You are on page 1of 19

Authors Accepted Manuscript

Effect of sintering temperature on mechanical


properties of magnesia partially stabilized zirconia
refractory
Lan Jiang, Shuqiang Guo, Yuyang Bian, Man
Zhang, Weizhong Ding
www.elsevier.com/locate/ceri

PII:
DOI:
Reference:

S0272-8842(16)30274-7
http://dx.doi.org/10.1016/j.ceramint.2016.03.136
CERI12507

To appear in: Ceramics International


Received date: 21 December 2015
Revised date: 18 March 2016
Accepted date: 18 March 2016
Cite this article as: Lan Jiang, Shuqiang Guo, Yuyang Bian, Man Zhang and
Weizhong Ding, Effect of sintering temperature on mechanical properties of
magnesia partially stabilized zirconia refractory, Ceramics International,
http://dx.doi.org/10.1016/j.ceramint.2016.03.136
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
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.

Effect of sintering temperature on mechanical properties of magnesia partially stabilized zirconia


refractory
Lan Jiang, Shuqiang Guo,*, Yuyang Bian, Man Zhang, Weizhong Ding
State Key Laboratory of Advanced Special Steel & Shanghai Key Laboratory of Advanced
Ferrometallurgy & School of Materials Science and Engineering, Shanghai University, Shanghai
200072, China
Corresponding

author. Tel.(Fax): +86 21 56338244. sqguo@shu.edu.cn

Abstract
The optimized sintering conditions for a 3.5wt% magnesia partially stabilized zirconia (Mg-PSZ)
refractory were proposed in our recent research. The influence of the sintering temperature on the
development of phase composition, microstructure, densification, thermal expansion and mechanical
strength were studied in detail by X-ray diffraction (XRD), scanning electron microscope (SEM),
He-pycnometer, high temperature dilatometry and three-point bending test. The samples sintered at
1670oC had the highest bend strength, the maximum densification, the lowest thermal expansion
coefficient (CTE), a homogeneous microstructure and a linear change in thermal expansion.
Keywords: A. Sintering; C. Mechanical properties; D. ZrO2; E. Refractories
1. Introduction
Vacuum induction melting (VIM) processing has been the primary melting method for
nickel-based superalloy because it does not impact melt chemistry and gives results in good
homogeneity of the melt [1]. But VIM is the only vacuum melting method which uses a refractory
crucible made of oxides such as Al2O3 and MgO [2]. Unfortunately, the refractory is a contamination
source of oxygen because of metal/refractory reactions leading to crucible disintegration. Oxygen is a
1

harmful trace element that exists both in solid solution and oxide inclusions in the superalloy. Oxide
inclusions can act as crack initiation sites and propagation paths, so it can dramatically affect the
properties of the nickel-based superalloy [3]. Therefore, it is important to design an appropriate
refractory to melt this superalloy. The stability of the refractory can be described by the following
sequence:Y2O3CaOZrO2Al2O3MgO [4]. It is apparent that Y2O3 is the most stable crucible
lining material. But Y2O3 is fairly expensive, which is a drawback [5]. CaO is not used because it is
very susceptible to humidity [6] and therefore not suitable for crucible lining in industrial furnaces.
MgO and Al2O3 crucible materials also cause a problem because of their unstability under high vacuum
melting. The dissociation of these refractories leads to oxygen pick-up in the melt, which could
eventually result in oxygen inclusion formation when the solubility limit is exceeded. Besides, the use
of MgO crucible accelerates the formation of high melting inclusions (MgAl2O4) which deteriorate the
cleanliness of the alloy [7]. Compared with refractories such as magnesia and alumina, zirconia
ceramics show better chemical stability. Aneziris et al. [8] reported that the low porosity ZrO2 based
materials have been used in the near shape steel casting for high corrosion resistance. Hence, ZrO2 is
the most suitable refractory for melting nickel-based superalloy.
However, pure zirconia (ZrO2) has three polymorphs depending on temperature: monoclinic (m)
up to1170oC, tetragonal (t) to 2370oC, and above this temperature, cubic (c) [9-10]. After sintering at
temperatures between 1500 and 1750oC, pure zirconia ceramics break into pieces at room temperature
[11]. Due to the destructive tm phase transformation, pure zirconia is used quite rarely [12]. However,
the effect of this phase transformation can be eliminated by doping zirconia with appropriate amount of
oxides such as MgO and CeO2 [13]. In view of the tremendous technological applications, PSZ
sintering has received much attention from scientists and technologists. Mechanical properties can be
2

enhanced by preparing PSZ ceramics that have high densities and small grain sizes after sintering [14].
Because of the different rates of pore closure, larger pores will develop into a porous microstructure
during densification. These pores can be removed only at a high sintering temperature and a long
sintering time [15]. Response surface methodology (RSM) was applied to determine the operational
conditions for the properties of PSZ. The sintering temperature and heating rate have different effects
on the densification and bending strength - sintering temperature is positive, heating rate is negative
[16]. Although Youngs modulus of the ZrO2-3mol% Y2O3 does not change with sintering temperature,
a slight decrease is observed in the hardness values above 1000 oC, which is attributed to microstructure
coarsening [17].
The Mg-PSZ refractory has already become one of the most important materials because of its
corrosion resistant, excellent chemical and thermomechanical properties [18]. Nevertheless, full
exploitation of these advantages can only be realized if the final products have the required
specifications comprising desired phase, density, microstructure and the bend strength [19]. Based on
the equilibrium phase diagram MgO-ZrO2 [20], the transformation temperature of tm is about
1240oC and the tetragonal and cubic solid solution becomes stable above 1400 oC. A feature of
technological importance is the cubic and tetragonal solid solution phase field [8]. The crystal structure
transformations involving volume changes exhibited thermal hysteresis, which may cause cracking
defects. The hysteresis was controlled by several microstructure and chemical factors, such as grain
size, tetragonal size, and the sintering process [21-22].
In this paper, the optimized sintering conditions for the 3.5wt% Mg-PSZ refractory are proposed.
The influence of the sintering temperature on the development of phase composition, microstructure,
densification, thermal expansion and mechanical strength were studied in detail.
3

2. Experimental Procedure
Commercial magnesia partially stabilized zirconia (3.5wt% Mg-PSZ) powders (provided by
Sunshine Capital Ltd, China) were selected as the starting material with a grain size distribution of
0-10m and d50=2.7m. Solid discs (20 mm in diameter) and bars of 3.5wt% Mg-PSZ were attained by
slip casting. In order to form suspensions, the 3.5wt% Mg-PSZ powders were dispersed in deionized
water with an organic binder and dispersant, then milled for 4 h. The organic binder and dispersants
were Arabic gum and Triethanolamine respectively. The stable suspensions with 78wt% solids content
were slip cast in plaster mold to obtain the green body. The green samples were sintered at different
temperatures of 1600, 1620, 1650, 1670, 1690 oC with a dwell time at temperature of 4 h.
The crystal structure of samples sintered at different temperatures was characterized by XRD
(D/MAX2200V PC). Morphology was examined using a scanning electron microscope (SEM, Hitachi
SU-1510). The amount of densification was obtained by the ratio of the densities of sintered samples
determined using a He-pycnometer (AccuPyc II 1340) to the theoretical density. Differential scanning
calorimetry analysis (DSC, Netzsch STA 449 F3) was employed to evaluate the phase transition
temperature of 3.5wt% Mg-PSZ powders in air from room temperature to 1350oC. The bend strength of
samples which were sintered at different temperatures was investigated using a three-point bending test
with a span length of 30mm on samples of 3mm thickness4mm width40 mm length. Five samples
were tested to obtain the average data.
Linear shrinkage (dL/L0) of the 3.5wt%Mg-PSZ powders was measured in air using a push-rod
dilatometer (Netzsch DIL 402C) from the room temperature to 1450C. The size of the sample bars for
the dilatometer analysis was 8mm length6 mm diameter.

3. Results and discussion


3.1. Sintering behavior of the 3.5wt% Mg-PSZ refractory
The sintering of green samples after drying took place in the dilatometer, which recorded the
shrinkage

value (dL/L0)

and the

rate of linear

shrinkage (d(dL/L0) of green bodies

during heating. The dynamic sintering curve of 3.5wt% Mg-PSZ green body in air is presented in Fig.1.
It can be seen that there was no shrinkage below 1000 oC, but dramatically declined between 1000oC
and 1240oC. The rapid shrinkage is caused by the formation of grain boundaries between the particles
and a reduction in porosity. The maximum rate of shrinkage around 1240 oC was related to the phase
transformation from monoclinic to tetragonal according to the equilibrium phase diagram MgO-ZrO2
[20]. A large volume change is associated with this transformation, which was believed to be the reason
for the catastrophic failure of zirconia ceramic [23]. It was noteworthy that the rate of shrinkage
increased at 1426oC, which may be caused by the phase transformation from tetragonal to cubic
according to the phase diagram of MgO-ZrO2. Thus, it was important to control sintering condition
around this temperature. With increasing temperature, the rate of linear shrinkage gradually decreased.
The DSC curve of 3.5wt% Mg-PSZ green samples is shown in Fig.2. Two endothermic peaks
occurred at 330oC and 1240oC respectively. Because decomposition temperature of Arabic gum and
Triethanolamine are about 330oC, the initial endothermic peak could be attributed to the decomposition
of organics which were introduced during the slip casting. At this temperature, the effect of organics
can be eliminated by prolonging soaking time during sintering. The other endothermic peak at 1240 oC
corresponded to the phase transformation, which was consistent with the phase diagram and sintering
curve.

Based on the equilibrium phase diagram MgO-ZrO2 [20], the dynamic sintering curve and DSC, a
controlled sintering process for 3.5wt% Mg-PSZ samples was determined as shown in Fig.3. From
room temperature to 330oC, a low heating rate of 1.5oC min-1 and a holding time of 60 min were used
to decompose the organic binder. Subsequently, the sintering temperature was increased to 1000oC
using a heating rate of 3oC min-1. From 1000oC to 1240oC, the initial sintering of the 3.5wt% Mg-PSZ
refractory occurred and a lower heating rate of 1.5oC min-1was used. A holding time of 120 min was
used at 1240oC to allow the Mg-PSZ phase to transform from monoclinic to tetragonal. From 1240oC
to 1426oC, a lower heating rate of 1.5oC min-1was used, and a hold time of 120 min at 1426oC due to
the phase transformation from tetragonal to cubic. Samples were further sintered at various final
sintering temperatures from 1600oC to 1700oC using a heating rate of 1.5oC min-1. The holding time at
the final temperature remained 240 min to allow grain growth and solidification. The optimal final
sintering temperature was determined by the thermomechanical properties and microstructure of
samples sintering at different temperatures.
3.2 Phase evolution
Polyphase ZrO2 is generally observed when the MgO content is below 13mol% [24]. Fig.4 shows
XRD patterns of 3.5wt% Mg-PSZ refractory sintered at different temperatures. The constituent phases
in the sintered samples were the tetragonal and cubic phases, with small amounts of monoclinic phase.
The (101)t and (111)c peaks overlap each other because they have nearly the same lattice parameters for
the tetragonal and cubic phases [13]. By XRD spectrum, with increasing sintering temperature, the
tetragonal and cubic phases were found to increase. When the sintering temperature are above 1600oC,
the tetragonal and cubic phases become stable on the basis of the equilibrium phase diagram
MgO-ZrO2 [20].
6

The relative amount of the monoclinic phase of zirconia is of considerable importance in attempts
to understand the mechanical properties of zirconia-based ceramics [25]. Therefore, it is necessary to
determine the amount of monoclinic phase. The amount of monoclinic phase Xm was calculated in
accordance to Eq. (1) as previously described [24]:
Xm = (Im(111) + Im(111) )/(It(101)+c(111) + Im(111) + Im(111) )

(1)

Where Im(111) , Im(111) and It(101)+c(111) represent the integrated intensities of the corresponding
peaks. The results of the amount of monoclinic phase Xm sintered at different final temperatures are
shown in Fig.5. It can be seen that the amount of m-phase approached nearly 80% at 1600oC. However,
the zirconia materials with high amounts of monoclinic phase (above 50%) made the mechanical
strengths and corrosion resistance worse [8]. Though high amounts of monoclinic phase are bad for

mechanical properties, a certain amount of monoclinic in zirconia materials is necessary for good
thermal shock performance. Aneziris et al. [26] reported that the zirconia based materials with an
amount of 13% m-phase can achieve a good thermal shock performance. At 1670oC, the amount of
m-phase was about 15%, which is similar value as the report, thus having a minimum value for good
thermal shock performance. When the sintering temperature reached 1690oC, the amount of monoclinic
phase would be predicted to be less than 10%.
3.3 Densification and Microstructure
The densification of zirconia specimens sintered at different final temperature is shown in Fig.6.
As can be seen from the result, the sintering temperature had a significant effect on the densification.
With rising sintering temperature, the densification initially increased and then decreased, reaching a
maximum value at 1670oC. When temperature reached 1690oC, because of grains growth, some pores
were trapped by grains, resulting in a slight reduction of densification [27]. The micrographs of the
7

samples sintered at different temperatures are presented in Fig.7. The green body contained a large
amount of pores, and there were only point contact between particles. At 1450oC, a microstructure with
clear grain boundaries was obtained. But many pores were also found on the surface (Fig.7a). With
increasing temperature, the pores decreased gradually (Fig.7b). From Fig.7c, it shows the grain grew
further and the particles contacted with each other more and more close. It is clearly observed the
agglomerations of the fine particles. It means that the decrease in the numbers of pores results in the
increase of densification with increased sintering temperature. At the same time, larger grain size
implies a reduction in the grain boundaries. However, the microphotograph revealed an exaggerated
grain growth, with grain sizes becoming as coarse as 60m when the sintering temperature was 1690oC
(Fig.7e). This behavior was in good agreement with densification. Higher temperatures promote
exaggerated grain growth rather than further densification. This indicated that the combined effects of
particle size and the numbers of pores are all equally critical factors for optimal densification. Besides,
the increased grain size may result in enhanced crack formation [28]. When the sintering temperature
reached 1670oC, the size of grain was only about 40m and the structure uniformity was not found to
have exaggerated grain growth (Fig.7d). The similar micrographs of 3.5wt% Mg-PSZ samples sintered
at 1670oC were found.
3.4 Thermal and Mechanical properties
Thermal stress resistance can be used to predict the thermal shock behavior of material. The
thermal shock parameter R was calculated using the equations [29]:
R = (1 )/E

(2)

where is the bend strength, E is Youngs modulus, is the thermal expansion coefficient, and is
Poissons ratio. Considering the parameter R, it is clear that better thermal shock behavior can be
8

achieved in materials with high strength and Poissons ratio, and with low values of thermal expansion
coefficient and Youngs modulus. Besides, the change of linear thermal expansion also affects thermal
shock behavior. In spite of the higher values of the thermal expansion, more thermal expansion led to
better thermal shock behavior of the samples.
The thermal expansion curves of 3.5wt% Mg-PSZ refractory during heating and cooling, which
were sintered at different temperatures, are shown in Fig.8. With increased sintering temperature, the
linear shrinkage at the temperature around 1240 oC and the linear thermal expansion at the temperature
around 1400oC were both reduced. It was considered that the tm transformation temperature during
heating and cooling was directly related to the thermal stabilities of the tetragonal and monoclinic
phase [30]. Sample sintered at 1600oC exhibited a remarkable thermal hysteresis. The tm phase
transformation took place at around 800oC on cooling, which was much lower than the mt
transformation on heating, causing the hysteresis loops and the associated linear change. The thermal
expansions trended to a linear change with increasing sintering temperatures. The thermal hysteresis
area of curves has been decreased for the samples when sintering temperatures are higher than 1650oC.
It means that the tetragonal and cubic phase of Mg-PSZ materials could be stabilized at room
temperature.
The thermal expansion coefficient at temperature T was calculated as [31]:
= (L/0 )/ dT

(3)

where L0 is the length of a sample at room temperature. The value of linear shrinkage (dL/L0) of
samples was measured using a dilatometer. The curves of thermal expansion coefficient (CTE) are
shown for the different sintered samples in the Fig.9. The lower value of CTE should improve thermal
shock behavior of the samples. Though the curve of CTE was the lowest when samples were sintered at
9

1600oC, the thermal expansion curves as shown in Fig.8 exhibited a remarkable thermal hysteresis,
which would result in poor thermal shock behavior. The thermal expansion coefficients increased with
increasing sintering temperature. This phenomenon was consistent with the amount of m-phase in
samples as discussed earlier. The value of CTE of monoclinic phase was lower than cubic and
tetragonal phases [32]. The amount of monoclinic phase decreased with increasing sintering
temperature, which lead to an increase in the CTE value. As can be seen from Figs.8 and Figs.9, when
the samples were sintered at 1670oC, the amount of m-phase was about 15%. This composition offers a
low thermal shock coefficient accompanied by an acceptable thermal hysteresis. With increased
sintering temperature, the decrease of the amount of m-phase result in an increase of the thermal
expansion coefficient. This is a reason why zirconia based material should contain a certain amount of
monoclinic phase.
The effect of sintering temperature on bend strength was investigated by studying densification
and grain size [33]. Fig.10 shows the bend strength of Mg-PSZ refractory sintered at different
temperatures which was tested at room temperature. The strength-change behavior was observed to be
similar to the strength-sintering temperature behavior. With increasing sintering temperature, ceramic
particles connected closer and the structure of the sintered compacts became stronger, which led to
higher strength. The highest bend strength of 634 MPa was obtained with 3.5wt% Mg-PSZ sintered at
1670oC. As can be seen from Figs.6 and Figs.7, at temperature higher than 1670oC, the grain size of the
3.5wt% Mg-PSZ refractory becomes larger and the densification of the 3.5wt% Mg-PSZ refractory
decreases gradually. It must be mentioned that higher temperatures cause a lower densification and
larger grains, which results in decreasing mechanical strength of the 3.5wt% Mg-PSZ refractory.
Therefore, 1670oC was the most suitable temperature for sintering Mg-PSZ refractory because of the
10

high mechanical properties and lower CTE.


4. Conclusions
In this work, 3.5wt% Mg-PSZ refractories were prepared by slip casing using commercially
produced MgO stabilized zirconia powders and then sintered under a controlled sintering process. The
optimized sintering conditions were determined using MgO-ZrO2 phase diagram, high temperature
dilatometry and differential scanning calorimetry analysis. A sintering temperature of 1670oC proved to
be the optimum final sintering temperature. The samples sintered at 1670 oC had the highest bend
strength of 634 MPa, the maximum densification and the lower value of CTE. Therefore, by selecting
this sintering temperature, it is possible to obtain optimized mechanical properties of Mg-PSZ for
potential applications in melting superalloy.
Acknowledgements
This study was financially supported by the ShangHai University and the Joint Funds of the National
Natural Science Foundation of China (U1560202).
References
[1] A. Choudhury, State of the art of superalloy production for aerospace and other application using
VIM/VAR or VIM/ESR, ISIJ. Int. 32 (1992) 563-574.
[2] J. Otubo, O.D. Rigo, C.M. Neto, P.R. Mei, The effects of vacuum induction melting and electron
beam melting techniques on the purity of NiTi shape memory alloys, Mat. Sci. Eng. A-Struct.
438(2006), 679-682.
[3] J.P. Niu, X.F. Sun, T. Jin, K.N. Yang, H.R. Guan, Z.Q. Hu, Investigation into deoxidation during
vacuum induction melting refining of nickel base superalloy using CaO crucible, Mater. Sci. Tech-lond.
19 (2003) 435-439.
11

[4] T. Degawa, T. Ototani, Refining of high purity Ni-base superalloy using calcia refractory, ISIJ. Int.
73 (1987) 1691-1697.
[5] S.K. Sadrnezhad, S.B. Raz, Interaction between refractory crucible materials and the melted NiTi
shape-memory alloy, Metall. Mater. Trans. B. 36 (2005) 395-403.
[6] E. Serris, L. Favergeon, M. Pijolat, M. Soustelle, P. Nortier, R.S. Grtner, T.Chopin, Z. Habib,
Study of the hydration of CaO powder by gassolid reaction. Cement Concrete Res. 41 (2011)
1078-1084.
[7] M. Jiang, X.H. Wang, J.J. Pak, Formation of low-melting-point inclusions in Al-deoxidized steel
refined by high-basicity calcium aluminate slag in ZrO2 crucible experiments, Metall. Mater. Trans. B.
45 (2014) 1248-1259.
[8] C.G. Aneziris, E.M. Pfaff, H.R. Maier, Corrosion mechanisms of low porosity ZrO2 based materials
during near net shape steel casting, J. Eur. Ceram. Soc. 20 (2000) 159-168.
[9] B. Basu, J. Vleugels, O. Van Der Biest, Transformation behaviour of tetragonal zirconia: role of
dopant content and distribution, Mat. Sci. Eng. A-Struct. 366 (2004) 338-347.
[10] M.H. Bocanegra-Bernal, S.D De La Torre, Phase transitions in zirconium dioxide and related
materials for high performance engineering ceramics, J. Mater. Sci. 37 (2002) 4947-4971.
[11] C. Piconi, G. Maccauro, Zirconia as a ceramic biomaterial, Biomaterials. 20 (1999) 1-25.
[12] M.A. Borik, V.T. Bublik, A.V. Kulebyakin, E.E. Lomonova, F.O. Milovich, V.A. Myzina, V.V.
Osiko, N.Y. Tabachkova, Phase composition, structure and mechanical properties of PSZ (partially
stabilized zirconia) crystals as a function of stabilizing impurity content, J. Alloy Compd. 586 (2014)
S231-S235.
[13] J.A. Brito-Chaparro, A. Reyes-Rojas, M.H. Bocanegra-Bernal, A. Aguilar-Elguezabal, J.
12

Echeberria, Elucidating of the microstructure of ZrO2 ceramics with additions of 1200oC heat treated
ultrafine MgO powders: Aging at 1420oC, Mater. Chem. Phys. 106 (2007) 45-53.
[14] I.R. Gibson, G.P. Dransfield, J.T.S. Irvine, Sinterability of commercial 8 mol% yttria-stabilized
zirconia powders and the effect of sintered density on the ionic conductivity, J. Mater. Sci. 33 (1998)
4297-4305.
[15] V. Srdi, M. Winterer, H. Hahn, Sintering behavior of nanocrystalline zirconia doped with alumina
prepared by chemical vapor synthesis, J. Am. Ceram. Soc. 83 (2000) 1853-1860.
[16] J. Li, J. Peng, S. Guo, L. Zhang, Application of response surface methodology (RSM) for
optimization of sintering process for the preparation of magnesia partially stabilized zirconia (Mg-PSZ)
using natural baddeleyite as starting material, Ceram. Int. 39 (2013) 197-202.
[17] A. Daz-Parralejo, A.L. Ortiz, R. Caruso, Effect of sintering temperature on the microstructure and
mechanical properties of ZrO2-3mol% Y2O3 sol-gel films, Ceram. Int. 36 (2010) 2281-2286.
[18] M.O. Suk, J.H. Park, Corrosion behaviors of zirconia refractory by CaO-SiO2-MgO-CaF2 slag, J.
Am. Ceram. Soc. 92 (2009) 717-723.
[19] S. Koley, A. Ghosh, A.K Sahu, R.Tewari, A.K. Suri, Correlation of compaction pressure, green
density, pore size distribution and sintering temperature of a nano-crystalline 2Y-TZP-Al2O3 composite,
Ceram. Int. 37 (2011) 731-739.
[20] C.F. Grain, Phase Relations in the ZrO2-MgO System, J. Am. Ceram. Soc. 50 (1967) 288-290.
[21] Q. Liu, S.L.An, W.H. Qiu, Study on thermal expansion and thermal shock resistance of Mg-PSZ,
Solid State Ionics. 121 (1999) 61-65.
[22] F. Abe, S. Muneki, K. Yagi, Tetragonal to monoclinic transformation and microstructural evolution
in ZrO2-9.7mol% MgO during cyclic heating and cooling, J. Mater. Sci. 32 (1997) 513-522.
13

[23] S. Yoon, C.J. Van Tyne, H. Lee, Effect of alumina addition on the microstructure and grain
boundary resistance of magnesia partially-stabilized zirconia, Curr. Appl. Phys. 14 (2014) 922-927.
[24] C.J. Howard, R.J. Hill, The polymorphs of zirconia: phase abundance and crystal structure by
Rietveld analysis of neutron and X-ray diffraction data, J. Mater. Sci. 26 (1991) 127-134.
[25] D.L. Porter, A.H. Heuer, Microstructural Development in MgO-Partially Stabilized Zirconia
(Mg-PSZ), J. Am. Ceram. Soc. 62 (1979) 298-305.
[26] C.G. Aneziris, E.M. Pfaff, H.R. Maier, Fine grained MgPSZ ceramics with titania and alumina or
spinel additions for near net shape steel processing, J. Eur. Ceram. Soc. 20 (2000) 1729-1737.
[27] N.F. Amat, A. Muchtar, M.J. Ghazali, N. Yahaya, Suspension stability and sintering influence on
yttria-stabilized zirconia fabricated by colloidal processing, Ceram. Int. 40 (2014) 5413-5419.
[28] B. Stawarczyk, M. zcan, L. Hallmann, A. Ender, A. Mehl, C.H. Hmmerlet, The effect of
zirconia sintering temperature on flexural strength, grain size, and contrast ratio, Clin. Oral. Invest. 17
(2013) 269-274.
[29] C. Aksel, P.D. Warren, Thermal shock parameters [R, R and R ] of magnesia-spinel composites,
J. Eur. Ceram. Soc. 23 (2003) 301-308.
[30] W.C.J. Wei, Y.P. Lin, Mechanical and thermal shock properties of size graded Mg-PSZ refractory,
J. Eur. Ceram. Soc. 20 (2000) 1159-1167.

[31] H. Hayashi, T. Saitou, N. Maruyama, H. Inaba, K. Kawamura, M. Mori, Thermal expansion


coefficient of yttria stabilized zirconia for various yttria contents, Solid State Ionics.176 (2005)
613-619.

[32] M.V. Swain, R.C. Gaevie, R.H.J. Hannink, Influence of thermal decomposition on the mechanical
properties of magnesia-stabilized cubic zirconia, J. Am. Ceram. Soc. 66 (1983) 358-362.
14

[33] L.F. Hu, C.A. Wang, Effect of sintering temperature on compressive strength of porous
yttria-stabilized zirconia ceramics, Ceram. Int. 36 (2010) 1697-1701.

Fig.1.Temperature dependence of shrinkage (dL/L0) and shrinkage rate (d(dL/L0)/dT) of the 3.5wt%
Mg-PSZ sample in the course of heating(5 oC /min).

Fig.2.Differential scanning calorimetry of 3.5wt% Mg-PSZ green body.

15

Fig.3. Heating pattern for sintering the 3.5wt% Mg-PSZ samples.

Fig.4.X-ray diffraction patterns of 3.5wt% Mg-PSZ powders sintered at different temperatures.

Fig.5. Amount of M-phase of 3.5wt% Mg-PSZ powders sintered at different temperatures.

Fig.6.Densification of 3.5wt% Mg-PSZ powers sintered at different temperatures.

16

Fig.7.SEM micrographs of 3.5wt% Mg-PSZ powers sintered at different temperatures: (a) 1450oC, (b)
1600oC, (c) 1650oC, (d) 1670oC, (e) 1690oC.

Fig.8.Thermal expansion curve of 3.5wt% Mg-PSZ samples sintered at different temperatures.

17

Fig.9. The CTE curve of 3.5wt% Mg-PSZ samples sintered at different temperatures.

Fig.10.The bend strength of 3.5wt% Mg-PSZ with sintering at different temperatures under at room
temperature.

18

You might also like