You are on page 1of 4

Materials Chemistry and Physics 59 (1999) 139±142

Densi®cation behavior and mechanical properties of pressureless-sintered


silicon carbide ceramics with alumina and yttria additions
J.H. She*, K. Ueno
Department of Energy Conversion, Osaka National Research Institute, Midorigaoka 1-8-31, Ikeda, Osaka 563-8577, Japan

Received 23 November 1998; received in revised form 8 January 1999; accepted 12 January 1999

Abstract

SiC powders were pressureless sintered to about 98% of the theoretical density by using 10 wt.% (Al2O3 ‡ Y2O3) as sintering additives.
The densi®cation mechanism was attributed to liquid-phase sintering due to the formation of a eutectic liquid between Al2O3 and Y2O3 at
sintering temperatures. SEM observation revealed that SiC exhibited a uniformly distributed ®ne-grained microstructure and a highly
intergranular fracture behavior. The maximum values of strength and toughness were up to 625 MPa and 7.5 MPa m1/2, respectively. The
improved toughness is considered to be mainly associated with the de¯ection of cracks along interphase boundaries due to a weak interface
as well as with the introduction of microcracks at the interface between SiC grains and the secondary phases due to a residual tensile stress
from thermal expansion mismatch. # 1999 Elsevier Science S.A. All rights reserved.

Keywords: Silicon carbide; Pressureless sintering; Mechanical properties; Residual stress; Interface; Microcrack

1. Introduction In the present work, the effects of Y2O3 contents in the


(Al2O3 ‡ Y2O3) additives on the densi®cation behavior and
Pressureless sintering has been considered as the most mechanical properties of pressureless-sintered a-SiC cera-
attractive manufacturing technology to obtain ceramic com- mics are investigated and discussed.
ponents with complex shapes, volume production, and low
cost. In the early 1970s, pressureless-sintered SiC ceramics
were successfully developed by Prochazka [1], who found 2. Experimental procedure
that small additions of boron and carbon can drastically
enhance the densi®cation of SiC via a solid-state sintering The starting materials used in this experiment were a-
mechanism. However, the applications of this boron- and SiC, Al2O3 and Y2O3 powders, which have an average
carbon-doped SiC have been seriously limited by its low particle size of 0.6, 0.5 and 0.8 mm, respectively. The
mechanical properties and high sintering temperatures. chemical composition of the a-SiC powder in wt.% was
In order to improve the mechanical properties and lower Si (total) 69.35, Si (free) 0.22, C (total) 29.98, C (free) 0.37,
the sintering temperatures of pressureless-sintered SiC cera- O (total) 0.42, Al2O3 0.18 and Fe2O3 0.12. To investigate the
mics, many investigations have been focused on a variety of in¯uences of Y2O3 contents in the (Al2O3 ‡ Y2O3) additives
other sintering additives over the past few years. Recently, it on the densi®cation behavior and mechanical properties of
has been shown [2±7] that either a- or b-SiC powders may a-SiC, four compositions were made by keeping the total
be pressureless-sintered to high density with the addition of amount of Al2O3 and Y2O3 at 10 wt.%. The theoretical
Al2O3 and Y2O3 at a relatively low temperature of densities of these compositions were calculated according
1850  20008C. The sintering mechanism in this system to the rule of mixtures and are given in Table 1.
is considered to be attributed to liquid-phase sintering via the SiC powders and the (Al2O3 ‡ Y2O3) additives were
formation of a eutectic liquid between Al2O3 and Y2O3. mixed in ethanol using high-purity a-SiC milling media.
When compared with boron- and carbon-doped SiC, Al2O3- After drying in air at 708C for about 8 h and sieving through
and Y2O3-doped SiC exhibits a notably improved strength a 75-mesh screen, the powder mixtures were uniaxially
and toughness. pressed into rectangular specimens of 45 mm
 8 mm  6.5 mm at 3 MPa in a carbon steel die. Subse-
*Corresponding author. quently, these green bodies were compacted in a cold

0254-0584/99/$ ± see front matter # 1999 Elsevier Science S.A. All rights reserved.
PII: S 0 2 5 4 - 0 5 8 4 ( 9 9 ) 0 0 0 3 9 - 5
140 J.H. She, K. Ueno / Materials Chemistry and Physics 59 (1999) 139±142

Table 1 aluminum garnet, YAG). Thus, the formation of Al2O3-


Theoretical densities of SiC-10 wt.% (Al2O3 ‡ Y2O3) YAG eutectic liquid should be responsible for the densi®ca-
Content of Y2O3 in Al2O3 ‡ Y2O3 (wt.%) Theoretical density (g/cm3) tion of SiC ceramics. Since the eutectic temperature of
Al2O3-YAG liquid phase is approximately 17608C [8], it
25.0 3.2846
37.5 3.2907 is possible to obtain maximum sintered densities at tem-
50.0 3.2969 peratures as low as 18508C. In this work, a high density up to
62.5 3.3031 96.8% was attained for SiC-10 wt.% (Al2O3 ‡ Y2O3) cera-
mics at 18508C when the relative content of Y2O3 to
isostatic press at 200 MPa, and sintered in a tightly-threaded (Al2O3 ‡ Y2O3) was 37.5 wt.%. It must be pointed out that
graphite crucible that was placed in a graphite resistance the presence of SiO2 on the surface of SiC powders may
furnace. Sintering was conducted in a ¯owing argon atmo- further enhance densi®cation through liquid-phase sinter-
sphere at a heating and cooling rate of 108C/min. Four ing, because it can decrease the melting point of Al2O3-YAG
different sintering temperatures of 18508C, 19008C, 19508C eutectic phase and increase the amount of liquid [9].
and 20008C were chosen, and the holding time was 1 h. On the other hand, it can be seen in Fig. 1 that sintered
Dimensions and weights of all individual specimens densities deceased sharply when the content of Y2O3 in
before and after sintering were determined to calculate (Al2O3 ‡ Y2O3) was 62.5 wt.%. According to the equili-
linear shrinkages and weight losses. Densities of the sin- brium phase diagram [8] between Al2O3 and Y2O3, this
tered specimens were measured by a conventional water- additive may form a eutectic liquid phase between
displacement method. Phase analysis was performed by 3Y2O35Al2O3 and 2Y2O3Al2O3. Due to a relatively high
X-ray diffractometry (XRD) with Cu Ka radiation. Flexural eutectic temperature of 3Y2O35Al2O3-2Y2O3Al2O3 than
strength was measured by three-point bending tests with a Al2O3-3Y2O35Al2O3, SiC grains cannot be fully wetted by
30 mm span at a cross-head speed of 0.5 mm/min. Fracture liquid phase, and thus the solution-precipitation process of
toughness was determined using the single edge notch beam SiC in the liquid phase is inhibited, resulting in the
(SENB) method with a notch width of 0.25 mm and a decreases in sintered density.
support span of 20 mm. The microstructure of fracture Fig. 2 shows the variation of sintering shrinkage with
surfaces and polished surfaces was observed by scanning Y2O3 addition for SiC containing 10 wt.% (Al2O3 ‡ Y2O3).
electron microscopy (SEM). As expected, the dependence of sintering shrinkage on Y2O3
content correlates well with that of sintered density. How-
ever, the change of shrinkage with temperature is not fully
3. Results and discussion correlative to that of density. At 20008C, maximum shrink-
age values reached but the highest densities were not
Fig. 1 presents the sintered densities of SiC-10 wt.% achieved. This is believed to be mainly associated with
(Al2O3 ‡ Y2O3) ceramics with different Y2O3 contents. some reactions between SiC and additives. The most pos-
As shown, the highest densities were achieved at Y2O3 sible and important reactions are [4]
contents between 25 and 50 wt.% of the total additives. This SiC s† ‡ Al2 O3 s† ! SiO g† ‡ Al2 O g† ‡ CO g† (1)
characteristic is quite important in conventional production
SiC s† ‡ 2Y2 O3 s† ! SiO g† ‡ 4YO g† ‡ CO g† (2)
of sintered SiC ceramics, because it reduces the sensitivity
of properties to composition. Furthermore, X-ray diffraction The escape of gaseous reactants (SiO, Al2O, YO and CO)
analysis revealed that these specimens were composed during sintering may result in the formation of `new' pores
primarily of a-SiC, Al2O3 and 3Y2O35Al2O3 (yttrium- and thus the decrease in sintered density. Since high tem-

Fig. 1. Relative density as a function of Y2O3 content for SiC-10 wt.% Fig. 2. Linear shrinkage as a function of Y2O3 content for SiC-10 wt.%
(Al2O3 ‡ Y2O3) ceramics sintered at 18508C (triangles), 19008C (circles), (Al2O3 ‡ Y2O3) ceramics sintered at 18508C (triangles), 19008C (circles),
19508C (diamonds) and 20008C (squares). 19508C (diamonds) and 20008C (squares).
J.H. She, K. Ueno / Materials Chemistry and Physics 59 (1999) 139±142 141

Fig. 3. Weight loss as a function of Y2O3 content for SiC-10 wt.% Fig. 5. Flexural strength as a function of Y2O3 content for SiC-10 wt.%
(Al2O3 ‡ Y2O3) ceramics sintered at 18508C (triangles), 19008C (circles), (Al2O3 ‡ Y2O3) ceramics sintered at 19008C (circles), 19508C (dia-
19508C (diamonds) and 20008C (squares). monds) and 20008C (squares).

peratures may enhance the evaporation of liquid phases and droplets can be seen on the surface of SiC grains due to a
the reactions between SiC and additives, maximum densi- relatively high amount of Al2O3, which is easy to volatilize
ties were not obtained at 20008C, where the sintering at high temperatures. In contrast, SiC-10 wt.%
shrinkage reached highest values. (Al2O3 ‡ Y2O3) ceramic with 62.5 wt.% Y2O3 in the addi-
During sintering, the volatilization of liquid phases as tives exhibited a `clean' surface. This phenomenon suggests
well as the reactions between SiC and additives may also that an excessive addition of Al2O3 is needed to suppress the
cause the loss of weight. As shown in Fig. 3, the weight loss evaporation of Al2O3 and therefore obtain high sintered
of SiC-10 wt.% (Al2O3 ‡ Y2O3) ceramics increases with densities. As shown in Fig. 1, the sintered densities
increasing temperature. In addition, it can be seen in Fig. 3 increased with increasing Al2O3 addition.
that the weight loss is higher at lower Y2O3 contents, i.e., The ¯exural strength of SiC-10 wt.% (Al2O3‡Y2O3)
Al2O3-rich compositions have higher weight loss than ceramics is illustrated in Fig. 5 as a function of Y2O3
Y2O3-rich compositions. content. Over the whole range of Y2O3 contents, the
Fig. 4 shows the fracture surfaces of SiC-10 wt.% strengths were in excess of 500 MPa, which is at the upper
(Al2O3‡Y2O3) ceramics with different Y2O3 contents. limit of boron- and carbon-doped SiC. Especially at a
For the specimen with 25 wt.% Y2O3 in the additives, many sintering temperature of 19508C, a high strength up to
625 MPa was achieved for SiC-10 wt.% (Al2O3 ‡ Y2O3)
ceramics with 25 wt.% Y2O3 in the additives. This should be
attributed to a relatively homogeneous microstructure with-
out large abnormal grains, which may act as ¯aw origins. In
addition, the strength of SiC-10 wt.% (Al2O3 ‡ Y2O3) cera-
mics decreased with increasing Y2O3 content. The decrease
in strength is estimated to result from the increase in residual
pores, i.e., the decrease in sintered density. As can be seen in
Fig. 1, the sintered density decreased with increasing Y2O3
content.
It has been reported [10] that the fracture mode in boron-
and carbon-doped SiC is predominantly transgranular. How-
ever, it is interesting to note in Fig. 6 that SiC-10 wt.%
(Al2O3 ‡ Y2O3) ceramics exhibited a highly intergranular
fracture behavior. This phenomenon is considered to be
related to the characteristics of secondary phases. As shown
in Fig. 6, the secondary phase is primarily located as a
continuous ®lm at the grain boundaries. Since the major part
of the intergranular phase was YAG, which has a larger
thermal expansion coef®cient than SiC, a residual tensile
stress would be generated at the interface between SiC and
YAG upon cooling down from sintering temperatures. The
existence of such a stress may weaken the interphase
boundaries, leading to intergranular fracture.
Fig. 4. Fracture surfaces of SiC-10 wt.% (Al2O3‡Y2O3) ceramics with Fig. 7 shows the fracture toughness of SiC-10 wt.%
(A) 25 wt.% and (B) 62.5 wt.% Y2O3 in the additives. (Al2O3 ‡ Y2O3) ceramics as a function of Y2O3 content.
142 J.H. She, K. Ueno / Materials Chemistry and Physics 59 (1999) 139±142

Fig. 6. SEM micrograph of a fracture surface for SiC-10 wt.% Fig. 8. SEM micrograph of Vickers indentation cracks in SiC-10 wt.%
(Al2O3 ‡ Y2O3) ceramics sintered at 19508C with 37.5 wt.% Y2O3 in (Al2O3 ‡ Y2O3) ceramics sintered at 19508C with 37.5 wt.% Y2O3 in the
the additives. additives.

When compared with boron- and carbon-doped SiC, Al2O3- 4. Conclusion


and Y2O3-doped SiC exhibited much higher toughness
values. Especially, a high fracture toughness in excess of SiC ceramics with an addition of 10 wt.%
7.5 MPa m1/2 was attained for SiC-10 wt.% (Al2O3 ‡ (Al2O3 ‡ Y2O3) were investigated by pressureless sintering
Y2O3) ceramics sintered at 20008C with 37.5±50 wt.% to determine the effects of Y2O3 content and sintering
Y2O3 in the additives. The notably improved toughness is temperature on the densi®cation behavior and mechanical
considered to be mainly associated with the crack de¯ection properties. A high sintered density up to 98% was achieved
process. It can be seen in Fig. 8 that SiC-10 wt.% at a sintering temperature of 19508C for SiC-10 wt.%
(Al2O3 ‡ Y2O3) ceramics showed signi®cant crack de¯ec- (Al2O3 ‡ Y2O3) with Y2O3 contents between 25 and
tion along interphase boundaries due to a weak interface. On 50 wt.% of the total additives. This is considered to be
the other hand, the residual tensile stresses at the interface attributed to the formation of a eutectic liquid between
between SiC and YAG, which were induced on cooling after Al2O3 and Y2O3 at sintering temperatures. Owing to a
sintering, may result in microcracking at the SiC/YAG homogeneous microstructure without exaggerated grains,
interface ahead of a primary propagating crack. These a high strength of about 625 MPa was achieved. SEM
microcracks dissipate the strain energy and shield the main observation revealed that SiC-10 wt.% (Al2O3 ‡ Y2O3)
crack, leading to the increase in fracture toughness. ceramics showed a highly intergranular fracture behavior
Additionally, it can be seen in Fig. 7 that the toughness of due to a residual tensile stress at the interface between SiC
SiC-10 wt.% (Al2O3 ‡ Y2O3) ceramics increased with grains and the secondary phases. On the other hand, SiC
increased sintering temperature. This phenomenon is con- exhibited signi®cant crack de¯ection along interphase
sidered to be related to the coarsening of SiC grains at a boundaries. As a result, a high toughness in excess of
relatively high sintering temperature, because an exagger- 7.5 MPa m1/2 was attained.
ated grain may enhance the crack bridging and branching,
leading to the increase in fracture toughness.
References

[1] Prochazka, S., in: J.J. Burke, A.E. Gorum, R.M. Katz (Eds.),
Proceedings of the Conference on Ceramics for High-Performance
Applications, Brook Hill Publishing Co., Chestnut Hill, 1975, p. 239.
[2] M. Omori, H. Takei, J. Mater. Sci. 23 (1988) 3744.
[3] R.A. Cutler, T.B. Jackson, in: V.J. Tenny (Ed.), Proceedings of the
Third International Symposium on Ceramic Materials and Compo-
nents for Engine, American Ceramic Society, Westerville, 1989,
p. 309.
[4] M.A. Mulla, V.D. Krstic, Am. Ceram. Soc. Bull. 70 (1991) 439.
[5] L.S. Sigl, H.J. Kleebe, J. Am. Ceram. Soc. 76 (1993) 773.
[6] N.P. Padture, J. Am. Ceram. Soc. 77 (1994) 519.
[7] S.K. Lee, C.H. Kim, J. Am. Ceram. Soc. 77 (1994) 1655.
[8] E.M. Levin, C.R. Robbins, H.F. McMurtie, in: M.K. Reser (Ed.),
Phase Diagram for Ceramists, American Ceramic Society, Colum-
Fig. 7. Fracture toughness as a function of Y2O3 content for SiC-10 wt.% bus, 1969, fig. 2344.
(Al2O3 ‡ Y2O3) ceramics sintered at 19008C (circles), 19508C (dia- [9] Y.W. Kim, M. Mitomo, J.G. Lee, J. Ceram. Soc. Jpn. 104 (1996) 816.
monds) and 20008C (squares). [10] S.K. Lee, Y.C. Kim, C.H. Kim, J. Mater. Sci. 29 (1994) 5321.

You might also like