Você está na página 1de 10

The influence of grain size on low-temperature degradation

of dental zirconia

Lubica Hallmann,1 Albert Mehl,2 Peter Ulmer,3 Eric Reusser,3 Johannes Stadler,4
Renato Zenobi,4 Bogna Stawarczyk,1 Mutlu Özcan,1 Christoph H. F. Hämmerle1
1
Clinic of Fixed and Removable Prosthodontics and Dental Material Science, Center of Dental Medicine,
University of Zürich, Zurich, Switzerland
2
Clinic of Preventive Dentistry, Periodontology and Cariology, Computer Assisted Restorations,
Center of Dental Medicine, University of Zürich, Zurich, Switzerland
3
Institute of Geochemistry and Petrology, ETH Zürich, Switzerland
4
Department of Chemistry and Applied Biosciences, ETH Zürich, Switzerland

Received 13 July 2011; revised 29 August 2011; accepted 6 September 2011


Published online 25 November 2011 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/jbm.b.31969

Abstract: The purpose of this study was to evaluate the of Zr-OH and Y-OH bonds. No increase of yttrium concentra-
influence of grain size and air abrasion on low-temperature tion on the grain boundaries of Y-TZP was detected, which
degradation (LTD) of yttria stabilized tetragonal zirconia poly- could be responsible for the destabilization of dental zirconia
crystalline (Y-TZP). Disc-shaped specimens were sintered at ceramics. A slight increase of diffusion-controlled t–m phase
1350, 1450, and 1600 C. Air abrasion was performed with dif- transformations was observed for all abraded specimens sin-
ferent abrasive particles. The specimens were stored for 2 h tered at 1350 and 1450 C. The size of abrasive particles did
at 134 C under 2.3 bar water vapor pressure. All specimens not play a crucial role on LTD of Y-TZP. The retardation of
were characterized by X-ray powder diffraction analysis diffusion-controlled t–m phase transformation was evident
(XRD), Raman spectroscopy, X-ray photoelectron spectros- for all abraded specimens sintered at 1600 C by comparison
copy (XPS), atomic force microscopy (AFM), and field emis- to non-abraded specimens. Conclusion: The LTD of Y-TZP
sion scanning electron microscopy (FESEM). Y-TZP sintered can be suppressed when the sintering temperature is set
at a temperature of 1350 C did not undergo the t–m phase between 1350 and 1450 C. VC 2011 Wiley Periodicals, Inc. J Biomed

transformation during accelerated aging. The diffusion-con- Mater Res Part B: Appl Biomater 100B: 447–456, 2012.
trolled t–m phase transformation initiated with the specimens
sintered at 1450 C. This transformation was remarkable for Key Words: dental zirconia (Y-TZP), degradation, grain size,
the specimens sintered at 1600 C. X-ray photoelectron spec- air abrasion, tensile stress
troscopy (XPS) measurements did not confirm the generation

How to cite this article: Hallmann L, Mehl A, Ulmer P, Reusser E, Stadler J, Zenobi R, Stawarczyk B, Özcan M, Hämmerle CHF. 2012. The
influence of grain size on low-temperature degradation of dental zirconia. J Biomed Mater Res Part B 2012:100B:447–456.

INTRODUCTION of a number of zirconia hip joints has drawn the attention


Dental zirconia ceramic, consisting of 3 mol % yttria stabi- of researchers to the behavior of Y-TZP ceramics in humid
lized tetragonal polycrystalline zirconia (Y-TZP), has excel- environments. Fractures are often associated with a large
lent mechanical, biomedical, aesthetic properties with its amount of monoclinic phase created during tetragonal–
semitranslucent and radiopacity features.1–3 The develop- monoclinic phase transformation in the human body.6,7 This
ment of CAD/CAM systems, the increase in aesthetic transformation reduces the mechanical properties of Y-TZP
demand and, worries about metallic hypersensitivity have and is known as low-temperature degradation (LTD).6 The
increased the application of Y-TZP in dentistry. Y-TZP has spontaneous transformation from the tetragonal into mono-
extraordinary strength and toughness due to the transfor- clinic structure initiates on the surface of Y-TZP via a stress
mation from the tetragonal to the monoclinic polymorph corrosion type mechanism and proceeds to the bulk of the
when external stress is applied.4,5 For these reasons, Y-TZP ceramic.4,7 The principal mechanism of LTD is diffusion-con-
is used to substitute the traditional metal-based fixed dental trolled, while the t–m modification transformation, once
prosthesis (FDPs). Zirconia full-ceramic copings, FDP frame- nucleated, is martensitic.4 The transformation proceeds
works, and implant suprastructures are introduced to most rapidly at temperatures of 200–300 C.8 The destabili-
replace the traditionally used metals.4 However, the failure zation of the tetragonal phase is accompanied by the

Correspondence to: L. Hallmann; e-mail: lubica.hallmann@zzm.uzh.ch

V
C 2011 WILEY PERIODICALS, INC. 447
formation of micro- and macrocracks and is dependent on Therefore, the purpose of the present study was to
the concentration of the stabilizer, and the grain size of ce- evaluate the effect of grain size, flaws, and pre-existing
ramic.6–10,11 monoclinic phase created during air abrasion on the low-
Y-TZP undergoes a stress-induced volume-expansion temperature degradation (LTD) of dental Y-TZP ceramic
phase transformation during air abrasion. Flaws or micro- under above-mentioned conditions. One of the most impor-
cracks can thereby be introduced in the Y-TZP surface.12,13 tant factors on LTD mechanism of dental zirconia is tensile
The pre-existing monoclinic phase on the air-abraded Y-TZP strain, which is grain size dependant. An additional goal of
surface hinders the propagation of the diffusion-controlled this study was to test the idea that flaws, micro-cracks, and
transformation during accelerated aging.4 pre-existing monoclinic phase play a certain role on the LTD
The aging of zirconia is a serious problem for the ortho- of Y-TZP, but the most important factor is the critical grain
pedic community.4 In conventional dental applications, the size. An attempt was made to ascertain the proper mecha-
zirconia core or framework is not in contact with the oral nism of LTD during the accelerated aging of Y-TZP ceramic.
environment and hard dental tissues due to veneering and The working hypothesis was that the resulting LTD of Y-TZP
luting materials and the problem seems not to be predomi- ceramic under hydrothermal condition is not primarily con-
nant.4 However, the luting resin could absorb water via den- trolled by the concentration gradient of yttrium on grain
tal tubules and the zirconia core and framework exposed to boundaries and/or the Zr-OH and Y-OH bonds formation
water may suffer from aging degradation.4 Additionally, resulting from OH– anion diffusion, but critically depends on
more and more fully ceramic zirconia restorations are the species that diffuse rapidly in the lattice of Y-TZP
inserted, where zirconia surfaces are directly opposed to ceramic.
the intraoral environment.
Several theoretical causes have been proposed to explain MATERIALS AND METHODS
the mechanism of LTD. Sato et al. suggested the chemical Specimen preparation
reaction of water with Zr-O-Zr bonds at the pre-existing sur- Y-TZP ceramic blanks (Cercon base, DeguDent, Hanau, Germany)
face flaws.9 This reaction is associated with the formation of (ZrO2, 5% Y2O3, < 2% HfO2, < 1% Al2O3 þ SiO2) were cut
Zr-OH bonds, releasing the strain that stabilizes the tetragonal with a cutting machine (Precision Diamond Wire Saw, Well,
phase, thus promoting the t–m phase transformation.9 The Switzerland) into discs (diameter: 25 mm, thickness: 3 mm)
phase transformation starts at the surface and propagates under copious water (n ¼ 216). The specimens were
into the interior.9 According to Lange et al., the water reacts sintered in a programmable oven (Cercon heat, DeguDent,
with Y2O3 at the surface of the zirconia and clusters of the sinter temperature was set to 1350 C for 2 h). Other
Y(OH)3 are formed, leading to the depletion of stabilizers specimens were sintered at 1450 and 1600 C for 2 h
from the grains and the spontaneous t–m transformation with heating and cooling rates of 10 C/min (Nabertherm,
takes places in the grains.10 Yoshimura et al. explained the Switzerland). The specimens were divided randomly into six
degradation of Y-TZP with the annihilation of oxygen vacan- main groups and each main group was divided into four
cies and about 60% of these vacancies are occupied by OH– subgroups (see Table I). The specimens were abraded with
anions.14 The OH– anions would migrate faster than O2– different air-borne particles (50 and 110 lm alumina,
because of less charge and similar size.14 These authors did 30 and 110 lm silica-coated alumina particles) at a pres-
not accept the depletion of yttrium as the result of Y(OH)3 sure of 2.5 bar for 15 s/cm2 from a distance of 10 mm
formation or the dissolution of Y3þ during hydrothermal (LEMAT NT4, Wassermann, Germany).
treatment. According to Hughes et al., the generation of O2– is After air abrasion, all specimens were first air-blown at
responsible for the rapid annealing of surface oxygen vacan- a pressure of 3 bar for 1 min and then ultrasonically
cies.15 The penetration of O2– into the lattice destabilizes the cleaned in 99.8% isopropanol (Merck, Darmstadt, Germany)
t-phase and the transformation or nucleation of the m-phase for 20 min (Brasonic Ultrasonic Cleaner, Danbury) at a
domains occurs.15 Oxygen vacancies play an important role in frequency of 42 kHz.
the stabilization of the metastable tetragonal and cubic struc- The acceleration aging of specimens was performed
ture.16 Fabris et al. employed a self-consistent tight-binding under autoclave conditions of 134 C and a water vapor
model to constrain that the stabilization of the cubic and tet- pressure of 2.3 bar for a duration of 2 h (GETINGE industriv
ragonal structure could be achieved by introducing oxygen 5 SE-471 31 Skärhamn, Sweden).
vacancies into the zirconia lattice.17 Guo related the LTD with Ceramic veneering simulation was performed in an oven
annihilation of oxygen vacancies that stabilize the metastable (Austromat M, Dekema, Freilassing, Deutschland) at 830 C
tetragonal phase through the OH– group.18 Chevalier proposes for 1 min. The heating rate was set to 55 C/min. Additional
that the filling of oxygen vacancies with O2– and not OH– ions ceramic veneering simulations were conducted at 820, 810,
may be responsible for low-temperature degradation of zirco- and 800 C. The holding time for these temperature treat-
nia.6 The O2– ions originate from the dissociation of water on ments was 1 min and the heating rate was 45 C/min.
the surface of Y-TZP ceramic.6
The processes of low-temperature degradation (LTD) Phase transformation
can effectively be simulated at 134 C, under a water vapor Tetragonal (t)–monoclinic (m) phase transformations of Y-
pressure of 2 bar. One hour of this treatment corresponds TZP ceramics were analyzed with X-ray powder diffraction
roughly to 4 years of aging in human environment.7 technique (Bruker, AXS D8 Advance) using Cu Ka1 X-rays.

448 HALLMANN ET AL. LOW-TEMPERATURE DEGRADATION


ORIGINAL RESEARCH REPORT

TABLE I. Preparation of Specimens (72 Groups With Each n ¼ 3 Specimens)


Specimens Treatment

Y-TZP sintered at 1350, Main group a


1450, and 1600 C Subgroups:
a0 control specimens
a00 aged specimens
a000 aged specimens þ first ceramic veneering stimulation
a0000 aged specimens þ fourth ceramic veneering stimulations
Main group b
Subgroups:
b0 abraded specimens with 50 lm alumina particles
b00 abraded specimens þ aging
b000 abraded and aged specimens þ first ceramic veneering stimulation
b0000 abraded and aged specimens þ fourth ceramic veneering stimulations
Main group c
Subgroups:
c0 abraded specimens with 110 lm alumina particles
c00 , c000 , c0000 the specimens are treated as specimens of main group b
Main group d
Subgroups:
d0 abraded specimens with 30 lm silica-coated alumina particles
d00 , d000 , c0000 ; the specimens are treated as specimens of main Group b
Main group e
Subgroups:
e0 abraded specimens with 110 lm silica-coated alumina particles
e00 , e000 , c0000 the specimens are treated as specimens of main group b
Main group f
Subgroups:
f0 abraded specimens with 110 lm alumina particles followed by 110 lm silica-coated
alumina particles (Rocatec system)
f00 , f000 , f0000 the specimens are treated as the specimens of main group b

Ceramic veneering simulation was performed at 830 C with heating rate of 55 C/min and 820, 810, and 800 C with heating rate of 45 C/min.
The holding time was 1 min for each temperature.

The diffraction profiles were acquired from 25.5 to 29 (2h) ance, floating-column ion gun, and an electron neutralizer
with a step size of 0.001 and counting time of 2 s/step, were used for charge compensation. The residual pressure
and from 25 to 37 (2h) with a step size of 0.006 and in the analysis chamber was below 5  10–7 Pa. The spec-
counting time of 2 s/step. trometer was calibrated according to ISO 15472:2001 with
Raman spectroscopy was additionally employed to inves- an accuracy of 60.1 eV.
tigate the t–m phase transformation. High-resolution AFM
and confocal Raman measurements were conducted on a
combined confocal Raman system and AFM instrument Surface topography
(Spectra Ntegra upright, NTMDT, Russia) equipped with a The topography of the Y-TZP ceramic surface before and af-
100  0.7 NA objective (Mitutoyo, Japan), a 632.8 nm HeNe ter aging was investigated using AFM and FESEM. The
laser (up to 3 mW), a 100  100  10 lm piezo scanner, images were acquired using a BRUKER N8 RADOS scanning
and non-contact cantilever-based AFM probes (ATEC-NC, probe microscope as follows:
NanoSensors, Switzerland) for atomic force microscopy. • Aged specimens which were sintered at 1350 and
1600 C: N8 RADOS equipped with NANOS 806, 80  80
lm2 maximal scan area, 6 lm Z-range, active vibration
Surface elemental composition
isolation, acoustic enclosure.
The surface chemistry of the specimens was investigated by
• Sintered specimens at 1350 and 1600 C: N8 NEOS
means of X-ray photoelectron spectroscopy (XPS). The anal-
equipped with NANOS 806, (80  80 lm2) maximal scan
yses were performed with a PHI Quantera SXM (ULVAC-PHI,
area, 6 lm Z-range.
Chanhassen, MN) spectrometer. All XPS spectra were
• Measurement Parameters:
collected at an emission angle of 45 using a beam size of
* Dynamic (intermittent contact) mode.
200 lm with a power of 50 W in the constant–analyzer–
* Free amplitude was varied between 10 and 170 nm.
energy (CAE) mode. High-resolution spectra were acquired
* Damping set point was varied 58 and 70%.
using a pass energy of 55 eV and a step size of 0.1 eV (full-
* Scan speed: 0.5 line per second.
width-at-half-maximum (FWHM) of the peak height for Ag
3d5/2 0 0.64 eV). Survey spectra were collected with a pass The phase images were acquired simultaneously with
energy of 280 eV and a step size of 1 eV. A high-perform- the topography images.

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | FEB 2012 VOL 100B, ISSUE 2 449
FIGURE 1. XRD patterns of Y–TZP specimens before and after aging at the range of 27.5–29 (2y). [Color figure can be viewed in the online
issue, which is available at wileyonlinelibrary.com.]

FESEM experiments were carried out with a Carl Zeiss t–m phase transformation was evident for all the abraded
Supra 50 VP FESEM (Oberkochen, Germany). The accelera- specimens sintered at 1600 C in comparison with non-
tion voltage of the cathode was 5–30 kV. All specimens abraded specimens [Figure 1(c,f)].
were palladium gold coated before FESEM examinations. Figure 2 reveals the effect of low-temperature degrada-
tion (LTD) at 134 C under a water vapor pressure of 2.3
RESULTS bar of Y-TZP ceramic sintered at different temperatures over
Phase identification a range of diffraction angles from 25 to 39 (2h). The X-ray
Figure 1 displays the X-ray diffraction pattern (XRD) of Y- diffraction patterns of non-aged and aged specimens sin-
TZP sintered at various temperatures (1350, 1450, and tered at 1350 C are similar and clearly indicate the missing
1600 C) over the range 27.5–29 (2h) before and after of a t–m phase transformation [Figure 2(a)]. The effects of
aging. The profiles are zoomed in this range because the accelerated aging on the t–m phase transformation for Y-
t–m phase transformation can visually be identified. For TZP ceramic sintered at 1450 and 1600 C are shown in Fig-
specimens sintered at 1350 C, the t–m phase transformation ure 2(c,e). Compared to the non-aged Y-TZP specimens, the
after aging for 2 h at 134 C and under water vapor pres- aged Y-TZP specimens sintered at 1600 C contains the high-
sure of 2.3 bar, was not observed [Figure 1(a), plot a00 ], est concentration of the monoclinic phase. The reverse m-t
whereas the specimens sintered at 1450 and 1600 C phase transformation took place after ceramic veneering
showed the t–m transformation [Figure 1(b,c), plots a00 ]. simulation. The intensity of the m(11-1) peak in this case
After air abrasion with different air-borne particles, a decreased accompanied by an increase of the t(111) peak
detectable monoclinic peak with orientation m(11-1) was intensity. However, a small proportion of monoclinic phase
discernable [Figure 1(a–f), plots; b0 –f0 ]. The intensity of this seems to be present in both cases, but for the specimen sin-
peak was increased after accelerated aging for abraded tered at 1600 C the amount of non-transformed monoclinic
specimens [Figure 1(a–f), plots b00 –f00 ]. In contrast, the inten- phase was higher than for the specimen sintered at 1450 C
sity of these peaks decreased after the ceramic veneering [Figure 2(c,e) and plots a0 , a00 , and a000 ]. After air abrasion,
simulation at 830 C for 1 min. [Figure 1(b–f), plots a000 –f000 ]. the monoclinic peak m(11-1) and a remarkable t(111) peak
The highest extent of t–m phase transformation was broadening accompanied by the reversed intensity of the
observed for Y-TZP sintered at 1600 C. A retardation of the tetragonal t(002) and t(020) peaks were observed in all

450 HALLMANN ET AL. LOW-TEMPERATURE DEGRADATION


ORIGINAL RESEARCH REPORT

FIGURE 2. XRD pattern of Y–TZP specimens before and after aging at the range of 25–37 (2y). [Color figure can be viewed in the online issue,
which is available at wileyonlinelibrary.com.]

XRD patterns. After accelerated aging, the intensity of m(11- peaks was evident after the fourth ceramic veneering
1) peak was only slightly increased for the abraded speci- stimulation.
mens sintered at 1350 C followed by the specimens sin- Figure 3 presents Raman spectra of Y-TZP ceramic. It is
tered at 1450 C, whereas the specimens sintered at 1600 C evident from the Raman spectra that major bands are
resulted significantly higher t–m phase transformation [Fig- located at 141, 255, 315, 462, and 641 cm–1. The peak at
ure 2(b,d,f,g)]. The intensity of the m(11-1) peak and 635 cm–1 has a shoulder at 603 cm–1. These peaks are char-
the broadening of the t(111) peak were reduced for all acteristic for the tetragonal phase of Y-TZP ceramic.19–22 The
specimens after the firths ceramic veneering stimulation. peaks at 175 and 373 cm–1 are characteristic for a mono-
After that no intensity change of the m(11-1) and t(111) clinic structure.23–25 The weak peak at 550 cm–1 likewise is

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | FEB 2012 VOL 100B, ISSUE 2 451
3d core level was not observed. Such a positive shift would
be indicative for the formation of M-OH bonds. No remark-
able change of the binding energy for the aluminium ion
was detected from the Al 2p spectra [Figure 4(e)]. Figure
4(f) shows the energy binding of the O 1s core level before
and after aging. The peak at low binding energy is assigned
to oxygen in the metal-O-metal bond. The shoulder at higher
binding energy 530.4–531 eV (Table III) could be assigned
to the OH bond, absorbed water, or oxygen–carbon bonds.

Surface topography
Figure 5 shows the AFM images of surface morphology of Y-
TZP ceramic, which was sintered at 1350 and 1600 C
before and after aging. No change of the surface morphology
and/or roughness of non-aged and aged Y-TZP ceramic, sin-
tered at 1350 C, was not observable. The roughness for
FIGURE 3. Raman spectra of Y-TZP specimens before and after aging.
[Color figure can be viewed in the online issue, which is available at non-aged and aged Y-TZP ceramic surface was 10.37 and
wileyonlinelibrary.com.] 8.97 nm, respectively. Figure 5(a0 –d0 ) presents AFM images
of the surface morphology of Y-TZP ceramic sintered at
attributed to the monoclinic phase (Table II). Comparing plots 1600 C, indicating that a change of grain topography
3a, b and c reveals that the distribution of the monoclinic occurred as the grain surfaces were not flat after acceler-
phase for aged Y-TZP sintered at 1600 C was not homogenous. ated aging. The monoclinic phase precipitated as dispersed
lenses in the tetragonal matrix phase. The roughness of Y-
TZP ceramic surface increased after accelerated aging to
Surface elemental composition 9.75 and 15.3 nm for non-aged and aged Y-TZP ceramic
The survey XPS spectra [Figure 4(a)] revealed that the ele- surfaces, respectively.
mental composition of the Y-TZP ceramic surface did not Figure 6 displays the FESEM images of non-aged and
change after accelerated aging at a temperature of 134 C aged Y-TZP ceramics, which were sintered at different tem-
under a water pressure of 2.3 bar for 2 h. The C 1s peak of peratures. The average grain size was small (about 0.21
all the specimens resulted from an overlaying contaminating lm) and not homogenous for the specimens sintered at
hydrocarbon layer, which was unavoidable during XPS mea- 1350 C. The same applied to the specimens sintered at
surement [Figure 4(b)]. Figure 4(c) compares the Zr 3d 1450 and 1600 C, with a grain size of 0.3 and 0.72 lm,
spectra obtained before and after accelerated aging for the respectively. After accelerated aging, the grains of specimens
Y-TZP sintered at three different temperatures. The binding sintered at temperature1350 and 1450 C appeared not to
energies of Zr 3d3/2 and Zr 3d5/2 were 183.2–183.6 eV and be uplifted in contrast to the specimen sintered at 1600 C.
180.8–182.2 eV, respectively (Table III). The binding energy The surface of aged Y-TZP sintered at 1600 C was damaged
of the Zr 3d core was shifted for 0.3 eV to negative values as a result of volume increasing during t–m phase transfor-
for aged specimens sintered at 1600 C. The Y 3d spectra mation and holes were created [Figure 6(f)]. Figure 6(h)
are shown in Figure 4(d). The binding energy of Y 3d3/2 shows the precipitation of the monoclinic phase as dis-
and Y 3d5/2 varied from 158.0 to 158.4 eV and 156.0 to persed lenses. The EDS measurements did not reveal any
156.5 eV, respectively. The Y 3d core peaks were shifted for change in the yttrium concentration before and after accel-
0.4 eV to negative values for the samples sintered at erated aging of Y-TZP ceramic sintered at different tempera-
1600 C. A positive shift of binding energy for Zr 3d and Y tures. Screening of elemental composition for grain surface

TABLE II. Wave Number (in cm1) of Raman Vibration Bands for the Tetragonal and Monoclinic Structure of Y-TZP Ceramics

Specimens Tetragonal Phase


 sh
Sintered specimen at 1350 C 141 254 314 458 601 637
Aged specimen (sintered at 1350 C) 141 254 314 458 601 sh
637
Sintered specimen at 1450 C 144 258 318 462 604 sh
641
Aged specimen (sintered at 1450 C) 143 257 316 462 604 sh
640
Sintered specimen at 1600 C 144 257 318 462 604 sh
640
(a) Aged specimen (sintered at 1600 C) 142 256 317 460 604 sh
639
(b) Aged specimen (sintered at 1600 C) 144 258 319 464 604 sh
640
(c) Aged specimen (sintered at 1600 C) 141 256 323 466 634
Specimens Monoclinic Phase
(b) Aged specimen (sintered at 1600 C 175 187 377
(c) Aged specimen (sintered at 1600 C) 175 187 374 549

452 HALLMANN ET AL. LOW-TEMPERATURE DEGRADATION


ORIGINAL RESEARCH REPORT

FIGURE 4. XPS spectra of Y-TZP ceramic surfaces before and after aging. [Color figure can be viewed in the online issue, which is available at
wileyonlinelibrary.com.]

and boundary concentration changes showed a homogenous of these stresses depend of the sintering temperature. The
distribution of yttrium across the surface and no concentra- slight increasing of the extent of the t–m phase transforma-
tion gradient in the grain-boundary region was observed tion observed for aged ceramic sintered at 1450 C implies
(Table IV). that this ratio was changed in favor of tensile stress in com-
parison to the ceramic sintered at 1350 C.
DISCUSSION Some groups of Y-TZP ceramic surfaces in this study
Low-temperature degradation (LTD) of dental zirconia by were abraded before accelerated aging. The aim was to
annealing in water vapor is dependent on several factors investigate the effect of micro-cracks that were introduced
such as crystal structure, grain size, the amount of dopant, on the Y-TZP ceramic surface during air abrasion. There are
residual stress. In this study, the stress-induced t–m phase studies which relate the propagation of the t–m phase
transformation of Y-TZP started at a sintering temperature transformation to the creation of micro-cracks during accel-
of 1450 C. The highest transformation rate observed for Y- erated aging.26,27 Micro-cracks served as preferential paths
TZP ceramic sintered at 1600 C can be explained by for water diffusion inside the bulk of the Y-TZP ceramic.28
increased tensile stresses, as larger grain sizes correspond The air abrasion induced the t–m phase transformation,
to a larger tensile stress (this work will be soon submitted). plastic deformations, flaws, grooves, cracks, and impinging
The grains of Y-TZP sintered at 1350 C were more spherical of air-borne particles on the surface of the ceramic. The
than the grains of Y-TZP sintered at 1450 C, while the grains compressive stresses created during the air abrasion serve
of Y-TZP sintered at 1600 C were flattened and enlarged. to inhibit micro-crack extension.4 Our study showed that
The stress distribution is more uniform for the spherical the air abrasion was responsible for the retardation of diffu-
grains than for the flat grains. The tensile and compressive sion controlled t–m phase transformation for Y-TZP ceramic
stresses are present in the sintered ceramic, but the ratios sintered at 1600 C. The slight increasing of the t–m phase

TABLE III. Binding Energies (in eV) of Zr 3d, Y 3d, Al 2p, and O 1s Core Levels

Specimens Zr 3d3/2 Zr 3d5/2 Y 3d3/2 Y 3d5/2 Al 2p O 1s O 1s



Sintered specimen at 1350 C 183.2 180.8 158.0 156.0 73.0 528.6 530.0
Aged sintered specimen at 1350 C 183.2 180.8 158.0 156.0 73.1 528.8 530.5
Sintered specimen at 1450 C 183.5 181.2 158.2 156.2 73.2 528.9 530.8
Aged sintered specimen at 1450 C 183.4 181.1 158.0 156.0 73.1 528.7 531.0
Sintered specimen at 1600 C 183.6 181.2 158.5 156.5 73.5 529.1 531.2
Aged sintered specimen at 1600 C 183.3 180.9 158.1 156.2 73.5 528.8 531.2

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | FEB 2012 VOL 100B, ISSUE 2 453
454
HALLMANN ET AL.
FIGURE 5. AFM images of Y-TZP ceramic surfaces before and after aging for specimens sintered at 1350 and 1600 C. [Color figure can be viewed in the online issue, which is available at
wileyonlinelibrary.com.]

LOW-TEMPERATURE DEGRADATION
ORIGINAL RESEARCH REPORT

FIGURE 6. FESEM images of the specimens sintered at 1350, 1450, and 1600 C before and after aging.

transformation for aged Y-TZP sintered at 1350 and 1450 C The theoretical atomistic study of Kushima et al. con-
after air abrasion in comparison with non-abraded Y-TZP ducted to explain the mechanism of oxygen anion transport
ceramic can be explained by the different stresses which in anion conducing ceramic reveled that the biaxial tensile
predominate in these ceramics before and after air abrasion. lattice stress is a very important factor for the migration of
After air abrasion, this ratio changed in favor of tensile oxygen vacancies by the hopping mechanism.28 The energy
stresses, which were responsible for contrasting behaviors barrier of migrating vacancies is affected by the migration
of sintered and abraded ceramics sintered at these tempera- space, cation-oxygen bond strength, and additional factors.
tures. The grain size of abrasive particles did not play a cru- The increase of cation–cation distances and the weakness of
cial role on LTD of Y-TZP sintered at different temperatures. the cation–oxygen bond strength decrease the migration
The binding energy shift of a core-electron is related to energy barrier of vacancies. The oxygen ions migrate after
the change of the chemical environment of the element. The the same mechanism as vacancies. These authors found that
negative shift of the Zr 3d binding energy for the specimen the macroscopic oxygen diffusivity increases exponentially
sintered at 1600 C after aging in this study indicates a in yttria stabilized zirconia up to a critical value of biaxial
change of the chemical state for the zirconium ion. Binding tensile stress, especially at lower temperatures. Beyond this
energy reduction of a metal ion is an indication for an critical value of lattice stress, the migration of oxygen
increasingly covalent bond formation,29 whereas the forma- decreases due to local relaxation.28 The higher t–m phase
tion of a Zr-OH bond shows a higher binding energy than transformation rate during accelerated aging observed for
that of ZrO2 for the same oxidation state.30 It may therefore sintered Y-TZP at 1600 C in this study confirms their
be suggested that the partial reduction of Zr4þ to Zr3þ took calculations.
place and the conversion of ZrO2 to suboxides ZrOx There is a debate32 about the species that are responsi-
occurred during aging process.31 The lack of a positive shift ble for stress induced t–m phase transformation of Y-TZP ce-
for the zirconium and yttrium ions shown in this study indi- ramic. The backward phase transformation m-t observed
cated that Zr-OH and Y-OH bonds were not introduced into during the first ceramic veneering simulation provides infor-
the Y-TZP lattice. The lack of concentration gradients for yt- mation about these species. This transformation took place
trium ions at the grain boundaries does not support the very fast at the beginning and thereafter the reversible reac-
theory of LTD occurring through the depletion of yttrium tion is retarded. The full reversible reaction took place at
ions from the lattice of Y-TZP ceramic. higher temperatures only. This implies that the species that

TABLE IV. Elemental Compositions in wt % of Non-aged and Aged Y-TZP Ceramic in Terms of Zr and Y
Zr (wt %) Y (wt %)
Grain Boundary Grain Boundary

Sintered zirconia at 1350 C 78.55 78.70 7.08 7.12
Aged zirconia (sintered at 1350 C) 78.37 77.97 6.85 6.91
Sintered zirconia at 1450 C 78.63 78.48 7.04 7.35
Aged zirconia (sintered at 1450 C) 79.35 79.27 7.12 7.21
Sintered zirconia at 1600 C 78.45 78.31 7.47 7.78
Aged zirconia (sintered at 1600 C) 78.69 78.96 8.53 7.78

JOURNAL OF BIOMEDICAL MATERIALS RESEARCH B: APPLIED BIOMATERIALS | FEB 2012 VOL 100B, ISSUE 2 455
are involved in this transformation moved very fast at first 6. Chevalier J, Gremillard L, Virkar VA, Clarke RD. The tetragonal-
and after that, due to the alteration of the lattice stress, the monoclinic transformation in zirconia:Lessons learned and future
trends. J Am Ceram Soc 2009;92:1901–1920.
migration barrier energy increased. The surface of Y-TZP 7. Gremillard L, Chevalier J, Epicier T, Deville S, Fantozzi G. Model-
serves as a catalyst for the dissociation of water into O2– ing the ageing kinetics of zirconia ceramics. J Euro Ceram Soc
and Hþ ions. Hþ ions, therefore, probably play a crucial role 2004;24:3483–3489.
8. Chevalier J, Cales B, Drouin MJ. Low-temperature aging of Y-TZP
on LTD of Y-TZP. For a full understanding of the LTD mecha-
ceramics. J Am Ceram Soc 1999;82:2150–2154.
nism, additional theoretical calculations at an atomistic scale 9. Sato T, Shimada M. Transformation of yttria-doped tetragonal ZrO2
are required. The biaxial tensile stress is one of the factors polycrystals by annealing in water. J Am Ceram Soc 1985;68:356–359.
which control the diffusion of vacancies and oxygen into the 10. Lange FF, Dunlop LG, Davis IB. Degradation during aging of
transformation-toughened ZrO2-Y2O3 materials at 250 C. J Am
lattice of Y-TZP. The role of hydrogen on the LTD mecha- Ceram Soc 1986;69:237–240.
nisms is not yet clear and request complementary measure- 11. Kelly RJ, Denry I. Stabilized zirconia as a structural ceramic: an
ments which are beyond the scope of this article. overview. Dental Mater 2008;24:289–298.
12. Zhang Y, Lawn RB, Malament AK, Thompson PV, Rekow DE.
Damage accumulation and fatigue life of particle-abraded
ceramics. Int J Prosthodont 2006;19:442–448.
CONCLUSION 13. Scherrer S.S, Lorente CM, Vittecoq E, Mestral F, Griggs AJ, Wis-
kott AWH. Fatigue behavior in water of Y-TZP zirconia ceramics
1. The degradation of Y-TZP ceramic at the condition of after abrasion with 30 lm silica-coated alumina particles. Dental
accelerated aging is strongly dependent on the grain Mater 2011;27:e28–e42.
14. Yoshimura M, Noma T, Kawabata K, Somiya S. Role of H2O on the
sizes.
degradation process of Y-TZP. J Mater Sci Lett 1987;6:465–467.
2. When the grain size has achieved the critical value, which 15. Hughes EA, John SH, Kountouros P, Shubert H. Moisture sensi-
is about 0.3 lm, the t–m phase transformation occurred. tive degradation in TiO2-Y2O3-ZrO2. J Euro Ceram Soc 1995;15:
3. Y-TZP sintered at 1600 C resulted the highest rate of deg- 1125–1134.
16. Kawata K, Maekawa H, Nemoto T, Yamamura T. Local structure anal-
radation (average grain size was about 0.72 lm). ysis of YSZ by Y-89 MAS-NMR. Solid State Ion 2006;177:1687–1690.
4. Flaws and cracks which were induced on the surface of 17. Fabris S, Paxton TA, Finnis WM. A stabilization mechanism of zirconia
Y-TZP through the air abrasion process were not an im- based on oxygen vacancies only. Acta Mater 2002;50:5171–5178.
18. Guo X. On the degradation of zirconia ceramics during low-tem-
portant factor for the LTD. The size of air-borne particles
perature annealing in water or water vapour. J Phys Chem Solids
was not a crucial factor on the LTD. The retardation of 1999;60:539–546.
LTD caused by air abrasion was observed for sintered 19. Li M, Feng Z, Ying P, Xin, Q, Li C. Phase transformation in the
specimens at 1600 C. surface region of zirconia and doped zirconia detected by UV
Raman spectroscopy. Phys Chem Chem Phys 2003;5:5326–5332.
5. The formation of Y-OH and Zr-OH bonds was not 20. Merle T, Guinebretiere R, Mirgorodsky A, Quintard P. Polarized
observed. The concentration gradient of yttrium ions on Raman spectra of tetragonal pure ZrO2 measured on epitaxial
the surface of grains or on the grain boundaries did not films. Phys Rev B 2002;65:144302-1–144302-6.
take place. 21. Quintard EP, Barberis P, Mirgorodsky PA, Mejean MT. Comparative lat-
tice-dynamical study of the raman spectra of monoclinic and tetrago-
6. A stable Y-TZP can be achieved from a sintering tempera- nal phases of zirconia and hafnia. J Am Ceram Soc 2002;85:1745–1749.
ture ranging between 1350 and 1450 C. 22. Lopez FE, Escribano SV, Panizza M, Carnasciali MM, Busca G.
Vibrational and electronic spectroscopic properties of zirconia
powders. J Mater Chem 2001;11:1891–1897.
ACKNOWLEDGMENTS 23. Daramola AD, Muthuvel M, Botte GG. Density functional theory
The authors thank Prof. N. Spencer and Mr. F. Mangolini from analysis of Raman frequency modes of monoclinic zirconium ox-
ide using Gaussian basis sets and isotopic substitution. J Phys
the Surface & Science Technology, Department of ETH Zürich Chem B 2010;114:9323–9329.
for XPS measurements, Dr. Igor Nemeth Bruker Nano GmbH 24. Kim KB, Hamaguchi H. Mode assignments of the Raman spec-
Berlin for AFM measurements, Dr. L Völkl of Degudent for the trum of monoclinic zirconia by isotopic exchange technique. Phys
Status Solidi B 1997;203:557–563.
Y-TZP blanks, Prof. Gauckler from the Nonmetallic Inorganic
25. Maczka M, Lutz GTE, Verbeek JH, Oskam K, Meijerink A, Hanuza
Materials, Swiss Federal Institute of Technology, Department J, Stuivinga M. Spectroscopic studies of dynamically compacted
of Materials of ETH Zürich for his advice. monoclinic ZrO2. J Phys Chem Solids 1999;60:1909–1914.
26. Chevalier J. What future for zirconia as a biomaterial? Biomateri-
als 2006;27:535–543.
27. Lawson S. Environmental degradation of zirconia ceramics. J
REFERENCES Euro Ceram Soc 1995;15:485–502.
1. Lung KYC, Kukk E, Hägerth T, Matinlinna PJ. Surface modification 28. Kushima A, Yildiz B. Oxygen ion diffusivity in strained yttria stabi-
of silica-coated zirconia by chemical treatments. Appl Surf Sci lized zirconia: where is the fastest strain? J Mater Chem 2010;20:
2010;257:1228–1235. 4809–4819.
2. Piconi C, Maccauro G. Zirconia as a ceramic biomaterial. Biomate- 29. Barr TL. Recent advances in x-ray photoelectron spectroscopy sty-
rials 1999;20:1–25. dies of oxides. J Vacuum Sci Technol A 1991;9:1793–1805.
3. Sato H, Ban S, Hashiguchi M, Yamasaki Y. Effect of autoclave 30. Li SY, Wong CP, Mitchell RAK. XPS investigations of the interac-
treatment on bonding strength of dental zirconia ceramics to tions of hydrogen with thin films of zirconium oxide II. Effect of
resin cements. J Ceram Soc Japan 2010;118:508–511. heating a 26 Å thick film after treatment with a hydrogen plasma.
4. Kosmač T, Oblak Č, Marion L. The effects of dental grinding and Appl Surf Sci 1995;89:263–269.
sandblasting on ageing and fatigue behavior of dental zirconia (Y- 31. Kumari L, Li ZW, Xu MJ, Leblanc MR, Wang ZD,. Li Y, Guo H, Zhang J.
TZP) ceramics. J Euro Ceram Soc 2008;28:1085–1090. Controlled hydrothermal synthesis of zirconium oxide nanostructures
5. Moser ME, Keller AB, Lienemann P, Hug P. Surface analytical and their optical properties. Crystal Growth Design 2009;9:3874–3880.
study of hydrothermally treated zirconia ceramics. Fresenius 32. Duong T, Limarga MA, Clarke RD. Diffusion of water species in
J Anal Chem 1993;346:255–260. yttria-stabilized zirconia. J Am Ceram Soc 2009;92:2731–2737.

456 HALLMANN ET AL. LOW-TEMPERATURE DEGRADATION

Você também pode gostar