ArticlePDF Available

High strength mullite-bond SiC porous ceramics fabricated by digital light processing

Tsinghua University Press
Journal of Advanced Ceramics
Authors:

Abstract and Figures

Fabricating SiC ceramics via the digital light processing (DLP) technology is of great challenge due to strong light absorption and high refractive index of deep-colored SiC powders, which highly differ from those of resin, and thus significantly affect the curing performance of the photosensitive SiC slurry. In this paper, a thin silicon oxide (SiO2) layer was in-situ formed on the surface of SiC powders by pre-oxidation treatment. This method was proven to effectively improve the curing ability of SiC slurry. The SiC photosensitive slurry was fabricated with solid content of 55 vol% and viscosity of 7.77 Pa·s (shear rate of 30 s⁻¹). The curing thickness was 50 μm with exposure time of only 5 s. Then, a well-designed sintering additive was added to completely convert low-strength SiO2 into mullite reinforcement during sintering. Complex-shaped mullite-bond SiC ceramics were successfully fabricated. The flexural strength of SiC ceramics sintered at 1550 ℃ in air reached 97.6 MPa with porosity of 39.2 vol%, as high as those prepared by spark plasma sintering (SPS) techniques.
This content is subject to copyright. Terms and conditions apply.
High strength mullite-bond SiC porous ceramics fabricated by
digital light processing
JianSun1, JingdeZhang1,2, XuZhang1, ZiheLi3, JianzhangLi4, SijieWei1,5, WeibinZhang1, WeiliWang1,
GuifangHan1,
1 Key Laboratory for Liquid-Solid Structural Evolution and Processing of Materials (Ministry of Education), School of Materials Science
and Engineering, Shandong University, Jinan 250061, China
2 Key Laboratory of Special Functional Aggregated Materials, Ministry of Education, Shandong University, Jinan 250100, China
3 Materials and Structures Research Centre, Department of Mechanical Engineering, University of Bath, Bath BA2 7AY, UK
4 National Engineering Research Centre of Ceramic Matrix Composite Manufacture Technology, Xi’an Golden Mountain Ceramic
Composites Co., Ltd., Xi’an 710118, China
5 School of Physics and Materials Science, Changji University, Changji 831100, China
Received: September 20, 2023; Revised: November 11, 2023; Accepted: November 27, 2023
©The Author(s) 2024. This is an open access article under the terms of the Creative Commons Attribution 4.0 International License
(CC BY 4.0, http://creativecommons.org/licenses/by/4.0/).
Abstract: FabricatingSiCceramicsviathedigitallightprocessing(DLP)technologyisofgreatchallengeduetostronglight
absorption and high refractive index of deep-colored SiC powders, which highly differ from those of resin, and thus
significantlyaffectthecuringperformanceofthephotosensitiveSiCslurry.Inthispaper,athinsiliconoxide(SiO2)layerwas
in-situformedonthesurfaceofSiCpowdersbypre-oxidationtreatment.Thismethodwasproventoeffectivelyimprovethe
curing ability of SiC slurry. The SiC photosensitive slurry was fabricated with solid content of 55 vol% and viscosity of
7.77Pa·s (shear rate of 30 s−1). The curing thickness was 50μm with exposure time of only 5 s. Then, a well-designed
sinteringadditive wasaddedto completelyconvertlow-strength SiO2intomullite reinforcementduringsintering. Complex-
shapedmullite-bondSiCceramicsweresuccessfullyfabricated.TheflexuralstrengthofSiCceramicssinteredat1550in
airreached97.6MPawithporosityof39.2vol%,ashighasthosepreparedbysparkplasmasintering(SPS)techniques.
Keywords: digitallightprocessing(DLP);SiCceramics;pre-oxidation;mullite-bondSiC;mechanicalproperties
1Introduction
Due to the advantages of high strength/hardness, high thermal
conductivity, low thermal expansion coefficient, and excellent
temperature/corrosion/wear resistance [15], silicon carbide (SiC)
ceramics are widely used in many fields including thermal
insulating, filter, catalyst carriers, biology, and aerospace [610].
These excellent properties of SiC ceramics are due to a strong
Si–C covalent bond. However, this results in difficulty in sintering.
In addition, the high strength and hardness of the SiC ceramics
make it difficult to machine or process [11,12], which limits the
complexity of SiC components and therefore their applications.
As a novel concept, the additive manufacturing technology can
realize direct fabrication of complex structure ceramics without
machining [13]. Among a range of additive manufacturing
techniques, since the digital light processing (DLP) technique has
high precision and relatively fast production speed, DLP has been
widely applied in the fabrication of complex ceramic components
[1417]. The mechanism of photocuring is based on the
polymerization of the photosensitive resin under the irradiation of
specific wavelengths of ultraviolet (UV) light [18,19]. In the first
step, ceramic powders are added to this photosensitive resin to
form uniformly dispersed ceramic slurry. Then, the slurry is cured
layer by layer with UV light exposure, and simultaneously ceramic
particles are wrapped inside the cured layer to form the designed
shape. The DLP technique has been successfully applied in the
printing of Al2O3 [2022], ZrO2 [23], and other oxide ceramics.
However, deep-colored SiC has strong light absorption and a
refractive index much higher than that of resin. This seriously
reduces the penetration depth of light, and the amount of light
that can be utilized by photosensitive resin, thereby decreasing the
curing thickness of SiC slurry [24].
To solve this problem, researchers decreased the solid content
of SiC inside the slurry and/or increased the particle size of SiC
powders [24,25]. However, the sample fabricated with low-solid-
content slurry showed large shrinkage after the remove of
polymer agents, which induced defects and thus poor mechanical
properties of the sintered samples. In addition, the large particle
size of SiC also resulted in low strength of fabricated parts.
Another method to increase the curing thickness of the SiC slurry
is to coat shell materials with low absorbance on the surface of SiC
particles by physical or chemical methods. As a result, the shell
layer can reduce the absorbance and improve curing ability
[2629]. Interestingly, high-temperature oxidation treatment of
the SiC powders in air can generate in-situ SiO2 shell layers with
low absorbance, which has been proven to increase the curing
depth of the SiC slurry [26,27].
However, SiO2 has low mechanical properties and undergoes
phase change during the cooling process accompanied by volume
changes. The mismatch of the thermal expansion coefficient
(CTE) between cristobalite SiO2 and SiC (CTEcristobalite =
Corresponding author.
E-mail: gfhan@sdu.edu.cn
Journal of Advanced Ceramics
2024, 13(1): 53−62
https://doi.org/10.26599/JAC.2024.9220835 ResearchArticle
https://doi.org/10.26599/JAC.2024.9220835
0.5×10−6 −1, CTESiC = 4.9×10−6 −1) [30] results in weak bonding
between them and therefore the poor performance of the SiC
ceramics. All these factors limit mechanical strength of the printed
SiC components. In our previous work [31], we sintered SiC
porous ceramics at 1550 in air with the strength comparable to
that fabricated from the spark plasma sintering (SPS) technique.
The oxidation-derived SiO2 from SiC at high temperatures in air
was converted into high-strength mullite bonding phases and
whisker reinforcement due to the rational design of the composite
sintering additives. In the Al(OH)3–Y2O3–CaF2 composite
additive, Al(OH)3 provides an Al2O3 source to consume SiO2, and
Y2O3 promotes the liquid mass transfer and the formation
reaction of mullite, while CaF2 introduces gaseous mass transfer
and enhances the formation of mullite whiskers. By converting
low-strength SiO2 to the mullite bond phase and introducing
mullite whisker reinforcement, this technique has been proven to
effectively enhance mechanical properties of the SiC ceramics.
Therefore, in this study, fine SiC powders with an average
particle size of 5.28 μm were pre-oxidized to form a thin SiO2
shell. The curing depth of the thus fabricated photosensitive slurry
was significantly increased, and complex-shaped SiC components
were successfully fabricated. SiO2 was completely converted to
mullite by carefully adjusting the added amount of
Al(OH)3–Y2O3–CaF2 sintering additives. The fabricated mullite-
bond SiC ceramics achieved high flexural strength comparable to
that fabricated by the SPS method. This work therefore provides a
feasible strategy to enhance the curing ability of the SiC slurry and
the mechanical property of sintered SiC at the same time, such
that we hope to excite the whole society to realize a low-cost and
precise photo-polymerization fabrication of complex-shaped deep-
colored ceramic materials such as SiC and Si3N4.
2Experimental
2.1Materials
Commercial SiC ceramic powders (purity 99.9%, d50 = 5.288 μm,
Nangong Ruiteng Alloy Material Co., Ltd., China) were chosen as
raw materials. Aluminum hydroxide (Al(OH)3, purity 99.9%,
average particle size of 10 μm, Aladdin Chemical Co., Ltd., China),
yttria (Y2O3, purity 99.9%, average particle size of 1 μm, Aladdin
Chemical Co., Ltd., China), and calcium fluoride (CaF2, purity
99.9%, average particle size of 1 μm, Shanghai Maclin Biochemical
Technology Co., Ltd., China) were used as sintering additives.
Tetramethyl ammonium hydroxide aqueous solution (TMAH,
25 wt%, AR, Aladdin Chemistry Co., Ltd., China) was used as the
dispersant (1.6 wt% of ceramic powders). Photosensitive resin
used in this paper was commercial resin containing monomers
and photoinitiators. The photoinitiator was 2,4,6-trimethylbenzoyl-
diphenyl phosphorus oxide (TPO). The monomer was a
dimethacrylate based substance, and the dispersant was KOS110
(Guangzhou Tai Runding New Materials Co., Ltd., China).
2.2Experimentalprocedure
The SiC powders were evenly spread into an Al2O3 crucible and
put into a chamber furnace for pre-oxidation in air atmosphere.
The heating rate was 10 /min and dwelled at 1200 for 2, 4,
and 6 h. The pre-oxidized powders were ball milled at 300 r/min
for 2 h and then sieved through a 100-mesh sieve.
In our previous work [31], the amount of Al(OH)3 was
optimized as 6.4 wt% of that of the SiC powders. The amount of
Y2O3 was optimized as 1.6 wt% of that of SiC and Al(OH)3
powders. The amount of CaF2 was optimized as 2.0 wt% of that of
the Al(OH)3 powders. Since we implemented pre-oxidation in this
study and the debinding process in air also introduced oxidation
of SiC, the amount of Al(OH)3 was increased to 10, 12.5, and 15 wt%
to completely consume additional SiO2. Also, the amount of Y2O3
and CaF2 was also increased accordingly. All these powders were
put into anhydrous ethanol (analytically pure) to form a
suspension with solid loading of 50 vol%, and TMAH was added
to the suspension as the dispersant to enhance the mixing
uniformity of SiC and additive powders. After ball milling for 4 h
with a planetary ball mill at a rotating speed of 300 r/min, the
slurry was dried in a vacuum drying oven. In the process of
obtaining composite powders by drying, the drying temperature
was 90 , and the time was more than 24 h. The dried powders
were then crushed and ground through a 100-mesh screen to
obtain composite powders.
The composite powders were mixed with commercial
photosensitive resin to obtain ceramic slurry with solid loadings of
45, 50, and 55 vol%, and KOS110 (5 wt% of the total weight of
ceramic powders) was used as the dispersant agent. After the
ceramic slurry was planetary ball milled for 5 h at a rotating speed
of 300 r/min, a vacuum drying oven (ZK-025, Shanghai
Experimental Instrument Factory, China) was used for deforming
of 1 h to obtain uniformly dispersed photocurable ceramic slurry.
The temperature of the ceramic slurry in the vacuum defoaming
process was controlled at room temperature of about 25 .
The DLP printer (PC5003A-50, Xi'an Dianyun Biotechnology
Co., Ltd., China) with a UV light wavelength of 405 nm was
used in this study. The intensity of the light source used was
30 mW/cm2. This instrument used a digital mask device (DMD)
with a resolution of 1920 × 1080 pixels. The dot resolution on the
printing plane was 50 μm, and the control accuracy in the printing
direction was around 10 μm. For the printing parameters, the
single layer thickness and the UV light exposure time were set to
50 μm and 3 s, respectively. Printing was conducted layer-by-layer
until the designed shape was obtained. Furthermore, within the
scope of this investigation, ceramic green bodies were prepared
through the printing process and subsequently subjected to
sintering to yield the final ceramic products. The heating rate in
the debinding process was 1 /min till 600 . After holding for
1 h, the temperature was raised to 1200 at a rate of 5 /min,
and finally to the sintering temperature of 1450, 1500, and 1550
at a heating rate of 10 /min and holding time of 2 h.
2.3Characterizations
An X-ray diffractometer (D/max 2500 PC, Rigaku, Japan, Cu Kα,
λ = 0.1548 nm) was used to obtain X-ray diffraction (XRD)
patterns for sintered ceramic materials. A scanning electron
microscope (SEM; JSM-7800, JEOL, Japan) equipped with an
energy dispersion spectrometer (EDS) was used to study the
microstructure of powders and cross-sections of the sintered
ceramic materials. The viscosity of the ceramic slurry was
measured by a rotating rheometer (R/S+ Rheometer, Brookfield,
USA), and the test temperature was maintained at 25 using a
temperature controller (Brookfield-TC). A square (20 mm ×
20 mm) cured single layer was obtained by exposing the SiC slurry
inside a glass dish for 5–50 s using the DLP 3D printer digital
mask device (DMD). After cleaning and erasing the uncured
stock, a digital micrometer (211-101, Dongguan Sanliang
Measuring Tools Co., Ltd., China) was used to measure the curing
thickness for at least 5 points. The flexural strength was measured
in a microcomputer control electronic universal testing machine
(4505, Instron Experimental Equipment Trading Co., Ltd., USA)
54 J. Sun, J. Zhang, X. Zhang, et al.
J Adv Ceram2024,13(1):53−62
using a three-point bending test (3PBT) with a loading speed of
0.5 mm/min and a span of 30 mm for at least 5 samples (3 mm ×
4 mm × 35 mm). Apparent porosity was determined by
Archimedes drainage hydrostatic method with at least 5 samples.
3Results and discussion
3.1Pre-oxidationofSiCpowders
The XRD patterns of the raw SiC powders and those after pre-
oxidation are shown in Fig. 1(a). The red line presented XRD of
the raw materials used in the experiment, which well matched the
standard diffraction pattern of 6H–SiC. After the pre-oxidation
treatment at 1200 for 2, 4, and 6 h, the crystal structure of the
SiC powders did not change significantly, but a new phase peak at
2θ of around 21.7°, which should be due to the formed cristobalite
(SiO2). A local magnification of the main diffraction peak of the
cristobalite phase is shown in Fig. 1(b). It can be seen that the
diffraction peak intensity of cristobalite gradually increased with
the prolongation of pre-oxidation time. The phase composition
change that occurred in the pre-oxidation process was that the SiC
powders oxidized in air at high temperatures generating
amorphous SiO2, which crystallized into cristobalite phases during
the cooling process as Reactions (1) and (2):
 +() + 
()()
The surface morphology of the raw SiC powders and that after
pre-oxidation were shown in Fig. 2. The original SiC powders
appeared as irregular shapes with relatively smooth surfaces as
shown in Fig. 2(a). After pre-oxidation, some tiny particles were
observed on the surface of the SiC particles, which might be due
to the chalking effect during pre-oxidation or ball milling
(Figs. 2(b)–2(d)). The surface of the SiC powders became rougher
with increasing the pre-oxidation time.
The element distribution of the SiC raw powders before
oxidation is shown in Figs. 3(a)–3(d), where C and Si were the
main elements on the surface. After the pre-oxidation treatment
(Figs. 3(e)–3(h)), besides C and Si, O was also detected on the
surface of the particles. The distribution of elements O and Si was
consistent with the profile of particles. Combined with the XRD
result in Fig. 1, it could be indicated that a scaly cristobalite layer
formed on the surface of the SiC powders after pre-oxidation
treatment.
The TEM image of pre-oxidized SiC powders is shown in
Fig. 4. A thin layer of SiO2 was found on the surface of the SiC
crystal, and its thickness was about 50–100 nm. The oxide shell
layer was non-uniform, which might be caused by the ball milling
process. Combined with the XRD results, the oxide layer was
cristobalite phase.
As particle size has a great effect on the viscosity of slurries [24],
Fig. 1  XRD patterns of raw SiC powders and those after pre-oxidation at 1200
for different time: (a) overall patterns; (b) local magnification of diffraction peak
at 21.7 ° in (a).
Fig. 2  SEM images of SiC powders before and after oxidation: (a) raw SiC
powders before oxidation; SiC powders oxidized at 1200 for (b) 2 h, (c) 4 h,
and (d) 6 h.
Fig. 3  SEM images and element distribution of SiC powders before and after oxidation: (a) raw SiC powders before oxidation; (e) SiC powders oxidized at 1200 for
4 h; (b–d) EDS mapping for (a) and (f–h) for (e).
High strength mullite-bond SiC porous ceramics fabricated by digital light processing 55
https://doi.org/10.26599/JAC.2024.9220835
the particle size distribution of SiC powders before and after
oxidation at 1200 for different time was measured and shown
in Fig. 5. Pre-oxidized powders were ball milled and then sieved. It
can be seen that, d50 of raw SiC powders was 5.288 μm, which
changed to 4.851, 5.047, and 5.346 μm after oxidation at 1200
for 2, 4, and 6 h. The variation was quite small. Therefore, it can
be concluded that the viscosity of slurry made from SiC powders
before and after oxidation is not strongly related to particle size
changes.
3.2PhotocuringabilityofSiC-basedslurry
With experimental verification, it was found that the solid loading
of the pure fine SiC powder-based slurry was difficult to reach
35 vol% without additives. That was because the dispersion
behavior of the SiC powders inside the photosensitive resin was
poor, and the sedimentation of the SiC powders occurred quickly.
As a result, the SiC powder-based slurry without additives was not
suitable for printing. Herein, the raw SiC powders and pre-
oxidized ones were mixed with the sintering additives and added
to commercial resin to prepare the photocurable SiC slurry with a
solid content of 45 vol%. The curing ability of the slurry was now
discussed.
The curing layer thickness of the SiC ceramic slurry varied with
the exposure time of UV light as shown in Fig. 6. For the slurry
fabricated from raw SiC powders, no obvious curing/
polymerization occurred until the UV light exposure time was
above 20 s. With the increase in the exposure time, the curing
layer thickness increased. When the exposure time reached 45 s,
the curing layer thickness was around 46 μm. For the slurry with
pre-oxidized SiC powders, the curing performance was greatly
Fig. 4  TEM image of surface of SiC powders after oxidation: (a) TEM image of SiC powders; (b) high magnification image of circle region in (a); (c) inverse fast Fourier
transform pattern of area in (b); (d) inverse fast Fourier transform pattern of area in (b).
Raw materials
d10=1.820
d50=5.288
d90=8.656
1200 °C/2 h
d10=0.979
d50=4.851
d90=7.495
1200 °C/4 h
d10=0.899
d50=5.047
d90=8.849
1200 °C/6 h
d10=2.136
d50=5.346
d90=8.580
Fig. 5  Particle size distribution curve of SiC powders before and after oxidation: (a) raw SiC powdesr before oxidation; SiC powders oxidized at 1200 for (b) 2 h,
(c) 4 h, and (d) 6 h.
56 J. Sun, J. Zhang, X. Zhang, et al.
J Adv Ceram2024,13(1):53−62
improved. Much shorter exposure time was required to obtain the
same curing thickness compared to the raw SiC powder-based
slurry. Among them, the ceramic slurry with the SiC powders
treated at 1200 for 4 h achieved the highest curing thickness at
all the range of the tested exposure time. To be specific, this slurry
exhibited a curing thickness as high as 51 μm with exposure time
of only 5 s, which was 40 s shorter than that of the control group.
As a result, the improved curing layer thickness can fully meet the
requirement for photocuring 3D printing, and therefore the
printing efficiency. According to the existing reports, the slurry
with un-treated SiC powders exhibited a curing thickness of < 30 μm
with exposure time of 90 s [24], whose exposure time is twice
while the curing thickness is thinner than that of the raw SiC
powder-based slurry in this work. In Ref. [27], 4.32 μm-sized SiC
powders were pre-oxidized at 1300 , and the fabricated slurry
showed a curing thickness of 59 μm with exposure time of 5 s,
which was comparable with that of the treated SiC powder-based
slurry in this work.
With the extension of the oxidation time, the thickness of the
oxide layer on the SiC surface would increase, and the curing
ability of the ceramic slurry would be improved. This was
consistent with other reports [27].
The influence of the pre-oxidation time on curing properties of
the ceramic slurry exhibited a non-linear trend, i.e., at the same
exposure time, the curing layer thickness first increased with the
pre-oxidation time increasing from 2 to 4 h, and decreased with
the pre-oxidation time further increasing to 6 h. To explore the
influencing factors, the viscosity of the four ceramic slurry at the
solid loading of 45 vol% was tested as shown in Fig. 7(a). It can be
clearly seen that the pre-oxidation treatment effectively reduced
the viscosity of the SiC slurry. Among them, the slurry with the
1200 /4 h treated SiC powders showed the highest viscosity in
the whole testing range of the shear rate. The viscosity values at
the shear rate of 30 s−1 are shown in Fig. 7(b). For the raw SiC
powder-based slurry, the viscosity was 2.543 Pa·s. For the slurry
made from the powders pre-oxidized at 1200 for 2, 4, and 6 h,
the viscosity decreased to 0.997, 1.369, and 0.859 Pa·s, respectively.
Fig. 6  Curve between curing thickness of SiC ceramic slurry with exposure time
for raw SiC powders and pre-oxidized SiC powders at 1200 for different
time.
Fig. 7  Viscosity of photosensitive SiC ceramic slurries: (a) viscosity at different shear rates and (b) viscosity at shear rate of 30 s1 with different oxidation time and fixed
solid content of 45 vol%; (c) viscosity at different shear rates and (d) the viscosity at shear rate of 30 s1 of slurry with different solid loadings prepared with SiC powders
pre-oxidized at 1200 for 4 h.
High strength mullite-bond SiC porous ceramics fabricated by digital light processing 57
https://doi.org/10.26599/JAC.2024.9220835
Effects of the solid loading on the viscosity of the ceramic slurry
fabricated from 4 h pre-oxidized SiC were further studied, and the
results are shown in Figs. 7(c) and 7(d). When the solid loading
increased from 45 to 50 vol%, the viscosity of the slurry slightly
increased from 1.369 to 3.112 Pa·s. With the solid loading further
increasing to 55 vol%, the viscosity of the slurry was still less than
10 Pa·s at a shear rate of 30 s−1, which is a criterion to judge
whether the slurry is suitable for printing [32]. To date, The
viscosity of the slurry with a pre-oxidized SiC particle size of
10 μm exceeded 10 Pa·s at a shear rate of 30 s−1 in a 55 vol% solid
content [26]. However, there have been few reports of the SiC
slurry with high solid loadings and low viscosity using powders
with a particle size of around 5 μm.
To investigate the effect of sintering additives on the viscosity,
three slurry were prepared separately as shown in Fig. 8. The
slurry A contained only pre-oxidized SiC powders. Based on this,
Al(OH)3 was added to form the slurry B. Also, in the slurry C,
along with pre-oxidized SiC powders, all additives, including
Al(OH)3, Y2O3, and CaF2 were added. When the solid load of the
slurry A reached 45 vol%, it showed a paste-like consistency with a
complete loss of flowability. Poor flowability was observed when
the solid content was 40 vol%. To facilitate the test, the solid
content of the slurry A was decreased to 35 vol%. The solid
content of 45 vol% was fabricated for the slurry B and C to
maintain the consistency with the above result.
As shown in Fig. 8(a), a strong shear-thinning behavior was
observed for the slurry A. The viscosity of both slurry B and C
almost became constant when the shearing rate was higher than
20 s−1. The viscosity of these slurry at a shear rate of 30 s−1 was
shown in Fig. 8(b). The visocisity of the slurry A was 1.375 Pa·s,
which changed to 2.232 and 1.369 Pa·s with the addition of
Al(OH)3 and Al(OH)3+Y2O3+CaF2, respectively. Considering the
lower solid content of the slurry A, we can conclude that the
addition of additives has a small effect on their viscosity.
Considering the best curing ability of the slurry made from the
1200 /4 h treated SiC powders and also the acceptable viscosity
of the slurry at a solid loading of 50 vol%, this slurry was used to
print SiC bodies in the following work. To validate the feasibility
of using the slurry to print complex structures, a range of models
were designed, and printed, and sintered, as shown in Fig. 9.
Complex shapes were successfully maintained after sintering. The
result proved that the combination of pre-oxidation with the
addition of sintering additives was an effective way to enhance the
curing ability of the SiC slurry. The line shrinkage was similar for
the x- and y-axes in the horizontal direction with an average
shrinkage of 12.5%, and 22.7% for the z-axis along the print
direction. The total volume shrinkage was 41%.
The roughness of the top surface and side surface along the
print direction of the green body were measured and shown in
Fig. 10. The average surface roughness (Ra) of these two surfaces
was less than 1 μm, indicating the smooth feature of printing
surfaces.
The surface microstructure of the green body is shown in
Fig. 11. The top surface of the green body is distributed with
particles, and no obvious defects such as bumps and dents were
observed (Figs. 11(a) and 11(b)). In the side surface along the
printing direction, as shown in Figs. 11(c) and 11(d), printing
layers were observed without any other visible defects. The locally
enlarged image in Fig. 11(d) demonstrated that the combination
between layers was good, and there was no obvious layering
boundary and no debonding.
The top surface microstructure of the sintered body is shown in
Figs. 12(a) and 12(b). There were many pore structures distributed
on the top surface, and obvious rod-like grains were observed
around the pores. The particles were bonded together through the
glass phase, and no obvious crack defects were observed. On the
side surface along the printing direction, as shown in Figs. 12(c)
and 12(d), wave-like features were observed inheriting the layer-by-
layer printing characteristic. No interlayer debonding was noticed
from the locally enlarged image in Fig. 12(d).
3.3PropertiesofprintedSiCceramics
To avoid the negative impacts of the cristobalite phase generated
by pre-oxidation and oxidative debinding/sintering on the
performance of the SiC ceramics, Al(OH)3 was added as the Al2O3
source to react with SiO2 and form mullite, as illustrated in
Reactions (3) and (4):
()+ 
+·()
In addition, Y2O3 and CaF2 were added to enhance the
formation of mullite and the length/diameter ratio of the formed
mullite reinforcement. Based on the dry pressing process in our
previous work, the optimal amount of Al(OH)3 was 6.4 wt% of
the weight of SiC powders. However, the debinding treatment in
air was required for ceramics printed by the DLP method due to a
large amount of addition of organic matter such as photosensitive
resin and binders, which prolonged the entire oxidation process.
Besides, the SiC powders were also pre-oxidized before making
Fig. 8  Viscosity of photosensitive SiC ceramic slurry: (a) viscosity at different shear rates and (b) viscosity at shear rate of 30 s−1 with different compositions.
58 J. Sun, J. Zhang, X. Zhang, et al.
J Adv Ceram2024,13(1):53−62
the slurry. Therefore, it was necessary to explore and adjust the
proportion of the sintering additives to completely convert the
cristobalite phase into the mullite phase.
In this paper, four different amounts of Al(OH)3, i.e., 6.4, 10,
12.5, and 15 wt%, were added into samples, and their phase
composition after sintering was shown in Fig. 13. The crystalline
phase of the sintered samples included 6H–SiC, cristobalite, and
mullite as shown in Fig. 13(a). With the increase of the Al(OH)3
amount, the diffraction peak of the mullite phase was enhanced
and reached the highest when Al(OH)3 was 15 wt%, indicating
that some cristobalite continued to react with the Al2O3 source.
The peak intensity of the cristobalite phase first increased with the
Al(OH)3 amount increasing from 6.4 to 10 wt%, and then
disappeared at the Al(OH)3 amount of 12.5 wt%. Therefore, the
cristobalite produced in the whole process was completely
consumed at the Al(OH)3 amount of 12.5 wt%.
Interestingly, at the amount of Al(OH)3 to 15 wt%, a tiny
diffraction peak of the cristobalite phase reappeared. Since the
cristobalite phase has been eliminated at the Al(OH)3 addition
amount of 12.5 wt%, Al(OH)3 should be excessive. However, no
diffraction peaks of Al2O3 were detected in the XRD result,
indicating that the Al2O3 source had been consumed. Therefore, it
could be inferred that the excessive addition of Al(OH)3 might
accelerate the oxidation of SiC and therefore produce an extra
amount of the cristobalite phase [31,33], in which further
investigation is expected in the future.
Effects of the sintering temperature on the phase composition
of the SiC ceramic materials with 12.5 wt% Al(OH)3 addition were
shown in Fig. 13(b). With the increase of the sintering
temperature, the intensity of the diffraction peak of the cristobalite
phase decreased gradually, indicating that the higher sintering
temperature enhanced the entire mullitization process.
Theoretically, at a lower sintering temperature of 1450 , the
content of cristobalite produced by oxidation during sintering
should be less than that at 1550 , since a higher temperature
induces severer oxidation of SiC. However, as shown in Fig. 13(b),
a larger amount of the cristobalite phase was observed for the SiC
ceramic materials sintered at 1450 compared to that sintered at
1550 . This was mainly because the sintering temperature of
1450 was not high enough to enable the mullite formation
reaction to completely conduct. As a result, increasing the
sintering temperature enhanced the formation reaction of mullite
and therefore the elimination of the cristobalite phase. As shown
in Fig. 13(b), with the increase of the sintering temperature, the
intensity of the diffraction peaks of mullite increased while that of
the cristobalite phase decreased. The cristobalite phase was out of
detection when the sintering temperature reached 1550 . In
summary, the optimal addition amount of Al(OH)3 and sintering
temperature were found to be 12.5 wt% and 1550 , respectively,
at which, oxidation-derived cristobalite was completely converted
into mullite phase, realizing mullite-bonded SiC ceramic
fabrication.
The microstructure and element distribution at the cross
section of the sintered SiC ceramic materials are shown in Fig. 14.
Although they were porous, the particles were bonded together,
Fig. 9  Photograph of complex structural SiC (a) green bodies and (b) sintered
bodies printed by DLP 3D printing.
Fig. 10  Surface roughness of green body: (a) 2D and (b) 3D image of top surface; (c) 2D and (d) 3D image of side surface along print direction where Ra is the average
roughness, Rq is the root-mean-square deviation, and Rmax is the maximum roughness depth.
High strength mullite-bond SiC porous ceramics fabricated by digital light processing 59
https://doi.org/10.26599/JAC.2024.9220835
Fig. 11  SEM images of printed green body: (a, b) top surface and (c, d) side surface in z-axis printing direction.
Fig. 12  SEM images of sintered ceramic body: (a, b) top surface and (c, d) side surface in z-axis printing direction.
Fig. 13  XRD patterns of SiC ceramics with different Al(OH)3 addition amounts and sintering temperatures: (a) different Al(OH)3 addition amounts sintered at
1550 /2 h; (b) different sintering temperatures with 12.5 wt% Al(OH)3 addition.
60 J. Sun, J. Zhang, X. Zhang, et al.
J Adv Ceram2024,13(1):53−62
and no loose individual SiC particles were observed. According to
the EDS mapping result, Si, O, and Al elements were distributed
uniformly. This suggested that the mullite phase was formed as a
bonding phase between SiC particles making the materials well
sintered. The low sintering temperature of 1550 in air can also
effectively reduce the fabrication cost of complex SiC components
compared to the traditional high temperature/pressure and
vacuum/atmosphere sintering processes.
The flexural strength and apparent porosity of the SiC ceramic
materials sintered at 1550 in air by the DLP process were
shown in Fig. 15. The average flexural strength reached 97.6 MPa,
and the apparent porosity was 39.2 vol%. For comparison, the
performances of the SiC ceramic materials fabricated by other
techniques in literature, including the flexural strength and
porosity, were collected and listed in Table 1. The flexure strength
achieved in this study was 94.8% higher than that sintered at 2000
(50.1 MPa) in an Ar atmosphere by the DLP process, whose
porosity information remained unclear [27]. In addition, the
excellent flexural strength achieved by the DLP process in this
work was also comparable to that fabricated by other techniques,
e.g., SPS [34] and dry pressing method in our previous work [31]
with similar porosity, and much higher than the SiC ceramic
materials fabricated by dry pressing and sintering from other
groups.
The flexural strength of SiC porous ceramics prepared in this
study was high compared to the sintered bodies with similar
apparent porosity, but high porosity was a limitation to the
ceramics’ properties. The density of sintered ceramics could be
improved by subsequent post-treatment processes, such as PIP,
CVI, and RMI techniques, to enhance their properties [8,3538].
Ding et al. [8] and Chen et al. [37] greatly enhanced the density
and strength of SiC and C/SiC by PIP post-treatment [8,37].
Other researchers combined PIP with CVI [36,38] or CVI with
RMI [35], and the properties of SiC-based composites were greatly
enhanced. These treatments provided practical ideas for our
future research.
4Conclusions
In this paper, complex-shaped SiC ceramic materials with high
flexural strength were successfully fabricated by the DLP
technique and low-temperature and air-atmosphere sintering.
With the pre-oxidation treatment, SiO2 shell layers formed on the
SiC powders. As a result, the printing ability of the SiC
photosensitive slurry was effectively enhanced even though the
SiC particle size was as fine as 5 μm: a high curing thickness was
achieved above 50 μm with exposure time as short as 5 s; the
viscosity of the SiC slurry with a solid loading of 55 vol% was as
low as 7.77 Pa·s at a shear rate of 30 s−1. Moreover, with the
rational design of the sintering additives, the oxidation-derived
cristobalite in the materials was completely consumed and
Fig. 14  Micromorphology and element distribution of cross section of sintered SiC ceramic materials.
Fig. 15  Flexural strength and apparent porosity of SiC ceramic materials
sintered at 1550 in air by DLP technique.
Table 1  Comparison of apparent porosity and flexural strength of SiC porous ceramics prepared by photocuring and traditional methods
Material; additive Process Sintering condition Flexural strength (MPa) Apparent porosity (vol%) Ref.
SiC; Al(OH)3+Y2O3+CaF2DLP Pressureless, 1550 , air 97.6 39.2 This work
SiC DLP Pressureless, 2000 , argon 47.9 [27]
SiC; Y2O3+Al2O3SLA Pressureless, 1950 , nitrogen 229.0 10.2 [39]
SiC SPS, 1800 , vacuum 103.0 35.7 [34]
SiC; MoO3+Al2O3Dry pressing Pressureless, 1000 , air 66.0 45.4 [40]
SiC; Al(OH)3+Y2O3+CaF2Dry pressing Pressureless, 1550 , air 113.0 40.3 [31]
SiC; fly ash+MoO3Dry pressing Pressureless, 1000 , air 38.4 36.4 [41]
SiC; Al2O3+graphite Dry pressing Pressureless, 1450 , air 27.5 44.4 [42]
High strength mullite-bond SiC porous ceramics fabricated by digital light processing 61
https://doi.org/10.26599/JAC.2024.9220835
converted into a higher-strength mullite bonding phase. The thus
fabricated SiC ceramic materials exhibited flexural strength of
97.6 MPa at apparent porosity of 39.2 vol%, which is comparable
to that fabricated by the SPS process. This work therefore provides
a novel method to fabricate DLP-process SiC ceramic materials
with excellent printing ability and high mechanical strength.
Acknowledgements
This work was supported by Shandong University−MSEA
International Institute for Materials Genome Joint Innovation
Center for Advanced Ceramics, and the Key Research and
Development Projects of Shaanxi Province (Nos. 2018ZDCXL-
GY-09-06 and 2021ZDLGY14-06).
Declarationofcompetinginterest
The authors have no competing interests to declare that are
relevant to the content of this article.
References
Naslain R, Guette A, Rebillat F, et al. Boron-bearing species in
ceramic matrix composites for long-term aerospace applications.
J Solid State Chem 2004, 177: 449–456.
[1]
Zhang JX, Huang R, Gu H, et al. High toughness in laminated SiC
ceramics from aqueous tape casting. Sciptar Mater 2005, 52: 381–385.
[2]
Yuan SW, Yang ZC, Chen GB. 3D microstructure model and ther-
mal shock failure mechanism of a Si3N4-bonded SiC ceramic refrac-
tory with SiC high volume ratio particles. Ceram Int 2019, 45:
4219–4229.
[3]
Liu JX, Tian C, Xiao HN, et al. Effect of B4C on co-sintering of SiC
ceramic membrane. Ceram Int 2019, 45: 3921–3929.
[4]
Duan WY, Yin XW, Li Q, et al. A review of absorption properties in
silicon-based polymer derived ceramics. J Eur Ceram Soc 2016, 36:
3681–3689.
[5]
Katoh Y, Snead LL, Henager CH Jr, et al. Current status and recent
research achievements in SiC/SiC composites. J Nucl Mater 2014,
455: 387–397.
[6]
Krenkel W, Berndt F. C/C–SiC composites for space applications
and advanced friction systems. Mater Sci Eng A 2005, 412: 177–181.
[7]
Ding GJ, He RJ, Zhang KQ, et al. Stereolithography 3D printing of
SiC ceramic with potential for lightweight optical mirror. Ceram Int
2020, 46: 18785–18790.
[8]
Katoh Y, Snead LL, Szlufarska I, et al. Radiation effects in SiC for
nuclear structural applications. Curr Opin Solid State Mater Sci 2012,
16: 143–152.
[9]
Padture NP. Advanced structural ceramics in aerospace propulsion.
Nat Mater 2016, 15: 804–809.
[10]
Rosso M. Ceramic and metal matrix composites: Routes and proper-
ties. J Mater Process Technol 2006, 175: 364–375.
[11]
Han DY, Mei H, Xiao SS, et al. A review on the processing technolo-
gies of carbon nanotube/silicon carbide composites. J Eur Ceram Soc
2018, 38: 3695–3708.
[12]
Chen Z, Sun XH, Shang YP, et al. Dense ceramics with complex
shape fabricated by 3D printing: A review. J Adv Ceram 2021, 10:
195–218.
[13]
Zhang F, Li ZA, Xu MJ, et al. A review of 3D printed porous ceram-
ics. J Eur Ceram Soc 2022, 42: 3351–3373.
[14]
Fu YL, Chen ZW, Xu G, et al. Preparation and stereolithography 3D
printing of ultralight and ultrastrong ZrOC porous ceramics. J Alloys
Compd 2019, 789: 867–873.
[15]
Hassanin H, Essa K, Elshaer A, et al. Micro-fabrication of ceramics:
Additive manufacturing and conventional technologies. J Adv Ceram
2021, 10: 1–27.
[16]
He RJ, Zhou NP, Zhang KQ, et al. Progress and challenges towards
additive manufacturing of SiC ceramic. J Adv Ceram 2021, 10:
637–674.
[17]
Liu G, Zhang XF, Chen XL, et al. Additive manufacturing of struc-
tural materials. Mater Sci Eng R Rep 2021, 145: 100596.
[18]
Rasaki SA, Xiong DY, Xiong SF, et al. Photopolymerization-based[19]
additive manufacturing of ceramics: A systematic review. J Adv
Ceram 2021, 10: 442–471.
Xing HY, Zou B, Lai QG, et al. Preparation and characterization of
UV curable Al2O3 suspensions applying for stereolithography 3D
printing ceramic microcomponent. Powder Technol 2018, 338:
153–161.
[20]
An D, Li HZ, Xie ZP, et al. Additive manufacturing and characteriza-
tion of complex Al2O3 parts based on a novel stereolithography
method. Int J Appl Ceram Technol 2017, 14: 836–844.
[21]
Nie GL, Li YH, Sheng PF, et al. Microstructure refinement-homoge-
nization and flexural strength improvement of Al2O3 ceramics fabri-
cated by DLP-stereolithography integrated with chemical precipita-
tion coating process. J Adv Ceram 2021, 10: 790–808.
[22]
Wu HD, Liu W, He RX, et al. Fabrication of dense zirconia-tough-
ened alumina ceramics through a stereolithography-based additive
manufacturing. Ceram Int 2017, 43: 968–972.
[23]
Ding GJ, He RJ, Zhang KQ, et al. Stereolithography-based additive
manufacturing of gray-colored SiC ceramic green body. J Am Ceram
Soc 2019, 102: 7198–7209.
[24]
Tang J, Guo XT, Chang HT, et al. The preparation of SiC ceramic
photosensitive slurry for rapid stereolithography. J Eur Ceram Soc
2021, 41: 7516–7524.
[25]
Cao JW, Miao K, Xiong SF, et al. 3D printing and in situ transforma-
tion of SiCnw/SiC structures. Addit Manuf 2022, 58: 103053.
[26]
Shi ZA, Wu JM, Fang ZQ, et al. Influence of high-temperature oxida-
tion of SiC powders on curing properties of SiC slurry for digital
light processing. J Adv Ceram 2023, 12: 169–181.
[27]
Brinckmann SA, Patra N, Yao J, et al. Stereolithography of SiOC
polymer-derived ceramics filled with SiC micronwhiskers. Adv Eng
Mater 2018, 20: 1800593.
[28]
He RJ, Ding GJ, Zhang KQ, et al. Fabrication of SiC ceramic archi-
tectures using stereolithography combined with precursor infiltra-
tion and pyrolysis. Ceram Int 2019, 45: 14006–14014.
[29]
Talwar DN, Sherbondy JC. Thermal expansion coefficient of 3C–SiC.
Appl Phys Lett 1995, 67: 3301–3303.
[30]
Li ZH, Chang ZX, Liu XR, et al. A novel sintering additive system for
porous mullite-bonded SiC ceramics: High mechanical performance
with controllable pore structure. Ceram Int 2022, 48: 4105–4114.
[31]
Zhang KQ, Xie C, Wang G, et al. High solid loading, low viscosity
photosensitive Al2O3 slurry for stereolithography based additive
manufacturing. Ceram Int 2019, 45: 203–208.
[32]
Ebrahimpour O, Dubois C, Chaouki J. Manufacturing process for in
situ reaction-bonded porous SiC ceramics using a combination of
graft polymerization and sol–gel approaches. Ind Eng Chem Res 2014,
53: 17604–17614.
[33]
Zhao PQ, Li QG, Yi RJ, et al. Fabrication and microstructure of liq-
uid sintered porous SiC ceramics through spark plasma sintering.
J Alloys Compd 2018, 748: 36–43.
[34]
Li JX, Liu YS, Nan BY, et al. Microstructure and properties of C/
SiC–diamond composites prepared by the combination of CVI and
RMI. Adv Eng Mater 2019, 21: 1800765.
[35]
Liu RJ, Wang F, Zhang JP, et al. Effects of CVI SiC amount and
deposition rates on properties of SiCf/SiC composites fabricated by
hybrid chemical vapor infiltration (CVI) and precursor infiltration
and pyrolysis (PIP) routes. Ceram Int 2021, 47: 26971–26977.
[36]
Chen YF, Zhang L, Zhao YN, et al. Mechanical behaviors of C/SiC
pyramidal lattice core sandwich panel under in-plane compression.
Compos Struct 2019, 214: 103–113.
[37]
Ortona A, Donato A, Filacchioni G, et al. SiC–SiCf CMC manufac-
turing by hybrid CVI–PIP techniques: Process optimisation. Fusion
Eng Des 2000, 51–52: 159–163.
[38]
Wang KJ, Liu RZ, Bao CG. SiC paste with high curing thickness for
stereolithography. Ceram Int 2022, 48: 28692–28703.
[39]
Xing ZH, Hu YH, Xiang DP, et al. Porous SiC−mullite ceramics with
high flexural strength and gas permeability prepared from photo-
voltaic silicon waste. Ceram Int 2020, 46: 1236–1242.
[40]
Das D, Nijhuma K, Gabriel AM, et al. Recycling of coal fly ash for
fabrication of elongated mullite rod bonded porous SiC ceramic
membrane and its application in filtration. J Eur Ceram Soc 2020, 40:
2163–2172.
[41]
Ding SQ, Zhu SM, Zeng YP, et al. Effect of Y2O3 addition on the
properties of reaction-bonded porous SiC ceramics. Ceram Int 2006,
32: 461–466.
[42]
62 J. Sun, J. Zhang, X. Zhang, et al.
J Adv Ceram2024,13(1):53−62
... Zhang et al. [127,128] prepared Pickering emulsion with varied oil to water ratios as the feedstock for DLP, where suitable rheological performance with shear-thinning behavior and elastic solid at low shear stress, as well as excellent stability free from phase separation and emulsion droplet change have been found for DLP process (Fig. 1 ). enhanced curing thickness over 50 m [129]. However, the synthesis of photosensitive organosilicon polymers, ceramic yield as well as higher sintering shrinkage (usually higher than 25%) are the main concerns for PDC. ...
... (a) Dynamic interfacial tension of suspensions composed of water, HDDA and mixed particles before and after adding modifier; (b) photographs of oil-based suspension and Pickering emulsions with different oil to water ratios after standing for two weeks; (c, d) the rheological properties of oil-based suspension and Pickering emulsions with different oil to water ratios, (c) is the viscosity as a function of shear rate and (d) is the modulus varies with shear stress; (e) diagram displaying the water amount and solid loading to obtain pastes with printable rheological properties and stability. Reproduced with permission from Ref. [127], © Elsevier B.V. 2023.Besides, ceramic particles with high refractive index and light absorption, involving Si3N4 and SiC, bring great challenges for preparing complex shape parts by DLP, which has been well solved by surface modification and the emerging polymerderived ceramics (PDC)[129][130][131]. To realize the DLP for SiC particles, pre-oxidation treatment for SiC has been conducted by Han et al. and suspension with high solid content of 55 vol.% and viscosity of 7.77 Pa·s was prepared, contributing to the ...
... Mullite-bonded SiC (MBS) porous ceramics, a type of reaction-bonding SiC ceramics materials with an in-situ mullite layer forming on the SiC particle surface to bond SiC particles together, is considered an excellent candidate material for hot gas infiltration due to a range of advantages e.g. low sintering temperature [3], sustainabilityusing waste as raw materials e.g. photovoltaic silicon waste [4] and coal fly ash [5], good properties including intrinsic porous structure, corrosion resistance and thermal shock resistance [6,7]. ...
Article
Mullite-bonded SiC porous ceramics are of potential in hot gas infiltration while the insufficient mechanical performance limits its application. In-situ mullite whisker reinforcement is a potential solution but the formation temperature mismatch between mullite whisker and bonding layer is a challenge. In that case, mullite whiskers massively form ahead of the bonding layer excessively consuming alumina leading to poor bonding layers between SiC particles and thus, the mechanical performance. Herein, the mismatch was solved by using CaF2 as a temperature-controlled reaction switch. As a result, both the bonding layer and whisker synchronously formed. The material showed significant reinforcement effects, and exhibited a flexural strength of 124.8 MPa and a fracture toughness of 1.53 MPa·m1/2 with an open porosity of 36.6 vol%, realizing a breakthrough in the mechanical performance. Interestingly, an increase in oxygen vacancy concentration with the sintering temperature was first reported, which may benefit the whisker growth.
Article
Full-text available
Barium titanate (BaTiO3) piezoelectric ceramics with triply periodic minimal surface (TPMS) structures have been frequently used in filters, engines, artificial bones, and other fields due to their high specific surface area, high thermal stability, and good heat dissipation. However, only a limited number of studies have analyzed the effect of various parameters, such as different wall thicknesses and porosities of TPMS structures, on ceramic electromechanical performance. In this study, we first employed vat photopolymerization (VPP) three-dimensional (3D) printing technology to fabricate high-performance BaTiO3 ceramics. We investigated the slurry composition design and forming process and designed a stepwise sintering postprocessing technique to achieve a density of 96.3% and a compressive strength of 250±25 MPa, with the piezoelectric coefficient (d33) reaching 263 pC/N. Subsequently, we explored the influence of three TPMS structures, namely, diamond, gyroid, and Schwarz P, on the piezoelectric and mechanical properties of BaTiO3 ceramics, with the gyroid structure identified as exhibiting optimal performance. Finally, we examined the piezoelectric and mechanical properties of BaTiO3 ceramics with the gyroid structure of varying wall thicknesses and porosities, thus enabling the modulation of ceramic electromechanical performance.
Article
Full-text available
Fabrication of silicon carbide (SiC) ceramics by digital light processing (DLP) technology is difficult owing to high refractive index and high ultraviolet (UV) absorptivity of SiC powders. The surface of the SiC powders can be coated with silicon oxide (SiO2) with low refractive index and low UV absorptivity via high-temperature oxidation, reducing the loss of UV energy in the DLP process and realizing the DLP preparation of the SiC ceramics. However, it is necessary to explore a high-temperature modification process to obtain a better modification effect of the SiC powders. Therefore, the high-temperature modification behavior of the SiC powders is thoroughly investigated in this paper. The results show that nano-scale oxide film is formed on the surface of the SiC powders by short-time high-temperature oxidation, effectively reducing the UV absorptivity and the surface refractive index (nʹ) of the SiC powders. When the oxidation temperature is 1300 ℃, compared with that of unoxidized SiC powders, the UV absorptivity of oxidized SiC powders decreases from 0.5065 to 0.4654, and a curing depth of SiC slurry increases from 22±4 to 59±4 μm. Finally, SiC green bodies are successfully prepared by the DLP with the the oxidized powders, and flexural strength of SiC sintered parts reaches 47.9±2.3 MPa after 3 h of atmospheric sintering at 2000 ℃ without any sintering aid.
Article
Full-text available
High strength and thermal shock resistant mullite-bonded porous SiC ceramics are essential for applications in catalyst, hot gas/solid filtration/separation fields. However, these properties were limited by the contradictory issue: higher temperature is necessary for the formation and densification of mullite bond phase, while lower temperature is preferred to avoid sever oxidation of SiC accompanied with detrimental cristobalite phase. Herein, a novel composite additive system, Al(OH)3+Y2O3+CaF2, was proposed to solve this problem for one shot. Flexural strength as high as 113 MPa (40.3 vol% porosity) was achieved for samples sintered in air, even higher than that fabricated by expensive SPS technique (103 MPa, 35.7 vol%). Critical quenching temperature reached 547 °C, around 227 °C improved. Meanwhile, additives acted as pore formers and small variation of them could effectively modify the size and shape of pores. The evolution of pore structure with additives was thoroughly evaluated and models were established for the first time. Correlations of phase composition, microstructure, porosity, pore size and shape to the performance of porous SiC were investigated. This work provides a flexible strategy to simultaneously control the mechanical performance and pore structure for low cost porous SiC ceramics.
Article
Full-text available
Silicon carbide (SiC) ceramic and related materials are widely used in various military and engineering fields. The emergence of additive manufacturing (AM) technologies provides a new approach for the fabrication of SiC ceramic products. This article systematically reviews the additive manufacturing technologies of SiC ceramic developed in recent years, including Indirect Additive Manufacturing (Indirect AM) and Direct Additive Manufacturing (Direct AM) technologies. This review also summarizes the key scientific and technological challenges for the additive manufacturing of SiC ceramic, and also forecasts its possible future opportunities. This paper aims to provide a helpful guidance for the additive manufacturing of SiC ceramic and other structural ceramics.
Article
Full-text available
In this study, the chemical precipitation coating (CP) process was creatively integrated with DLP-stereolithography based 3D printing for refining and homogenizing the microstructure of 3D printed Al2O3 ceramic. Based on this novel approach, Al2O3 powder was coated with a homogeneous layer of amorphous Y2O3, with the coated Al2O3 powder found to make the microstructure of 3D printed Al2O3 ceramic more uniform and refined, as compared with the conventional mechanical mixing (MM) of Al2O3 and Y2O3 powders. The grain size of Al2O3 in Sample CP is 64.44% and 51.43% lower than those in the monolithic Al2O3 ceramic and Sample MM, respectively. Sample CP has the highest flexural strength of 455.37±32.17 MPa, which is 14.85% and 25.45% higher than those of Samples MM and AL, respectively; also Sample CP has the highest Weibull modulus of 16.88 among the three kinds of samples. Moreover, the fine grained Sample CP has a close thermal conductivity to the coarse grained Sample MM because of the changes in morphology of Y3Al5O12 phase from semi-connected (Sample MM) to isolated (Sample CP). Finally, specially designed fin-type Al2O3 ceramic heat sinks were successfully fabricated via the novel integrated process, which has been proven to be an effective method for fabricating complex-shaped Al2O3 ceramic components with enhanced flexural strength and reliability.
Article
Full-text available
The SiCf/SiC composites have been manufactured by a hybrid route combining chemical vapor infiltration (CVI) and precursor infiltration and pyrolysis (PIP) techniques. A relatively low deposition rate of CVI SiC matrix is favored ascribing to that its rapid deposition tends to cause a ‘surface sealing’ effect, which generates plenty of closed pores and severely damages the microstructural homogeneity of final composites. For a given fiber preform, there exists an optimized value of CVI SiC matrix to be introduced, at which the flexural strength of resultant composites reaches a peak value, which is almost twice of that for composites manufactured from the single PIP or CVI route. Further, this optimized CVI SiC amount is unveiled to be determined by a critical thickness t0, which relates to the average fiber distance in fiber preforms. While the deposited SiC thickness on fibers exceeds t0, closed pores will be generated, hence damaging the microstructural homogeneity of final composites. By applying an optimized CVI SiC deposition rate and amount, the prepared SiCf/SiC composites exhibit increased densities, reduced porosity, superior mechanical properties, increased microstructural homogeneity and thus reduced mechanical property deviations, suggesting a hybrid CVI and PIP route is a promising technique to manufacture SiCf/SiC composites for industrial applications.
Article
Full-text available
Additive manufacturing (AM), also known as three-dimensional (3D) printing, has boomed over the last 30 years, and its use has accelerated during the last 5 years. AM is a materials-oriented manufacturing technology, and printing resolution versus printing scalability/speed trade-off exists among various types of materials, including polymers, metals, ceramics, glasses, and composite materials. Four-dimensional (4D) printing, together with versatile transformation systems, drives researchers to achieve and utilize high dimensional AM. Multiple per�spectives of the AM of structural materials have been raised and illustrated in this review, including multi�material AM (MMa-AM), multi-modulus AM (MMo-AM), multi-scale AM (MSc-AM), multi-system AM (MSy�AM), multi-dimensional AM (MD-AM), and multi-function AM (MF-AM). The rapid and tremendous development of AM materials and methods offers great potential for structural applications, such as in the aerospace field, the biomedical field, electronic devices, nuclear industry, flexible and wearable devices, soft sensors, actuators, and robotics, jewelry and art decorations, land transportation, underwater devices, and porous structures.
Article
Successful preparation of photopolymerizable ceramic slurries using non-oxide ceramics and their 3D printing is highly challenging as the ceramic powders often possess high UV-light absorbance. Even though pre-oxidation of the powders could improve the slurry printability, the residual oxygen element in the material would eventually deteriorate the material purity and performance. In this study, a separated redox route (pre-oxidation, 3D printing and carbothermal reduction) was proposed to overcome these challenges. SiC ceramic powder was pre-oxidized into [email protected]2 core-shell form to lower the UV-light absorbance. It was found that the ultraviolet (UV) light absorbance value of SiC powder at 405 nm decreased 20% from 0.357 to 0.284 after pre-oxidation process. The UV-light penetration depth of [email protected]2 ceramic slurries at 40 ~ 55 vol% solid loading ranging from 14.81 to 12.82 μm, which surpassed two times that for 40 vol% raw SiC ceramic slurry. Different-shaped structures of [email protected]2/resin green ceramic bodies were successfully fabricated by vat photopolymerization 3D printing. The oxygen element resided was then eliminated by carbothermal reduction to avoid detriment of SiO2 to the mechanical performance of the SiC ceramics. Phenolic epoxy acrylate resin with 14 wt% pyrolytic carbon (PyC) yield was chosen as the carbon source and thus the [email protected]2/resin green body pyrolyzed into [email protected]2/PyC ceramics after 1200 °C. It was found that after the heating temperature was further raised to 1600°C, most of the pre-introduced SiO2 shells on the surface of SiC particles were in situ transformed into SiC nanowires through carbothermal reduction reaction between SiO2 and PyC, and the content of oxygen element in the ceramic matrix sharply dropped from 20.21 to 2.08%. The results demonstrated that through the proposed redox route, the photopolymerization-related issues including high UV-light absorbance and pre-oxidation-induced impurity for 3D printing of SiC ceramic slurries can be tackled. Finally, high purity porous SiCnw/SiC ceramic components with different structures were produced. This study provides a promising route for the preparation and 3D printing of photopolymerization ceramic slurries using non-oxide ceramics possessing high UV-light absorbance.
Article
In this study, the method and mechanism of the improvement of the curing thickness of SiC by stereolithography (SLA) were investigated, the rheological properties of SiC paste were optimised, and the problem of cracking problem during debinding was solved. The best moulding process for preparing SiC ceramics with the Ceramaker 900 equipment was studied. Oxidation at 1180 °C for 1 h and hydroxylation with 0.6 mol/L H2O2 were the best modification processes for SiC powder. SiC paste with good rheological properties and high curing thickness was prepared using a resin system of BPA, HDDA, TPGDA, TMPTA, DPHA, and A-BPEF (mass ratio = 9:7:5:20:15:6). The best debinding process for SLA SiC was segmented debinding. The bulk density, total porosity, and room temperature bending strength of the SiC ceramics obtained are 2.13 g/cm³, 10.2%, 229 MPa, respectively. The nanoindentation elastic modulus and hardness of SLA SiC ceramics were found to be 18.7 and 1.66 GPa, respectively. Through this process, a lattice structure of SiC parts with good surface quality was prepared.
Article
Three-dimensional (3D) printing of ceramics has gained widespread attentions in recent years. Many excellent reviews have reported the printing of ceramics. However, most of them focus on printing of dense ceramics or general ceramic aspects, there is no systematical review about 3D printing of porous ceramics. In this review paper, the 3D printing technologies for fabricating of porous ceramic parts are introduced, including binder jetting, selective laser sintering, direct ink writing, stereolithography, laminated object manufacturing, and indirect 3D printing processes. The techniques to fabricate hierarchical porous ceramics by integrating 3D printing with one or more conventional porous ceramics fabrication approaches are reviewed. The main properties of porous ceramics such as pore size, porosity, and compressive strength are discussed. The emerging applications of 3D printed porous ceramics are presented with a focus on the booming application in bone tissue engineering. Finally, summary and a perspective on the future research directions for 3D printed porous ceramics are provided.
Article
Stereolithography is one of the most widely used additive manufacturing techniques for preparing high precision and complex ceramic components. Due to the high optical absorbance and refractive index of SiC powder, the rapid stereolithography of SiC ceramics components has become a key challenge. Here, we innovatively use graded silica to improve the curing thickness, rheological and settling performance of the slurry. And we presented a preparation method of SiC ceramic slurry for stereolithography with high solid content, low viscosity, low sedimentation rate and high curing thickness. The printable precision of the slurry is more than 75 μm, the dynamic viscosity is less than 2 Pa·s, and the 24 h sedimentation height is less than 5%. This strategy demonstrates a tantalizing possibility and promising prospect to rapid stereolithography of large size SiC ceramic green body.