You are on page 1of 12

Effect of alumina-doping on grain boundary

Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

segregation-induced phase transformation in yttria-stabilized


tetragonal zirconia polycrystal
Koji Matsui,a) Nobukatsu Ohmichi, and Michiharu Ohgai
Synthetic Chemicals Production, Tosoh Corporation, Shunan, Yamaguchi 746-8501, Japan
Hidehiro Yoshida
National Institute for Materials Science, Tsukuba, Ibaraki 305-0047 Japan
Yuichi Ikuhara
Institute of Engineering Innovation, School of Engineering, The University of Tokyo, Bunkyo-ku,
Tokyo 113-8656, Japan

(Received 14 March 2006; accepted 3 May 2006)

The microstructure in a small amount of Al2O3-doped Y2O3-stabilized tetragonal


zirconia polycrystal (Y-TZP) sintered at 1100–1650 °C was examined to clarify the
effect of Al3+ ions segregated at grain boundaries on cubic-formation and grain-growth
processes. The sintering rate in Y-TZP was remarkably enhanced by Al2O3-doping. In
addition, at temperatures >1500 °C, grain growth remarkably proceeded, and the
fraction of the cubic phase increased in comparison with that of undoped Y-TZP.
High-resolution electron microscopy and nanoprobe x-ray energy dispersive
spectroscopy measurements revealed that no amorphous layer existed along the
grain-boundary faces in Al2O3-doped Y-TZP and that Y3+ and Al3+ ions segregated at
grain boundaries over widths of ∼10 and ∼6 nm, respectively. At 1100 °C, Al3+ ions
started to segregate at grain boundaries, and the segregation peak of Al3+ ions
increased as the sintering temperature increased. Cubic-formation and grain-growth
behaviors in Al2O3-doped Y-TZP were reasonably interpreted by taking into account
the effect of Al3+ ions segregating along grain boundaries.

I. INTRODUCTION found in Y-TZP in comparison to that in Y2O3-stabilized


Yttria-stabilized tetragonal zirconia polycrystal (Y- cubic zirconia. The model of the Zener’s pinning effect
TZP) is an important structural ceramic with excellent was firstly proposed by Lange.2 Subsequently, Sakuma
mechanical properties, such as high fracture toughness, and Yoshizawa3,4 quantitatively analyzed the grain-
strength, and hardness, and is used for products such as growth process using a rate equation based on Zener’s
optical fiber connectors, grinding media, and precision pinning model to interpret the grain growth of dual-phase
mechanical parts. The hydrolysis process is the best in- ZrO2 generated after the equilibrium partitioning. Grain-
dustrial manufacture method for producing Y-TZP pow- boundary characterization is a key to reveal the grain-
der. To develop new Y-TZP powder with forming and growth mechanism. The grain-boundary structure in Y-
sintering characteristics by the hydrolysis process, it is TZP was intensively investigated by high-resolution
particularly important to clarify the sintering mechanism electron microscopy (HREM), and it had been concluded
in Y-TZP. that an amorphous phase existed in grain boundaries.5–10
So far, two models based on the solute-drag effect and However, Ikuhara et al.11 first reported that, from HREM
Zener’s pinning effect have been proposed to understand and nanoprobe x-ray energy dispersive spectroscopy
the grain-growth behavior in Y-TZP. The model of the (EDS) measurements in Y-TZP, no amorphous phase is
solute-drag effect in Y-TZP was first proposed by Lee observed along grain boundaries, but Y3+ ions segregate
and Chen1 to explain the slow grain-growth behavior at grain boundaries over a width of 4–6 nm. The grain-
boundary structure was clarified by their observation re-
sults. On the other hand, using the Rietveld analysis,
a)
Address all correspondence to this author. Ohmichi et al.12 investigated the formation behavior of
e-mail: k_matui@tosoh.co.jp the cubic phase in Y-TZP and reported that the amount of
DOI: 10.1557/JMR.2006.0274 cubic phase increased and the Y2O3 concentration within

2278 J. Mater. Res., Vol. 21, No. 9, Sep 2006 © 2006 Materials Research Society
K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP

tetragonal phase decreased as the sintering temperature used to measure the average grain sizes of the sintered
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

increased.12 We reported in-depth analysis of microstruc- bodies. SEM specimens were polished with 3-␮m dia-
ture change in Y-TZP during the sintering process, using mond paste and then thermally etched for 1 h at a tem-
HREM– and scanning transmission electron microscopy perature of 50 °C lower than sintering temperature of
(STEM)–nanoprobe EDS techniques, and the cubic- each specimen in air. The average grain size was meas-
formation and grain-growth mechanisms can be ex- ured by the Planimetric method.18
plained using the Grain Boundary Segregation-Induced
Phase Transformation (GBSIPT) model and the solute- C. XRD measurements
drag effect of Y3+ ions segregating along the grain X-ray diffraction (XRD) profiles were measured at
boundary, respectively.13,14 room temperature using a powder diffractometer system
On the other hand, from the industrial standpoint, it is (Model MXP3, MAC Science Co. Ltd, Tokyo, Japan) and
particularly important to clarify the role of Al2O3, which recorded using Cu K␣ radiation, under 40 kV and 30 mA,
is one of the additives accelerating sintering. The role of at a slit width of 1°. The fractions (mass%) of cubic (fc)
Al2O3 on the microstructure and properties of Y-TZP has and tetragonal (ft) phases in the sintered bodies were
been reported previously by several researchers.15–17 determined by the Rietveld method. The Rietveld calcu-
However, the effect of Al2O3 on the microstructural for- lation was conducted in a manner identical to those re-
mation process in Y-TZP has not yet been clarified. ported by Ohmichi et al.14
In the present study, we examined the microstructure
of grain boundaries and grain interiors in a small amount D. Transmission electron microscopy and
of Al2O3-doped Y-TZP sintered at 1100–1650 °C to sys- nanoprobe EDS measurements
tematically investigate the effect of Al2O3-doping on
cubic-formation and grain-growth processes in detail. The microstructure was examined by field emission
The Al2O3-doping effect was discussed by comparing the transmission electron microscopy (TEM; Model 002BF,
results of the present analysis with previous analysis for Topcon, Tokyo, Japan). Specimens for TEM observa-
undoped Y-TZP.14 tions were mechanically ground to a thickness of
∼0.1 mm, further dimpled to a thickness of ∼10 ␮m, and
then ion-milled for electron transparency. Specimens sin-
II. EXPERIMENTAL PROCEDURES tered at 1100 and 1200 °C were very porous and there-
fore were reinforced by permeating the epoxy resin be-
A. Specimen preparation
fore sample preparation. HREM observations were con-
A small amount of Al2O3-doped Y-TZP was prepared ducted to examine the grain-boundary structure, using
as follows. Y-TZP powder containing 2.9 mol% (5.2 Topcon 002BF (Tokyo, Japan) with a point-to-point
mass%) Y2O3 and 0.31 mol% (0.25 mass%) Al2O3, and resolution of 0.17 nm. Nanoprobe EDS measurements
with a specific surface area of 15 m2/g (TZ-3YE grade, were performed to examine quantitatively segregation of
Tosoh, Tokyo, Japan), manufactured by the hydrolysis Y3+ and Al3+ ions at grain boundaries, using a Noran
method was used as a starting raw material. This powder Voyager System in the Topcon 002BF (Tokyo, Japan)
was uniaxially pressed into a disk under 70 MPa pres- with a probe size of 0.5 nm. Four to six grain boundaries
sure. The resulting green compacts were sintered at in one specimen were analyzed. STEM and nanoprobe
1100–1650 °C for 2 h at a constant heating rate of EDS analyses were performed to examine the Y3+ ion
100 °C/h in air. The specimens of the resulting Al2O3- distribution within grains, using a Noran Voyager Sys-
doped Y-TZPs are termed 3YE. The experimental con- tem in the Topcon 002BF with a probe size of 1 nm.
dition to prepare the specimens is the same as the con-
dition to sinter Y-TZP powder without Al2O3 described
in our previous paper.14 The Y-TZP powder without III. RESULTS
Al2O3 (TZ-3Y grade, Tosoh, Tokyo, Japan) used in our A. Densification, grain growth, and T → C
previous paper had a Y2O3 concentration of 2.9 mol% phase transformation
(5.2 mass%) and a specific surface area of 15 m2/g.14 The
Figure 1 shows the relationship between relative den-
previous data (the undoped Y-TZPs) are termed 3Y.
sity and sintering temperature for the 3YEs (Al2O3-
doped Y-TZPs) sintered at 1100–1650 °C. The relative
B. Density and grain size measurements density of 3YE increased with increasing sintering tem-
The density of sintered bodies was measured by the perature and attained 99% at 1350 °C. The 3Y (undoped
Archimedes method. In the case of the sintered bodies Y-TZP) data reported in our previous paper14 are also
with relative density of <80%, the density was calculated plotted in Fig. 1. The relative density of 3YE was higher
from the weight and the size. Scanning electron micros- than that of 3Y at sintering temperatures of <1500 °C.
copy (SEM; Model S-4500, Hitachi, Tokyo, Japan) was Figure 2 shows typical SEM images of 3YEs sintered at

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2279


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP

compared with 3Y data,14 it was seen that the decreasing


Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

ft of 3YE showed nearly the same tendency as that of 3Y,


whereas the increasing fc of 3YE exhibited nearly the
same tendency as that of 3Y up to 1500 °C, but when this
temperature was exceeded, the fc value of 3YE increased
more than that of 3Y as the sintering temperature in-
creased. If we make an analogy to our previous results,13
we can presume that the slight difference in fc is caused
primarily by the fact that the Y2O3 concentration within
the tetragonal phase of 3YE decreases more rapidly than
that of 3Y.
These experimental results revealed that when a small
amount of Al2O3 was doped to Y-TZP, the sintering rate
was remarkably enhanced, grain growth proceeded re-
markably, and the fc increased slightly at temperatures of
>1500 °C.

B. Internal structure
TEM measurements were performed to observe the
FIG. 1. Dependence of relative density and average grain size in Y- change in the internal structure in 3YE with an increase
TZPs with sintering temperature. (䊉) and (䊏) indicate the relative in the sintering temperature. Figure 4 shows typical TEM
density and the average grain size of 3YE, respectively. (嘷) and (䊐) images of 3YEs sintered at (a) 1100 °C, (b) 1200 °C, (c)
indicate the relative density and the average grain size of 3Y, respec-
1300 °C, (d) 1500 °C, and (e) 1650 °C. In the 3YE sin-
tively.14
tered at 1100 °C, the network structure joined through
the neck parts among starting particles was observed.
When the sintering temperature attained 1200 °C, the
(a) 1100 °C, (b) 1300 °C, (c) 1500 °C, and (d) 1650 °C.
form of each particle and the network structure were no
The grain size of 3YE increased as the sintering tempera-
longer seen, and grains and large pores were observed. At
ture increased. To quantitatively examine the behavior of
1300 °C, the large pores disappeared. The grain-growth
grain growth with increasing temperature, the average
behavior is thus consistent with the result shown in
grain sizes of 3YE were determined by the Planimetric
Fig. 1.
method. The grain size of 3YE sintered at each tempera-
ture is also plotted in Fig. 1. The grain size increased
gradually up to 1400 °C and then rapidly above this tem- C. Grain-interior microstructure
perature. Comparing with previous 3Y data,14 it can be To examine the change in the microstructure with T →
seen that the grain-growth behavior of 3YE showed the C phase transformation in detail, the Y3+ ion distribution
same tendency as that of 3Y up to 1500 °C, but when this of grain interiors in 3YE was analyzed using STEM and
temperature was exceeded, the grain size of 3YE in- element-mapping nanoprobe EDS techniques. Figure 5
creased more rapidly than that of 3Y as the sintering shows typical STEM images and Y K␣ and Zr K␣ map-
temperature increased. ping images obtained by the STEM-nanoprobe EDS
Figure 3 shows the relationship between the fc or ft method in 3YEs sintered at (a) 1100 °C, (b) 1200 °C, (c)
value and the sintering temperature for 3YEs sintered at 1300 °C, (d) 1500 °C, and (e) 1650 °C, respectively. Be-
1100–1650 °C. The crystal phase in 3YE was monopha- cause the Zr K␣ maps of Figs. 5(a)–5(e) showed the same
sic tetragonal zirconia (i.e., ft ⳱ 100 mass%) up to homogeneous distribution, the distribution of Zr4+ ions is
1200 °C. At 1300 °C, the cubic phase appeared and the almost the same irrespective of the sintering tempera-
value of fc was 12.7 mass%. The value of fc for 3YE tures. However, when the Y K␣ mapping patterns of
increased with increasing sintering temperature and Figs. 5(a)–5(e) are compared, Y3+ ions are drastically
reached 22.9 mass% at 1650 °C. The value of ft de- changed to heterogeneously distribute with the increas-
creased with increasing sintering temperature. In our pre- ing sintering temperature. In the specimens sintered at
vious paper, we reported that the amount of the cubic 1300 °C and below [Y K␣ maps in Figs. 5(a)–5(c)], most
phase formed by the tetragonal-to-cubic (T → C) phase of the grains were tetragonal phase because the distribu-
transformation increases with both decreasing ft value tion of Y3+ ions in grain interiors was nearly homoge-
and Y2O3 concentration within the tetragonal phase.13 neous. At 1500 °C [Y K␣ map, Fig. 5(d)], the regions
Therefore, the present result can be also interpreted by with high Y3+ ion concentrations were formed in grain
the same cubic-formation process. When 3YE data were interiors, and the regions started to form from grain

2280 J. Mater. Res., Vol. 21, No. 9, Sep 2006


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 2. SEM images of 3YEs sintered at (a) 1100 °C, (b) 1300 °C, (c) 1500 °C, and (d) 1650 °C.

boundaries and triple junctions. When the sintering tem- faces. This observation agrees with the previous
perature increased to 1650 °C [Y K␣ map, Fig. 5(e)], the data.11,13,14,19,20 Taking into account the results of
regions with high Y3+ ion concentrations were further HREM, Rietveld, and STEM-nanoprobe EDS analyses, it
partitioned. These observation results were nearly the can be concluded that 3YEs sintered at 1100 and 1200 °C
same as those of 3Y.14 were composed of a single-crystal particle with mono-
According to Fig. 3, the fc increased with increasing phasic tetragonal.
sintering temperature. Assuming that the high Y3+ ion Figure 7 shows typical Y-distribution profiles across
concentration regions are cubic phase, the present result the grain in 3YEs sintered at (a) 1100 °C, (b) 1200 °C,
agrees qualitatively with the fc behavior in Fig. 3. It is (c) 1300 °C, (d) 1500 °C, and (e) 1650 °C. The nano-
therefore concluded that the high Y3+ ion concentration probe EDS measurements were performed with the probe
regions generated in grain interiors transform from te- size of 0.5 nm at every 1 nm across the grain boundaries.
tragonal to cubic phases and the cubic phase regions are In the 3YE sintered at 1100 °C, Y3+ ions started to seg-
partitioned as the sintering temperature increases. It is regate at grain boundaries. The Y2O3 concentration at the
presumed that this T → C phase transformation occurs by grain boundary is ∼5 mol%, which is a little bit higher
redistribution of Y3+ ions in grain interiors and the re- than that in the grain interior. This segregation profile
distribution process is related to the grain boundary. represents the initial step of the segregation of Y3+ ions.
At 1300 °C, the segregation peak of Y3+ ions clearly
D. Grain-boundary segregation
appeared. For the profiles in Figs. 7(a)–7(e), the Y2O3
To probe the formation behavior of the grain-boundary concentrations in grain interiors were in the range of
segregation of Y3+ ions with increasing sintering tem- 1–4 mol%, which corresponds to that in the tetragonal
perature, grain-boundary structures and composition in phase. It was therefore confirmed that Y3+ ions segre-
3YE were measured using HREM and nanoprobe EDS gated at grain boundaries over a width of ∼10 nm in all
techniques, which are very effective in analyzing a lo- grain boundaries between tetragonal and tetragonal
calized region. Figure 6 shows typical HREM images of phases (T-T grain boundaries). The segregation behavior
the grain-boundary faces in 3YE sintered at (a) 1100 °C, of Y3+ ions obtained at the present experiments revealed
(b) 1200 °C, (c) 1300 °C, (d) 1500 °C, and (e) 1650 °C. a tendency similar to that of 3Y.14
The images have been taken with the boundary edge-on For the 3YEs sintered at 1300 °C and below, Y3+ ions
to observe the grain-boundary structure directly. In segregate at all grain boundaries, and the Y2O3 concen-
the HREM images of Figs. 6(a)–6(e), no amorphous or trations in grain interiors correspond to that in the te-
second phase was observed along the grain-boundary tragonal phase. Therefore, it is concluded that the cubic-

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2281


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP

considered to be the possibility of transforming from


Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

tetragonal to cubic phases. To examine this possibility,


the fraction of the cubic phase ( f⬘c) is determined by
assuming that the grain is a sphere with a diameter of D
and that the crystal structure of the grain-boundary vi-
cinity at which Y3+ ions segregate is the cubic phase. The
calculated value of f⬘c is given by Eq. (1)14:

D3 − 共D − ␻兲3
f⬘c共mass%兲 ≅ × 100 . (1)
D3

Here, ␻ is the width of the segregation of Y3+ ions.


Substituting the average grain size (D ≅ 200 nm) and the
width (␻ ≅ 6 nm) of the segregation of Y3+ ions at
1300 °C in Eq. (1), f ⬘c ≅ 9 mass% is obtained. This
calculated value is slightly smaller than the fc value
(12.7 mass%) at 1300 °C determined by the Rietveld
analysis. Therefore, this result reveals that most of the
grain boundaries at 1300 °C in which Y3+ ions segregate
have transformed from tetragonal to cubic phases, and
further, a part of cubic-phase regions formed in grain
boundaries and triple junctions may extend slightly in the
grain interior.
Because a small amount of Al2O3 was doped to 3YE,
not only Y3+ ions but also Al3+ ions were observed to
segregate at grain boundaries over a width of ∼6 nm in
T-T grain boundaries [Figs. 7(a)–7(e)]. In the 3YE sin-
tered at 1100 °C, the small segregation peak of Al3+ ions
was observed, and Al3+ ions started to segregate at grain
boundaries. At 1200 °C, the segregation peak of Al3+
ions clearly appeared. The segregation peak of Al3+ ions
increased as the sintering temperature increased. At
1650 °C, the segregation-peak height of Al3+ ions
reached ∼4.6 mol% Al2O3 and was ∼0.84 times that of
Y3+ ions (∼5.5 mol% Y2O3).
As can be seen in Figs. 5(d) and 5(e), the interfaces
between cubic and tetragonal phases are also formed in a
single domain in 3YEs sintered at 1500 °C and over. As
shown in Fig. 8, the different segregation profile of Y3+
ions from the T-T grain-boundaries was also observed at
1650 °C. The HREM image corresponding with Fig. 8
is also shown in Fig. 9. In the right side of the grain
FIG. 3. Dependence of the fractions of cubic and tetragonal phases in interior, the Y2O3 concentration is over 6 mol%, which
Y-TZPs with sintering temperature. (嘷) and (䊐) indicate 3YE and
3Y,14 respectively. indicates that the right side grain is a cubic phase. It was
therefore assigned to the grain boundary between cubic
and tetragonal phases (C-T grain boundary). In the C-T
phase regions with high Y3+ ion concentration in grain grain boundary, Al3+ ions segregated at grain boundaries
interiors shown in Figs. 5(d) and 5(e) started to form from over a width of ∼6 nm.
grain boundaries and the triple junctions in which Y3+ The present results reveal that though Al3+ ions seg-
ions segregate. This mechanism can be reasonably inter- regate in grain boundaries, the cubic-formation behavior
preted using the GBSIPT model proposed in our previous in 3YE can be explained using the GBSIPT mechanism
papers.13,14 recognized for 3Y.14 It is therefore thought that the
Analogizing from the above conclusion, the region of Al2O3-doped in Y-TZP does not greatly influence the
the segregation peak appearing clearly at 1300 °C is GBSIPT mechanism.

2282 J. Mater. Res., Vol. 21, No. 9, Sep 2006


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 4. Conventional bright-field TEM images of 3YEs sintered at (a) 1100 °C, (b) 1200 °C, (c) 1300 °C, (d) 1500 °C, and (e) 1650 °C.

IV. DISCUSSION boundary. As a result, Y3+ ions segregate slightly and a


segregation profile of Y3+ ions is formed in grain bound-
Based on the present analytical results, the effect of aries, as shown in Fig. 7(a). Subsequently, at the inter-
Al2O3-doping on cubic-formation and grain-growth mediate sintering stage, neck growth between particles
mechanisms in Y-TZP can be considered as follows. The completes and then grain formation and growth occur. At
grain-boundary segregation process of Y3+ ions and the the first step in the intermediate stage, neck growth con-
cubic-formation mechanism from the initial to final sin- tinues to proceed, and as a result, the size of the single
tering stages in 3YE are summarized schematically in crystal with monophasic tetragonal increases. The Y3+
Fig. 10. At 1100 °C classified as the initial sintering ion distribution of particle interiors is homogeneous, but
stage, neck formation and growth between particles oc- the segregation peak of Y3+ ions starts to form in grain
cur [Fig. 10(a)]. Each particle is the single crystal com- boundaries. At 1200 °C, neck growth is completed as the
posed of monophasic tetragonal and the distribution of particle form originating in the starting powder disap-
Y3+ ions in particle interiors is homogeneous. In the pears and the grains are formed [Fig. 4(b)].
grain-boundary faces formed in the neck parts, particles At 1300 °C, classified as the final sintering stage
are joined directly without amorphous or second phase. [Fig. 10(b)], the Y3+ ion distribution of grain interiors is
When the grain boundary starts to form with neck almost homogeneous [Fig. 5(c)], but the well-defined
growth, Y3+ ions existing in the grain-boundary neigh- segregation peak of Y3+ ions is formed in grain bound-
borhood diffuse and start to segregate in the grain aries [Fig. 7(c)]. Grain boundaries migrate with the

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2283


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 5. STEM images and Y K␣ and Zr K␣ mapping images by STEM-nanoprobe EDS method, of 3YEs sintered at (a) 1100 °C, (b) 1200 °C,
(c) 1300 °C, (d) 1500 °C, and (e) 1650 °C. The white broken lines in the mapping images indicate the grain boundaries. Bright parts in the Y K␣
mapping images correspond to regions with high Y3+ ion concentrations.

2284 J. Mater. Res., Vol. 21, No. 9, Sep 2006


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 6. HREM images of the grain-boundary faces in 3YEs sintered at (a) 1100 °C, (b) 1200 °C, (c) 1300 °C, (d) 1500 °C, and (e) 1650 °C.

well-defined segregation peak of Y3+ ions as the sin- growth, but also the T → C phase transformation, occur
tering temperature increases (i.e., this behavior corre- at the final stage. A part of the crystal phase in Y-TZP
sponds to grain growth). On the other hand, not only becomes the cubic phase, which is thermodynamically
grain-boundary segregation of Y 3+ ions and grain stable as the sintering temperature increases. Because the

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2285


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 7. Y- and Al-distribution profiles across the T-T grain boundaries in 3YEs sintered at (a) 1100 °C, (b) 1200 °C, (c) 1300 °C, (d) 1500 °C,
and (e) 1650 °C. (嘷) and (䊐) indicate Y and Al, respectively.

regions in which Y3+ ions segregate in grain boundaries regions of the cubic phase, which is thermodynamically
and triple junctions have a tendency to cause T → C stable, Y3+ ions in grain interiors dominantly migrate in
phase transformation, the cubic phase starts to form from the direction of grain boundaries and triple junctions with
grain boundaries and triple junctions when the sintering grain growth (i.e., the GBSIPT structure is formed). As a
temperature at which the cubic phase is thermodynami- result of such behavior of Y3+ ions, cubic grains in Y-
cally stable is attained. We named such a new diffusive TZP are formed via the GBSIPT structure.
transformation phenomenon from a grain boundary This cubic-formation mechanism is applicable from a
GBSIPT.13 When this temperature is exceeded, to extend certain step in the final stage where the segregation peak

2286 J. Mater. Res., Vol. 21, No. 9, Sep 2006


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP

Y3+ ion concentrations are formed in the grain interior


Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

[Figs. 10(b)–10(c)]. These regions with high Y3+ ion


concentrations immediately change into the cubic phase
by T → C phase transformation. When the cubic regions
increase, the Y2O3 concentrations of the tetragonal re-
gions further decrease by diffusion of Y3+ ions to the
cubic regions. When the sintering temperature increases
further, regions of the cubic phase proceed to the regions
of tetragonal phase with grain growth, and then grains
with only cubic phase are formed [Fig. 10(d)]. Thus, the
present results in 3YE are also reasonably interpreted
using the GBSIPT mechanism recognized for 3Y.14
Grain growth in Y-TZP proceeds in intermediate and
final stages of sintering. In particular, the grain size in-
creases rapidly at the final stage. We reported that since
Y3+ ions segregated at T-T and C-T grain-boundaries
during sintering, based on the theory of the impurity-drag
effect derived by Cahn,21 the grain-growth mechanism in
Y-TZP is reasonably interpreted by the solute-drag effect
of Y3+ ions segregating along grain boundaries.13,14
In the present experiments, it is revealed that when a
small amount of Al2O3 is doped to Y-TZP, the densifi-
FIG. 8. Y- and Al-distribution profiles across the C-T grain bound- cation rate is remarkably enhanced at initial and inter-
aries in 3YE sintered at 1650 °C. (嘷) and (䊐) indicate Y and Al, mediate stages, and the grain size increases with the same
respectively. tendency up to 1500 °C and rapidly at temperatures of
>1500 °C. Because the densification process occurring at
initial and intermediate stages is caused primarily by
pore shrinkage with neck growth, it is thought that the
densification behavior is determined by the characteris-
tics of the grain boundary relating directly to neck
growth. At 3YE sintered at 1100 °C, Al3+ ions slightly
segregate at grain boundaries over a width of ∼6 nm. In
this segregated amount, no difference in the relative den-
sities of 3YE and 3Y appears (i.e., the relative density of
3YE equals that of 3Y). However, at 1200 °C, in which
the segregation peak of Al3+ ions appears clearly, the
relative density of 3YE increases more rapidly than that
of 3Y. At 1300 °C, the segregation peak of Al3+ ions
grows further, and the relative density has already at-
tained in the range of the final stage. These results ap-
parently reveal that because the densification rate in the
initial and intermediate stages is enhanced as the segre-
gation peak of Al3+ ions grows, Al3+ ions segregated at
grain boundaries over a width of ∼6 nm relate directly to
acceleration of neck growth.
When the sintering temperature increases, the segre-
FIG. 9. HREM image of the grain-boundary face in 3YE sintered at gation peak of Al3+ ions grows further and at 1650 °C,
1650 °C. segregation-peak heights of Al3+ ions are attained in the
range of about 4.6–5.9 mol% Al2O3. Because the grain
of Y3+ ions clearly appears. At 1300 °C, most grain size of 3YE increases more rapidly than that of 3Y at
boundaries in which Y3+ ions segregate have transformed temperatures >1500 °C, it is thought that when the
from tetragonal to cubic phases. In this step, when the amount of Al3+ ions segregated at grain boundaries ex-
sintering temperature increases, the widths of Y3+ ions ceeds a certain value, Al3+ ions accelerate grain growth.
segregating in grain boundaries and triple junctions in- According to the impurity-drag theory,21 when an impu-
crease with grain growth, and then the regions with high rity exists in a grain boundary, grain growth is depressed

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2287


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

FIG. 10. Schematic illustration of the cubic-formation mechanism from the initial to final sintering stages for 3YE. The gray parts of the
grain-boundary indicate segregation of Y3+ ions. The white and gray regions of grain interior represent the tetragonal and cubic phases,
respectively. GB stands for grain boundary.

because the grain boundary migrates by dragging the which is formed by the GBSIPT mechanism. In the pres-
impurity segregated in the grain boundary. However, be- ent results, at temperatures >1500 °C, at which the dif-
cause the effect of Al3+ ions segregated at grain bound- ference in grain growth appears, the fc of 3YE also in-
aries is different from the grain-growth behavior pre- creases more than that of 3Y. Therefore, it is thought that
dicted from the impurity-drag theory, it cannot be inter- Al3+ ions segregated at grain boundaries also accelerate
preted using this theory. On the other hand, Yoshida the formation of the cubic phase.
et al.22 have reported the small amount of dopant effect
on the grain-growth behavior in Y-TZP. They have ex-
perimentally revealed that the growth constant related to V. CONCLUSIONS
the grain-boundary diffusivity increased as the ionic ra- In the present study, the microstructure in a small
dius of dopant decreased and argued that the dopant cat- amount of Al2O3-doped Y-TZP at 1100–1650 °C was
ions segregated at grain boundaries change grain- measured to clarify the effect of Al2O3-doping on cubic-
boundary diffusivity.22 Taking into account the present formation and grain-growth processes. The following re-
results based on the result reported by Yoshida et al.,22 it sults were obtained.
is concluded that Al3+ ions segregated at grain bound- When a small amount of Al2O3 was doped to Y-TZP,
aries have the effect of enhancing grain-boundary diffu- the sintering rate was enhanced remarkably, grain growth
sion because the ionic radius of Al3+ ions (0.068 nm) is proceeded remarkably, and the fc slightly increased at
smaller than that of Y3+ ions (0.104 nm).23 Therefore, it temperatures >1500 °C.
is supposed that at temperatures >1500 °C, the predomi- No amorphous or second phase existed along the
nant effect that enhances grain-boundary diffusion is grain-boundary faces, but Y3+ and Al3+ ions segregated
caused by Al3+ ions rather than the solute-drag effect of at grain boundaries over widths of ∼10 and ∼6 nm, re-
Y3+ ions, and as a result, the grain-growth behavior of spectively. The segregation peak of Al3+ ions at the grain
3YE is enhanced more rapidly than that of 3Y. boundary increased as the sintering temperature in-
Furthermore, Al3+ ions segregated at grain boundaries creased.
also affect the formation behavior of the cubic phase, Al3+ ions segregated at grain boundaries relate directly

2288 J. Mater. Res., Vol. 21, No. 9, Sep 2006


K. Matsui et al.: Effect of alumina-doping on grain boundary segregation-induced phase transformation in Y-TZP

to acceleration of the densification rate in the initial and for the superplasticity mechanism. Acta Metall. Mater. 43, 1211
Downloaded from https://www.cambridge.org/core. Gadjah Mada University, on 28 Feb 2019 at 02:29:09, subject to the Cambridge Core terms of use, available at https://www.cambridge.org/core/terms. https://doi.org/10.1557/jmr.2006.0274

intermediate stage of sintering. Furthermore, at tempera- (1995).


11. Y. Ikuhara, P. Thavorniti, and T. Sakuma: Solute segregation at
tures >1500 °C, Al3+ ions segregated at grain boundaries grain boundaries in superplastic SiO2-doped TZP. Acta Mater. 45,
enhanced grain growth and the fc. 5275 (1997).
12. N. Ohmichi, K. Kamioka, K. Ueda, K. Matsui, and M. Ogai:
Phase transformation of zirconia ceramics by annealing in hot
REFERENCES water. J. Ceram. Soc. Jpn. 107, 128 (1999).
13. K. Matsui, H. Horikoshi, N. Ohmichi, M. Ohgai, H. Yoshida, and
1. I.G. Lee and I-W. Chen: Sintering and grain growth in tetragonal
Y. Ikuhara: Cubic-formation and grain-growth mechanisms in te-
and cubic zirconia, in Sintering ’87, edited by S. Somiya,
tragonal zirconia polycrystal. J. Am. Ceram. Soc. 86, 1401 (2003).
M. Shimada, M. Yoshimura, and R. Watanabe (Elsevier Applied
14. K. Matsui, N. Ohmichi, M. Ohgai, H. Yoshida, and Y. Ikuhara:
Science, London, UK, 1988), pp. 340, 345.
Grain boundary segregation-induced phase transformation in
2. F.F. Lange, D.B. Marshall, and J.R. Porter: Controlling micro-
yttria-stabilized tetragonal zirconia polycrystal. J. Ceram. Soc.
structures through phase partitioning from metastable precursors:
Jpn. 114, 230 (2006).
The ZrO2–Y2O3 system, in Ultrastructure Processing of Ad-
vanced Ceramics, edited by J.D. Mackenzie and D.R. Ulrich 15. S.N.B. Hodgson, J. Cawley, and M. Clubley: The role of Al2O3
(Wiley, New York, 1988), pp. 519, 532. impurities on the microstructure and properties of Y-TZP.
3. Y. Yoshizawa and T. Sakuma: Evolution of microstructure and J. Mater. Proc. Technol. 92–93, 85 (1999).
grain growth in ZrO2–Y2O3 alloys. ISIJ Int. 29, 746 (1989). 16. T.S. Suzuki, Y. Sakka, K. Morita, and K. Hiraga: Enhanced su-
4. T. Sakuma and Y. Yoshizawa: The grain growth of zirconia dur- perplasticity in a alumina-containing zirconia prepared by colloi-
ing annealing in the cubic/tetragonal two-phase region. Mater. Sci. dal processing. Scripta Mater. 43, 705 (2000).
Forum 94–96, 865 (1992). 17. I.M. Ross, W.M. Rainforth, D.W. McComb, A.J. Scott, and
5. T. Stoto, M. Nauer, and C. Carry: Influence of residual impurities R. Brydson: The role of trace additions of alumina to
on phase partitioning and grain growth processes of Y-TZP ma- yttria-tetragonal zirconia polycrystals (Y-TZP). Scripta Mater. 45,
terials. J. Am. Ceram. Soc. 74, 2615 (1991). 653 (2001).
6. M.L. Mecartney: Influence of an amorphous second phase on the 18. T. Yamaguchi: Characterization techniques of ceramic: Properties
properties of yttria-stabilized tetragonal zirconia polycrystals (Y- of sintered bodies. Ceram. Jpn. 19, 520 (1984).
TZP). J. Am. Ceram. Soc. 70, 54 (1987). 19. J. Zhao, Y. Ikuhara, and T. Sakuma: Grain growth of silica-added
7. T.G. Nieh, D.L. Yaney, and J. Wadsworth: Analysis of grain zirconia annealed in the cubic/tetragonal two-phase region. J. Am.
boundaries in a fine-grained, superplastic, yttria-containing, te- Ceram. Soc. 81, 2087 (1998).
tragonal zirconia. Scripta Metall. 23, 2007 (1989). 20. P. Thavorniti, Y. Ikuhara, and T. Sakuma: Microstructural char-
8. T. Hermansson, H. Swan, and G. Dunlop: The role of the inter- acterization of superplastic SiO2-doped TZP with a small amount
granular glassy phase in the superplastic deformation of Y-TZP of oxide addition. J. Am. Ceram. Soc. 81, 2927 (1998).
zirconia, in Euro-Ceramic, Vol. 3, edited by G. du With, 21. J.W. Cahn: The impurity-drag effect in grain boundary motion.
R.A. Terpsta, and R. Metselaar (Elsevier Applied Science, Lon- Acta Metall. 10, 789 (1962).
don, UK, 1989), pp. 329–333. 22. H. Yoshida, H. Nagayama, and T. Sakuma: Small dopant effect on
9. M.M.R. Boutz, C.S. Chen, L. Winnubst, and A.J. Buggraaf: Char- static grain growth and flow stress in superplastic TZP. Mater.
acterization of grain boundaries in superplastically deformed Y- Trans. 44, 935 (2003).
TZP ceramics. J. Am. Ceram. Soc. 77, 2632 (1994). 23. R.D. Shannon: Revised effective ionic radii and systematic studies
10. S. Primdahl, A. Thölén, and T.G. Langdon: Microstructure ex- of interatomic distances in halides and chalcogenides. Acta Crys-
amination of a superplastic yttria-stabilized zirconia: Implications tallogr., Sect. A 32, 751 (1976).

J. Mater. Res., Vol. 21, No. 9, Sep 2006 2289

You might also like