TaC and SiC, forming a liquid Si–O–Ta–C phase and
252
cleaning the surface of impurities, further contributing to the densification.
253
(3) The reason for the plate-like growth of SiC was that SiC dissolved in the liquid
254
phase in the sintering process and grew along the elongation direction to reduce the
255
surface energy. Some nanoscaled Ta2Si was observed and was thought to be
256
generated by an increasing amount of Ta and Si dissolving in the liquid and
257
reprecipitating into nanoscaled Ta2Si particles, promoting a flexural strength of
258
550 MPa.
259
Acknowledgements
260
This work was supported by the National Natural Science Foundation of China
261
under Grant Nos. 51202175 and 51521001; the 111 Project under Grant No.
262
B13035; and the Joint Fund under Grant No. 6141A02022255.
263
Reference
264
[1] Yang H., Fang H., Yu H., et al, Low temperature self-densification of high 12
265
strength bulk hexagonal boron nitride, Nat. Commun. 10 (2019).
266
[2] Silvestroni L., Kleebe H. J., Fahrenholtz W. G., et al, Super-strong
267
materials for temperatures exceeding 2000°C, Sci. Rep. 7 (2017) 40730.
268
[3] Zeng Y., Wang D., Xiong X., et al, Ablation-resistant carbide
269
Zr0.8Ti0.2C0.74B0.26 for oxidizing environments up to 3,000°C, Nat.
270
Commun. 8 (2017) 15836.
271
[4] Cho J., Li Q., Wang H., et al, High temperature deformability of ductile
272
flash-sintered ceramics via in-situ compression, Nat. Commun. 9 (2018)
273
2063.
274
[5] Zhang X., Hilmas G. E., Fahrenholtz W. G., Densificaiton and mechanical
275
properities of TaC-based ceramics, Mater. Sci. Eng. A. 501 (2009) 37–43.
276
[6] Zhao G., Huang C., Liu H., et al, Preparation of in-situ growth TaC
277
whiskers toughening Al2O3 ceramic matrix composite, Int. J. of Refract. Met.
278
Hard Mater. 36 (2013) 122-125.
279
[7] Glechner T., Mayrhofer P. H., Holec D., et al, Tuning structure and
280
mechanical properties of Ta-C coatings by N-alloying and vacancy population,
281
Sci. Rep. 8 (2018) 17669.
282
[8] Cedillos-Barraza O., Manara D., Boboridis K., et al, Investigating the highest
283
melting temperature materials: A laser melting study of the TaC-HfC system, Sci. Rep.
284
6 (2016) 37962.
285
[9] Wu W. W., Zhang G. J., Kan Y. M., et al, Reactive Hot Pressing of ZrB2–SiC–ZrC
286
Ultra High Temperature Ceramics at 1800°C, J. Am. Ceram. Soc. 89 (2006) 13
287
2967-2969.
288
[10] Ağaoğulları D., Gökçe H., Duman İ., et al, Characterization investigations of
289
ZrB2/ZrC ceramic powders synthesized by mechanical alloying of elemental Zr, B
290
and C blends, J. Eur. Ceram. Soc. 32 (2012) 1447-1455.
291
[11] Cedillos-Barraza O., Grasso S., Al Nasiri N., et al, Sintering behaviour, solid
292
solution formation and characterisation of TaC, HfC and TaC–HfC fabricated by spark
293
plasma sintering, J. Eur. Ceram. Soc. 36 (2016) 1539-1548.
294
[12] Gu J., Fu Z., Pressure enhanced densification of TaC ceramics during flash
295
spark plasma sintering, J. Am. Ceram. Soc. 102 (2019) 98-103.
296
[13] Gu J., Xiang M., Ji W., et al, Synthesis, densification, and microstructure of
297
TaC TaB2 SiC ceramics, J. Am. Ceram. Soc. 101 (2018) 5400-5410.
298
[14] Zhang X., Hilmas G. E., Fahrenholtz W. G., Hot pressing of tantalum carbide
299
with and without sintering additives, J. Am. Ceram. Soc. 90 (2007) 393–401.
300
[15] Kelly J. P., Graeve O. A., Mechanisms of pore formation in high-temperature
301
carbides: Case study of TaC prepared by spark plasma sintering, Acta Mater. 84 (2015)
302
472-483.
303
[16] Lu Z., Liu L., Geng G., et al, Effect of boron addition on microstructure,
304
mechanical properties and oxidation resistance of TaC ceramics, Ceram. Int. 45 (2019)
305
6712-6717.
306
[17] You Y., Tan D. W., Guo W. M., et al. TaB2 powders synthesis by reduction of
307
Ta2O5 with B4C, Ceram. Int. 2017, 43(1): 897-900.
308
[18] Demirskyi D., Vasylkiv O., Analysis of the high-temperature flexural strength 14
309
behavior of B4C–TaB2 eutectic composites produced by in situ spark plasma sintering,
310
Mater. Sci. Eng. A. 697 (2017) 71-78.
311
[19] Zhang X., Hilmas G. E., Fahrenholtz W. G., Densificaiton, mechanical prop-
312
erties, and oxidation resistance of TaC–TaB2 ceramic, J. Am. Ceram. Soc. 91
313
(2008) 129–32.
314
[20] Liu L., et al., Microstructure and mechanical properties of the spark plasma
315
sintered TaC/SiC composites, Mater. Sci. Eng. A. 529 (2011) 479-484.
316
[21] Liu H., Liu L., Ye F., et al., Microstructure and mechanical properties of the
317
spark plasma sintered TaC/SiC composites: Effects of sintering temperatures, J. Eur.
318
Ceram. Soc. 32 (2012) 3617-3625.
319
[22] Deng P., Li S., Xin Z., et al, Effect of graphite content on properties of B4C
320
W2B5 ceramic composites by in situ reaction of B Gr WC[J]. J. Am. Ceram. Soc.
321
101 (2018) 3617–3626
322
[23] Lan W., Shu X., Yi F., et al, Rapid fabrication and phase transition of Nd and Ce
323
co-doped Gd2Zr2O7 ceramics by SPS, J. Eur. Ceram. Soc. 2018,38(7): 2863-2870.
324
[24] Wu C., Fang P., Luo G., et al., Effect of plasma activated sintering parameters on
325
microstructure and mechanical properties of Al-7075/B4C composites. J. Alloys.
326
Compd. 615 (2014) 276-282.
327
[25] Shen Q., Wu C., Luo G., et al, Microstructure and mechanical properties of
328
Al-7075/B4C composites fabricated by plasma activated sintering, J. Alloys. Compd.
329
588 (2014) 265-270.
330
[26] Rangaraj L., Divakar C., Jayarm V., Fabrication and mechanisms of densifica15
331
tion of ZrB2-based ultra high temperature ceramics by reactive hot pressing, J. Eur.
332
Ceram. Soc. 30 (2010) 129–38.
333
[27] Zhang S. C., Hilmas G. E., Fahrenholtz W. G., Mechanical properties of sintered
334
ZrB2–SiC ceramics, J. Eur. Ceram. Soc. 31 (2018) 93–901
335
[28] Sun Y., J., Kulkarni K., Aindow M., et al., Influence of electric current on
336
microstructure evolution in Ti/Al and Ti/TiAl3 during spark plasma sintering, J.
337
Alloys. Compd. 648 (2015) 1097-1103.
338
[29] Conrad H., Effects of electric current on solid state phase transformations in
339
metals, Mater. Sci. Eng. A287 (2000) 227.
340
[30] Demirkesen G., Effect of Al2O3 additions on the acid durability of a
341
Li2O-ZnO-SiO2 glass and its glass-ceramic, Ceram. Int. 29 (2003) 463-469.
342
[31] Xiang L., Cheng L., Fan X., et al, Effect of interlayer on the ablation properties
343
of laminated HfC–SiC ceramics under oxyacetylene torch, Corros. Sci. 93 (2015)
344
172-179.
345
[32] Rangaraj L., Divakar C., Jayaram V., Fabrication and mechanisms of
346
densification of ZrB2-based ultra high temperature ceramics by reactive hot pressing,
347
J. Eur. Ceram. Soc. 30 (2010) 129-138.
348
[33] Pienti L., Silvestroni L., Landi E., et al. Microstructure, mechanical properties
349
and oxidation behavior of TaC-and HfC-based materials containing short SiC fiber,
350
Ceram. Int. 41 (2015) 1367-1377.
351
[34] Silvestroni L., Pienti L., Guicciardi S., et al. Strength and toughness: The
352
challenging case of TaC-based composites. Composites Part B, 72 (2015) 10–20.
353
[35] Bernhardt J., Schardt J., Starke U., et al. Epitaxially ideal oxide–semiconductor 16
354
interfaces: Silicate adlayers on hexagonal (0001) and (0001̄) SiC surfaces. Appl. phy.
355
let. 74 (1994) 1084-1086.
356 357
Figure description:
358
Figure 1. (a) SEM morphology of the TaC and SiC mixture after ball milling, (b)
359
oxide impurities on the SiC surface, (c) oxide impurities on the TaC surface
360
Figure 2. XRD patterns of the TaC/SiC composites sintered under different
361
temperatures
362
Figure 3. SEM images of the microstructures of the TaC/SiC composites sintered
363
under different temperatures. (a) 1500 °C; (b) 1600 °C; (c) 1700 °C; (d) 1800 °C
364
Figure 4. (a) TEM micrographs of the TaC/SiC composite sintered at 1600 °C. (b), (c)
365
The selected area electron diffraction pattern of the TaC and the SiC, respectively. (d)
366
EDS analysis of the glassy phase. (e), (f) HRTEM image of the interface between
367
TaC/SiC and the SiC/glassy phase.
368
Figure 5. (a) TEM micrographs of the TaC/SiC composite sintered at 1800 °C. (b)
369
selected area HRTEM images of triangular grain boundary betwen TaC, SiC and
370
glassy phase. (c) selected area electron diffraction pattern of SiC (d), (e) the inverse
371
FFT images of the selected area and interplanar spacing measurement of this area,
372
respectively.
373
Figure 6. Schematic diagram of the sintering mechanism of the TaC/SiC composite
374
Figure 7. Fracture SEM images of the microstructures of the TaC/SiC composites
375
sintered under different temperatures. (a) 1500 °C; (b) 1600 °C; (c) 1700 °C; (d)
376
1800 °C 17
377
Table description:
378
Table 1 The comparison of fracture strength of TaC-SiC system composites by several
379
methods
380
Table 2 Sintering condition, relative density, and flexural strength of the
381
investigated materials
18
Table 1 Comparison of fracture strength of TaC/SiC system composites by several methods Sintered sample
Additive volume fraction (vol%)
Fracture strength
Hot pressing[33]
TaC - 10 TaSi2 - 15 SiC fibers
432MPa
Hot pressing[33]
TaC - 10 TaSi2 - 15 SiC
506 MPa
TaC - 10 MoSi2 - 15 SiC
348 MPa
TaC - 20 SiC
682 MPa
TaC - 17.78 SiC
550 MPa
Cold isostatically pressing/Hot pressing[34] SPS [21] PAS(this work)
Table 2 Sintering condition, relative density, and flexural strength of the investigated materials Sintering Condition
Relative Density (%)
Flexural Strength σ (MPa)
1500
,5 min
98.4
465
1600
,5 min
99.3
493
1700
,5 min
99.5
526
1800
,5 min
99.8
550
We declare that we have no financial and personal relationships with other people or organizations that can inappropriately influence our work, there is no professional or other personal interest of any nature or kind in any product, service and/or company that could be construed as influencing the position presented in, or the review of the manuscript entitled “Sintering Mechanism and Microstructure of TaC/SiC Composites Consolidated by Plasma-Activated Sintering”.