Sei sulla pagina 1di 10

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

Research Paper

Factors affecting the mechanical behavior of Y-TZP


Ferhan Egilmeza,n, Gulfem Erguna, Isil Cekic-Nagasa, Pekka K. Vallittub,
Lippo V.J. Lassilab
a

Gazi University Faculty of Dentistry, Department of Prosthodontics, Ankara, Turkiye


Institute of Dentistry, Department of Biomaterials Science and Turku Clinical Biomaterials Centre TCBC,
University of Turku, Turku, Finland
b

ar t ic l e in f o

abs tra ct

Article history:

Aim: The aim was to evaluate the inuence of sandblasting with various propulsion pressures

Received 7 December 2013

on the phase transformation, exural strength and Weibull modulus of a yttria stabilized

Received in revised form

tetragonal zirconia polycrystalline ceramic (Y-TPZ). In addition, the susceptibility of cyclic

6 May 2014

loading and low-temperature degradation under two different conditions (chemical and

Accepted 10 May 2014

thermal aging) was investigated.

Available online 15 May 2014

Materials and methods: The specimens [105 bar-shaped specimens (19.3  14.5  1.3 mm3)] were

Keywords:

equally divided into seven groups (n15) according to the test protocols. The specimens in

Y-TZP zirconia

control group received no surface treatment (Group A). Groups B1, B2 and B3 were airborne-

Surface treatments

particle abraded with 110 m Al2O3 particles at 200 kPa, 400 kPa and 600 kPa (2 bar, 4 bar and

Cyclic fatigue

6 bar) pressure, respectively. Group C was submitted to mechanical cyclic loading under 200 N

Thermal and chemical degradation

20,000 times sinusoidal loading/unloading at a frequency of 10 Hz between 10% and 100% load

Flexural strength

in distilled water at 37 1C. Group D was submitted to thermal degradation in an autoclave at


134 1C under additional 200 kPa pressure for 5 h. Group E was immersed in 4% acetic acid at
8075 1C for 168 h as chemical degradation testing. Following each treatment protocols, the
three-point exure test was used to calculate the exural strength. Additionally, X-ray
diffraction analysis was used to estimate the relative amount of monoclinic phase. The
reliability of strength was assessed through the Weibull distribution. Statistical analysis was
conducted with one-way ANOVA and Tukey's pairwise multiple comparisons. The treated and
fractured surfaces were observed with SEM.
Results: Statistical analysis revealed that there were signicant differences among the exural
strength results of all tested groups (F4.510, po0.05). Group B2 demonstrated the highest
strength values, whereas Group E showed the lowest (625.867123.57; 466.56791.50, respectively). Weibull moduli of all tested groups were statistically signicant and ranging from 4.3 to
8.3. A greater amount of monoclinic phase was determined in the specimens of D, E and B3
groups (25.43%, 20.89% and 19.71%, respectively). However, lower amount was observed in
groups A, B1 and B2 (10.02%, 13.35% and 15.19%, respectively).

n
Correspondence to: MutlukentMah, 10. Cadde 2065. Sk. No:15 Beysukent, Ankara, Turkey. Tel.: 90 312 203 41 92;
fax: 90 312 223 92 26.
E-mail addresses: ferhanegilmez@gmail.com, fegilmez@gazi.edu.tr (F. Egilmez).

http://dx.doi.org/10.1016/j.jmbbm.2014.05.013
1751-6161/& 2014 Elsevier Ltd. All rights reserved.

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

79

Conclusions: The present study suggested that exural strength of zirconia was signicantly
decreased by chemical degradation. In addition, surface conditioning, cyclic fatigue and
thermal, chemical degradation conditions signicantly changed the structural reliability of
the material's strength.
& 2014 Elsevier Ltd. All rights reserved.

1.

Introduction

The development of advanced dental ceramics has led to the


application of partially stabilized zirconia, which has gained
increased popularity in contemporary dentistry due to its
high biocompatibility, low bacterial surface adhesion, high
exural strength, toughness due to transformation toughening mechanism, and cosmeticesthetic properties (Egilmez
et al., 2013; Guess et al., 2012). Pure zirconium dioxide
possesses three crystallographic forms that are stable at a
different range of temperatures under atmospheric pressure.
The monoclinic phase is stable up to approximately 1170 1C.
The structure is tetragonal between 1170 1C and 2370 1C and
cubic above 2370 1C and up to the melting point (Denry and
Kelly, 2008). Nonetheless, the tetragonal phase, which is
essentially metastable, may transform into monoclinic state.
Alloying pure zirconia with stabilizing oxides such as CaO,
MgO, Y2O3, or CeO2 has been shown to maintain the tetragonal structure at room temperature (Denry and Kelly, 2008).
However, the phase transformation may be triggered by the
external stresses such as thermal aging (Chevalier et al.,
1999), grinding (Kosmac et al., 1999; Garvie et al., 1975) and
sandblasting (Porter and Heuer, 1977). This transformation
exhibits 34% volume expansion which induces compressive
stresses (Piconi and Maccauro, 1999), thereby closing the
crack tip and preventing further crack propagation (Porter
and Heuer, 1977). On the other hand, the deep surface aws
introduced by severe grinding and sandblasting act as stress
concentrators and may cause strength degradation (Green, 1983).
The other negative physical characteristic of Y-TZP is known as
aging or low-temperature degradation (LTD) (Chevalier et al.,
1999). This phenomenon initiates in isolated surface grains,
between which water is incorporated into the zirconia lattice
through dissolution of ZrOZr bonds and lling of oxygen
vacancies (Guo, 2004). After the surface is saturated with monoclinic phase, LTD proceeds into the bulk material (Yoshimura
et al., 1987). This process may reduce the strength, toughness,
and density of zirconia signicantly (Kim et al., 2009), a problem
that is further exacerbated by exposure of the restoration to
cyclic stresses such as chewing (Studart et al., 2007a). LTD may
also increase surface roughness of the zirconia which may cause
increased wear on antagonist teeth (Lughi and Sergo, 2010).
To assess the strength of brittle materials such as ceramics, exural strength is generally considered a meaningful
and reliable method (Pittayachawan et al., 2007). The failure
stresses of brittle materials are statistically distributed as a
function of the aw size distribution in the material (Bona
et al., 2003). Weibull statistics that shows the structural
reliability or probability of fracture of brittle materials is used
to determine this distribution (Pittayachawan et al., 2007).

The main parameters involved are the Weibull modulus (m)


and characteristic strength (s) (Weibull, 1951). A high Weibull
modulus indicates a smaller error range, a higher level of
structural integrity and potentially greater structural reliability of the material (Bona et al., 2003).
In addition to mechanical properties, the chemical
solubility of a dental ceramic plays a major role for material
selection. The acidity of uids or food substances in the oral
cavity may cause dissolution of a ceramic material, which in
turn may be the result of the loss of surface luster or
weakening of the material (Esquivel-Upshaw et al., 2001).
It is especially important in the clinical conditions where the
material is exposed to thermal and mechanical cycling in a
chemically active aqueous environment over long periods. In
these circumstances, the impact-induced aws may grow to
become stress intensiers, facilitating fracture initiation at
lower levels of applied stress (Kosmac et al., 2000); therefore,
it is important to know what stress a material can withstand
for material selection.
The aim of this investigation was to assess the inuence
of sandblasting with various propulsion pressures on the
phase transformation behavior, three-point exural strength
and Weibull modulus of a yittria stabilized tetragonal zirconia
polycrystalline ceramic (Y-TPZ). Additionally, the susceptibility
of cyclic loading and low-temperature degradation under two
different conditions adopted for testing the chemical solubility
and thermal aging was investigated. The following null hypotheses were tested: (1) different sandblasting pressure, cyclic
loading and chemical and thermal aging have no effect of the
exural strength and reliability of Y-TZP; (2) different sandblasting pressure, cyclic loading and chemical and thermal aging do
not increase the monoclinic phase content of Y-TZP.

2.

Materials and methods

2.1.

Preparation of the specimens

Several blocks of Lava Frame Y-TPZ ceramic (Lot#: 367366,


367367, 3M ESPE AG, Seefeld, Germany) were used to make
105 bar-shaped (25  20  1.5 mm3) specimens. The blocks
were cut with a diamond blade (Leco 330 CA 600 grit size
mounted on Struers Accutom 50, Copenhagen, Denmark).
No polishing procedure was performed after cutting the
ceramic blocks. The specimens were ultrasonically cleaned
(Quantrex 90, L&R Ultrasonics, Kearny, NJ) for 10 min in 961
ethanol and deionized water in sequence and air dried. The
dimensions of the specimens were determined using a digital
caliper (Liaoning MEC Group, Dalian, China) with an accuracy
of 0.01 mm. After that, they were sintered into Lava Therm

80

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

Sintering Furnace (3M ESPE, St. Paul, MN, USA) according to


the manufacturer's recommendation. The nal dimensions
of the specimens were 19.3  14.5  1.3 mm3. Then, they were
allocated randomly into 7 groups (n 15). Group codes are
given in Table 1. The specimens in group A assigned as
Control and received no surface treatment, LTD procedure
or cyclic loading.

2.2.

Surface treatment

One surfaces of the specimens in groups B1, B2 and B3 were


airborne-particle abraded with 110 m Al2O3 particles (Rocatec Pre, 3 M ESPE, St. Paul, MN) at a direction perpendicular to
the surface and from a distance of about 10 mm for 20s.
In group B1, airborne-particle abrasion was applied under
200 kPa; in group B2, 400 kPa and in group B3, 600 kPa
pressure. Air-abrasion was performed on the tensile side of
the specimens in relation to loading during the three-point
bending test.

2.3.

Fatigue testing

The specimens in group C were submitted to mechanical


cyclic loading using a Universal Testing Machine (Lloyd LR
30K Plus; Lloyd Instruments Ltd, Fareham, UK) with a threepoint bend test jig. Two hardened steel knife edges with a
radius of 0.870.01 mm formed the supports for the specimen.
The span between the supports was 13.070.01 mm. The
midpoint of each specimen was located and then placed
centrally between the supports. Cyclic loading was performed
under 200 N 20,000 times sinusoidal loading/unloading at a
frequency of 10 Hz between 10% and 100% load. The loading
was performed with a third steel knife edge with a radius of
0.870.01 mm perpendicular to the long axis of the bar in
distilled water at 37 1C. In order to avoid sliding movements
of the specimens, and to distribute the tension in a more
uniform way during cycling, a polyethylene strip (3 M ESPE)

(12 mm  10 mm) was placed on the compression side during


cyclic loading.

2.4.

Thermal and chemical degradation testing

Thermal degradation testing was performed for the specimens in group D and simulated in an autoclave (Cisa S.p.A.,
Pomezia, Rome, Italy) at 134 1C under additonal 200 kPa
pressure for 5 h.
For the chemical degradation testing, the specimens in
group E were washed with distilled water and dried before
immersing in 4% acetic acid. Each specimen was placed
separately in a glass-dish and 20 ml acetic acid solution was
added into each dish. Then they were stored at 8075 1C for
168h in a drying furnace (Precision Scientic, Virginia, USA).

Uniaxial exural strength test (three-point bend test)

2.5.

All experimental groups were subjected to monotonic uniaxial loading to determine the critical load for fracture. Same
specimen holder and Universal Testing Machine were used
for fatigue testing and monotonic loading. The load at a
crosshead speed of 0.5 mm/min was applied to the midpoint
between the supports by means of a third steel knife edge
across the 14.5-mm-wide face along a line perpendicular to
the long axis of the bar until fracture occurred. The test was
conducted at room temperature (2271 1C) with a relative
humidity of 7075%.
The following formula was used to calculate the uniaxial
exural strength [modulus of rupture (s)] of each specimen
(ISO 6872, 1995):
s

3Wl
2bd2

where W is the breaking load (N); l is the test span (mm); b is


the width of the specimen (mm); and d is the thickness of the
specimen (mm).

Table 1 Experimental groups (names and codes), mean exural strength and standard deviation values (SD),
characteristic strength (r0) and Weibull modulus (m).
Group
names

Control
200 kPa Al2O3
400 kPa Al2O3
600 kPa Al2O3
Cyclic loaded 2000
times
Autoclave
Acetic Acid

Group
codes

Flexural
strength7Std
deviation (MPa) (r)

Weibull
characteristic
strength (Std
error)
(MPa) (r0)

Respective
condence
ntervals (Cl) (95%)
of r0
Lower

Upper

Weibull
modulus (m)
(Std error)

A
B1
B2
B3
C

597.266795.831a
538.8857116.072a,b
625.8607123.571a
614.970797.015a
525.7087113.609a,b

638.423
581.679
671.345
655.758
570.088

(29.037)a
(24.878)b
(26.105)c
(23.973)c
(30.848)d

581.540
532.943
620.205
608.795
509.657

695.306
630.415
722.485
702.721
630.519

6.043
6.296
6.925
7.468
5.058

D
E

578.119779.725a,b
466.564791.494b

612.550 (21.814)e
506.803 (24.327)f

569.816
459.146

655.284
554.460

7.700 (1.483)b
5.281 (1.017)d

Different superscript letters indicate statistical difference inside the respective column.

(1.073)a
(1.342)a
(1.475)a,b
(1.509)b
(0.980)c

Respective
condence
ntervals (Cl)
(95%) of m
Lower

Upper

3.940
3.666
4.035
4.511
3.137

8.145
8.925
9.815
10.425
6.979

4.795
3.289

10.604
7.273

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

2.6.

Phase analysis by X-ray diffraction

X-ray diffraction analyses (APD 2000 Pro Generator, GNR


Analytical Instruments, Italy) were conducted to determine
quantitative relative amount of the tetragonal to monoclinic
phase of the fractured surfaces of the specimens (n 2/group).
Specimens were scanned with CuK1,2 (1.54060 ) and spectra
were collected in the 2 range 20401 at a step interval of 1 s
and step size of 0.021. Monoclinic peak intensity ratio was
calculated using the following equation given by Garvie and
Nicholson (1972):
Xm

Im111 Im111
Im111 Im111 It111

where It and Im is the integrated intensity area under the


peaks of the tetragonal (111), monoclinic (111) and monoclinic
(111) peaks around 301, 311, and 281, respectively. Experimental monoclinic volume content, Vm, was calculated with the
formula suggested by Toraya et al. (1984):
Vm

1:311:Xm
0:311:Xm 1

In addition, the thickness of the transformed surface layer


(TZD) of tested specimens was calculated using the X-ray
determination method proposed by Kosmac et al. (1999):
TZD

sin
1
ln
2
1 Xm

where 151 is the angle of reection, 0.0642 the absorption coefcient and Xm the relative monoclinic fraction
obtained from XRD analysis.

2.7.

SEM analysis

Two specimens from each group were prepared for SEM


analysis to evaluate the surface structure. The specimens
were sputter-coated (Polaron SC 502, Fisons Instruments, UK)
with gold and observed under SEM (JSM-6060 LV; JEOL Ltd.,
Tokyo, Japan) for the examination of representative zirconia
surfaces after uniaxial exural strength test.

2.8.

Statistical analysis

Statistical analysis was conducted with KolmogorovSmirnov


test to evaluate the normal distribution of variables. Then,
homogenity test (Levene test) was used to verify the homogeneity of variance. Mean uniaxial exural strength values
(MPa) of tested specimens in each group were analyzed by
one-way ANOVA and Tukey's pair-wise multiple comparisons with a signicance level of 5%. Analysis was undertaken
using Statistics Package for the Social Sciences (SPSS 16.0 for
Windows, Microsoft, USA).
The reliability of strength was assessed through the
Weibull distribution. The method proposed by Quinn and
Quinn (2010) was chosen to perform the Weibull regression
analysis. For each of the seven test groups, the stress values
were ranked in ascending order, i 1, 2, 3,, N, where N is the
total number of test specimens and i is the ith datum. Thus,
the lowest stress for each conguration represents the rst
value (i 1), the next lowest stress value is the second datum
(i2), etc., and the highest stress is represented by the Nth

81

datum. This enables a ranked probability of failure, Pf (si), to


be assigned to each datum according to the following formula:
Pf si

i0:5
N

Then, the Weibull analysis was performed for calculation


of cumulative fracture probability (Pf) as function of applied
stress (Quinn and Quinn, 2010). The form of the Weibull
equation (for the two parameter distribution) is shown as:
  m 
s
Pf 1 exp 
s
where the constant m (the Weibull modulus) determines the
slope of the distribution function and characterizes the
spread of the failure data with respect to s (fracture stress).
s is known as the scale parameter or characteristic strength,
that the stress level at which 63.21% of the specimens have
failed.
Weibull scale and shape parameters, standart errors and
95% condence levels of the data were obtained using R
Statistical Software (Foundation for Statistical Computing,
Vienna, Austria). As a further test for the statistical difference
of the scale and shape parameters, an independent samples t
test was carried out between the two data sets for each
experimental procedure. p-values of o0.05 were taken to
indicate the statistical signicance of the data.

3.

Results

3.1.
Uniaxial exural strength test and Weibull analysis
results
Mean and standard deviation of exural strength results are
given in Table 1. Statistical analysis revealed that there were
signicant differences among the exural strength results of
all tested groups (F 4.510, po0.05) (Fig. 1). Groups A, B2 and
B3 showed the highest mean exural strength values, statistically similar to groups B1, C and D; and signicantly
different from group E. There were no signicant differences

Fig. 1 Box-plot diagram of the distribution of exural


strength data according to tested groups.

82

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

in the mean exural strength values among B1, C, D and E


groups.
Total Weibull probability for fracture diagramme of tested
groups was shown in Fig. 2. According to the t test, Weibull
moduli of all tested groups were statistically signicant.
While, Weibull distribution presented the highest shape
value for group D (m7.7071.48), group C demonstrated the
lowest value (m 5.0670.98). Signicant difference was found
between Group A and Group B3 (p 0.01), Group A and Group

Fig. 2 Curves representing calculated values of total


Weibull probability for fracture in tested groups.

C (p 0.02), Group A and Group D (p 0.003) and; Group A and


Group E (p 0.016) (Table 1). Moreover, intergroup comparison
demonstrated that the shape values of Group B1 were
signicantly lower than Group B3 values (p 0.041).

3.2.

X-ray diffraction results

XRD patterns of tested groups are demonstrated in Fig. 3.


The tetragonal phase combined with a small amount of
monoclinic phase was observed for all tested specimens.
Additionally, detectable monoclinic peaks with a marked
preference of the m(111) orientations were appeared, indicating phase transformation (t-m), in the XRD pattern of A, B3,
D and E groups. However, monoclinic m(111) peak was barely
seen in groups B1 and B2. On the other hand, the sharp T(111)
peak, observed in specimens of B1, C, D and E which indicates
the enhancement of the crystallinity of the tetragonal phase.
A similar XRD pattern was obtained in groups D and E. Also,
this pattern was similar with groups A and B2. Moreover,
groups C and B3 exhibited similar patterns.
The relative amount of the monoclinic phase detected by
XRD on the tested specimens was reported in Table 2.
A greater amount of monoclinic phase was determined in
the specimens of groups D, E and B3 (25.43%, 20.89% and
19.71%, respectively). However, lower amount was observed
in groups A, B1 and B2 (13.35%, 10.02% and 15.19%,

Fig. 3 XRD patterns of zirconia specimens following different surface treatments, cyclic fatigue, thermal and chemical
degradation.

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

83

Table 2 Mean values of peak intensity ratios, monoclinic volume ratios, relative amounts of monoclinic phase and
transformed zone depth of the specimens according to the tested groups.
Tested groups

Xm (Peak
intensity ratio)

Vm (Monoclinic
volume ratio)

Relative amounts of
monoclinic phase (wt%)

TZD (Transformed
zone depth) (lm)

A
B1
B2
B3
C
D
E

0.08229
0.060842
0.094395
0.125001
0.103118
0.165563
0.133205

0.10519
0.078283
0.120223
0.157743
0.130986
0.206424
0.167685

13.35%
10.02%
15.19%
19.71%
16.50%
25.43%
20.89%

0.189
0.138
0.218
0.294
0.239
0.398
0.314

Fig. 4 Representative scanning electron microscopy of the surface of tested specimen in each group (Original magnication
4000  , bar 5 m). Micrographs with lower magnication are shown as inserts.

respectively). Additionally, transformed zone depth, ranging


between 0.138 to 0.398 mm, revealed higher values for groups
D and E. The Vm and TZD values were higher in group B3 than
B2 and B1 groups.

3.3.

SEM results

Fig. 4. shows the surface morphology of the tested specimens in


each group at  2000 and  4000 magnications. There were
marked differences among the tested groups. SEM images of the
surface of specimens in groups A, C, D and E were characterized
by a relatively smooth surface with regularly distributed

tetragonal zirconia crystal grains. The specimens in groups B1,


B2 and B3 exhibited a surface morphology consisting of uniformly damaged and edge-shaped micro-rough surface texture
with shallow pits and distributed micro-irregularities. Fig. 5.
demonstrates the representative SEM of the fractured surface of
a tested specimen after unaxial exural testing. The hackle
lines surrounded the failure origin (white arrows) on the tensile
surface of the specimen were easily seen on the micrographs at
 150 magnication. The principal mode of failure was intergranular fracture between the zirconia granules. Transgranular
fracture was also observed. In addition, small amounts of
porosity and aws were detected in the fracture areas.

84

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

Fig. 5 Representative scanning electron microscopy of the


fractured surface of a tested specimen after three-point
bending test (Original magnication 150  , bar 100 m).
Fracture load direction determined by the hackle line
direction (white arrows) may easily be seen.

4.

Discussion

In this study, a signicant difference in exural strength


results was found amongst the tested groups, leading to the
rejection of the rst null hypothesis. The results showed that
sandblasting with Al2O3 at 200, 400 or 600 kPa did not affect
the exural strength of zirconia, since there was no statistic
difference among the control group and the sandblasted
groups.While several previous studies did not report any
improvements on the mechanical properties of zirconia by
air-abrasion (Curtis et al., 2006; Zhang et al., 2006), the
investigators demonstrated the enhancement in the mechanical strength of these ceramics after air-abrasion (Guazzato
et al., 2004; Sato et al., 2008). Our nding is in accordance with
a recent study (Souza et al., 2013) which presented nonsignicant effect of the air-abrasion with 110 mm Al2O3 either
at 200 kPa or 350 kPa pressure on the biaxial exural strength
of zirconia specimens.
Additionally, in this three-point bending test setup, specimen size was selected bigger (19.3  14.5  1.3 mm3) to be
more exposed to different treatments explain partly lower
strength values in this study. It was reported that the
increase in specimen size increases the probability of nding
a larger defect present and therefore a reduction in strength.
The area or volume of a material tested modies the resultant strengths in this manner (Ashby and Jones, 1986).
In the current study, the specimens in group C were
subjected to cyclic loading in distilled water at 37 1C to
simulate the clinical conditions. Cyclic fatigue in water was
demonstrated to have a high impact on the lifespan of
zirconia materials yielding to signicantly lower results than
mechanical cycling which was performed in dry conditions
(Studart et al., 2007b). On the other hand, various results were
reported in the literature (Larsson et al., 2007; Ozcan et al.,
2013) which may be attributed to the variations of the cyclic
loading conditions and protocols used. In the present study,
cyclic loading was performed under 200 N 20,000 times

sinusoidal loading/unloading at a frequency of 10 Hz between


10% and 100% load. Since the aim of this study was to
evaluate the effect of cyclic loading on zirconia without
fracturing the specimens prior to exural tests, it was decided
to perform 20,000 cycles with a 200 N load. With regard to the
previous point, 10 Hz is relatively high compared to typical
mastication frequency varies from 1 to 2 Hz. This cycling
parameter was selected simply to keep the tests to a reasonable duration which was also conrmed by a study (Zhang
et al., 2006). In addition, this high frequency is not anticipated
to have an inuence on the results, since a previous study
(Zhang et al., 2004) has demonstrated that fatigue-related
strength degradation does not depend on the frequency of
loading. Our results showed that, the specimens in group C
exhibited similar exural strength values compared with
control.
Due to the phase transformation of zirconia ceramics of
which is thermally activated and accelerated by the presence
of water, steam autoclave treatments at increased temperatures have been shown to serve as a good method for
accelerated testing of LTD (Chevalier et al., 1999; Lee et al.,
2012). This method was specied in the ISO standard (ISO
13356, 2008) as the simulation between 5 years and 20 years
of exposure at body temperature (Chevalier et al., 1999). The
specimens in group D, submitted to steam in an autoclave at
134 1C and 0.2 MPa for 5 h, had similar strength results
compared with control. Amaral et al. (2013) found that, the
zirconia specimens which were subjected to steam autoclave
treatments at 12771 1C at 150 kPa additional air pressure for
12 h, had signicant increases in strength values after LTD.
Controversially, Flinn et al. (2012) evaluated the accelerated
aging characteristics of different zirconia materials at 50, 100,
150, and 200 h and reported the signicant decrease in
exural strengths of tested specimens after 200 h at 134 1C
and 0.2 MPa.
Furthermore, chemical degradation testing was performed
in group E specimens by immersing in 4% acetic acid at
8075 1C for 168 h in the present study. Chai et al. (2007)
reported that, the effect of 4% acetic acid used for 1 week at
80 1C was like the immersion in articial saliva at 22 1C for 22
years. It has been shown in a previous study (Ardlin, 2002)
that 99% ZrO2 specimens were resistant even to 168 h of 4%
acetic acid reux without compromising their subsequent
exural strength. In contrast to this study, our study demonstrated that exural strength of group E was signicantly
lower than control. These conicting results may be the
consequence of using different zirconia materials. Kosmac
et al. (2000) reported that, the diffusion-controlled transformation of zirconia strongly depends on the grain size.
A larger grain size, which resulted in higher damage tolerance of the mechanically treated Y-TZP ceramics, may thus
turn out to be a disadvantage during prolonged aging below
100 1C in an acidic environment. The authors speculated that
along with propagating transformation during aging, exural
strength would raise at the beginning. However, the formation of microcracks and the buildup of tensile residual
stresses could lead to a decrease in strength with progressing
t-m transformation.
The failure behavior of tested specimens can best be
discussed in terms of the Weibull parameters given in

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

Table 1. Weibull analysis was used in this study to provide


information related to the overall performance of tested
groups, rather than relying on the mean exural strength and
standard deviation. Weibull modulus (m) is used to describe
the variation of the strength distribution as a result of aws
and microcracks which may develop within the microstructure (Pittayachawan et al., 2007). The lower the value of the
Weibull modulus is indicative of the variation in surface state
which contains more aws and defects in the material and
decreased reliability. Conversely, the higher values of Weibull
modulus characterize the low inherent aw density and high
structural reliability (Pittayachawan et al., 2007). Another
factor that should be considered is the number of test specimens in using Weibull analysis (Quinn and Quinn., 2010).
As a general rule-of-thumb, approximately 30 test specimens
provide adequate Weibull strength distribution parameters.
However, numerous studies have addressed the level of
structural reliability of different dental materials with lower
numbers of specimens by Weibull analysis (Fennis et al.,
2005; Bona et al., 2003). It was reported that, the optimal
number of test specimens depends on many variables,
including cost of material and testing procedures, the values
of the distribution parameters and the desired precision for
an intended application (Quinn and Quinn., 2010). Therefore,
interpretation of the results of the current study should be
made carefully, since 15 specimens in each group were
subjected to Weibull analysis. However, it is believed the
results have some useful validity, as the results of the current
study follow the Weibull curve quite well.
Statistical analysis demonstrated that the Weibull modulus of all tested groups was signicantly different, ranging
from 5.06 to 7.7 which were in the expected range for ceramic
materials and similar to those values achieved in the previous studies (Kosmac et al., 1999; Bona et al., 2003).
The results of the current study revealed that sandblasting
with Al2O3 at 600 kPa pressure resulted in greater m values
when it was compared with control although mean biaxial
exural strength values were similar. Additionally, LTD conditions seem to increase the m values of control group.
However, cyclic loading and acetic acid treatment signicantly decreased the structural reliability of the material.
Therefore, the second hypothesis that, surface conditioning
procedures and LTD conditions tested in this study would not
change the structural reliability of the material could be
rejected. Furthermore, intergroup comparison demonstrated
that, the shape values of Group B1 were signicantly lower
than Group B3 values. Numerous studies reported that
sandblasting lowered the Weibull modulus of Y-TZP ceramics
(Kosmac et al., 1999; Guazzato et al., 2005b). Likely, alumina
abrasion has been reported to introduce deep surface aws
which can act as stress concentrators and become strengthlimiting factors if aw length extends beyond the surface
compressive layer (Luthardt et al., 2004; Wang et al., 2008). On
the other hand, according to the statistical analysis of m
values for Group D were signicantly the highest which was in
accordance with a previous study by Amaral et al. (2013). The
researchers (Amaral et al., 2013) implied that the amount of
m-phase on the zirconia surface is directly related to the
strength owing to the presence of compressive stresses on
the tensile surface, which resulted from the t-m phase

85

transformation in the crystalline structure. Similarly, the


results of the present study revealed that monoclinic zirconia
content was increased to 25.43% in group D, whereas groups
B1, B2 B3 consisted of monoclinic phase in the range of E10
20%. Thus, the third hypothesis that sandblasting with various
propulsion pressures, cyclic loading or chemical and thermal
LTD degradation would not lead to an increase in the monoclinic phase content in the Y-TPZ ceramic could be rejected. It
was observed that Xm values were increased with the increase
of air-abrasion pressure (B1: 10.02%, B2: 15.19% and B3: 19.71%).
Lower amounts of monoclinic phase of sandblasted groups
could be attributed to the formation of a protective layer of
residual compressive stresses on the zirconia surface as a
consequence of phase transformation (t-m) in the airabraded area (Kosmac et al., 1999; Guazzato et al., 2004;
Ozcan et al., 2013). The reported data of the present study on
phase composition of sandblasted groups are in agreement
with previous studies (Kosmac et al., 1999; Guazzato et al.,
2005a).
Moreover, 20.89% monoclinic content was determined in
group E. Upon immersion in acetic acid under the conditions
adopted for chemical degradation testing in the current
study, a signicant amount of tetragonal zirconia has transformed to monoclinic. The differences in the Xm values
between tested groups also conrmed that statistically signicant strength degradation has occurred compared to the
control. Therefore, based on the results of our study, the
authors speculated that, Y-TZP ceramics are susceptible to
low temperature degradation in an acidic environment.
A previous study by Kosmac et al. (2008) evaluated the
chemical solubility of different dental ceramics by extracting
the specimens with the 4% acetic acid solution by reuxing
for 16 h. In contrast to our results, although the researchers
demonstrated signicant amount of tetragonal zirconia,
strength degradation was not observed. These conicted
results in these two studies might be related with the
different extraction times.
Sandblasting was considered to be a gentle process, during
which considerably less material is removed from the surface
(Kosmac et al., 1999). However, in the current study, the
thickness of the transformed surface layer was found larger
in groups B2 and B3 than in group A, although exural
strength values were similar among these groups. The
authors speculated that the length of surface aws, which
were introduced by sandblasting, seemed to exceed largely
the thickness of the compressive surface layer, since the
strength of the material was not affected. Another important
nding of our study was that, specimens of group D, subjected to LTD, exhibited a degradation layer of about 0.4 mm
which was calculated almost 2 times higher than the control
specimens. As mentioned before, this group has the highest
Weibull modulus. This nding is in agreement with a recent
study (Amaral et al., 2013) of which TZD of zirconia specimens were increased from 0.07 mm to 1.55 mm after LTD
procedure. The authors Amaral et al. (2013) reported that
the formation of a compression layer might be responsible for
the increase in zirconia reliability, since small cracks, aws,
and defects that would have led to fracture. Thus, this
fracture probably be healed by the volume increase in the
grains (approximately 4%).

86

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

The limitation of this study was that the in vitro design


did not completely mimic the intraoral environment. Therefore, further study designs should be created to investigate
the fatigue behavior of various dental Y-TZP ceramics under
clinical conditions.

5.

Conclusions

Within the limitations of the current study, the following


conclusions can be drawned:
(1) Flexural strength of a zirconia ceramic was signicantly
decreased by chemical degradation.
(2) Sandblasting with Al2O3 at at 600 kPa pressure, cyclic
loading and LTD conditions including chemical degradation tested in this study signicantly changed the reliability of the material strength.
(3) The specimens subjected to thermal degradation exhibited the highest t-m phase transformation whereas this
group has the highest Weibull modulus.

references

Amaral, M., Valandro, L.F., Bottino, M.A., Souza, R.O., 2013. Lowtemperature degradation of a Y-TZP ceramic after surface
treatments. J. Biomed. Mater. Res. B Appl. Biomater. 101,
13871392.
Ardlin, B.I., 2002. Transformation-toughened zirconia for dental
inlays, crowns and bridges: chemical stability and effect of
low-temperature aging on flexural strength and surface
structure. Dent. Mater. 18, 590595.
Ashby, M.F., Jones, D.R.H., 1986. Eng. Mater., vol. 2. Pergamon
Press, Oxford, ISBN: 0-08-032532-7.
Bona, A.D., Anusavice, K.J., DeHoff, P.H., 2003. Weibull analysis
and flexural strength of hot-pressed core and veneered
ceramic structure. Dent. Mater. 19, 662669.
Chai, J., Chu, F.C., Chow, T.W., Liang, B.M., 2007. Chemical
solubility and flexural strength of zirconia-based ceramics.
Int. J. Prosthodont. 20, 587595.
Chevalier, J., Cales, B., Drouin, J.M., 1999. Low-temperature aging
of Y-TZP ceramics. J. Am. Ceram. Soc. 82, 21502154.
Curtis, A.R., Wright, A.J., Fleming, G.J.P., 2006. The influence of
simulated masticatory. J. Dent. 34, 317325.
Denry, I., Kelly, J.R., 2008. State of the art of zirconia for dental
applications. Dent. Mater. 24, 299307.
Egilmez, F., Ergun, G., Cekic-Nagas, I., Vallittu, P.K., Ozcan, M.,
Lassila, L.V., 2013. Effect of surface modification on the bond
strength between zirconia and resin cement. J. Prosthodont.
22, 529536.
Esquivel-Upshaw, J.F., Chai, J., Sansano, S., Shonberg, D., 2001.
Resistance to staining, flexural strength, and chemical
solubility of core porcelains for all-ceramic crowns. Int. J.
Prosthodont. 14, 284288.
Fennis, W.M., Tezvergil, A., Kuijs, R.H., Lassila, L.V., Kreulen, C.M.,
Creugers, N.H., Vallittu, P.K., 2005. In vitro fracture resistance
of fiber reinforced cusp-replacing composite restorations.
Dent. Mater. 21, 565572.
Flinn, B.D., deGroot, D.A., Mancl, L.A., Raigrodski, A.J., 2012.
Accelerated aging characteristics of three yttria-stabilized
tetragonal zirconia polycrystalline dental materials.
J. Prosthet. Dent. 108, 223230.

Garvie, R.C., Hannink, R.H., Pascoe, R.T., 1975. Ceramic steel.


Nature 258, 703704.
Garvie, R.C., Nicholson, P.S., 1972. Phase analysis in zirconia
systems. J. Am. Ceram. Soc. 55, 303305.
Green, D.J., 1983. A technique for introducing surface
compression into zirconia ceramics. J. Am. Ceram. Soc. 66,
C178C179.
Guazzato, M., Albakry, M., Ringer, S.P., Swain, M.V., 2004. Strength,
fracture toughness and microstructure of a selection of allceramic materials. PartII. Zirconia-based dental ceramic.
Dent. Mater. 20, 449456.
Guazzato, M., Quach, L., Albakry, M., Swain, M.V., 2005a. Influence
of surface and heat treatments on the flexural strength of
Y-TZP dental ceramic. J. Dent. 33, 918.
Guazzato, M., Albakry, M., Quach, L., Swain, M.V., 2005b. Influence
of surface and heat treatments on the flexural strength of a
glass infiltrated alumina/zirconia-reinforced dental ceramic.
Dent. Mater. 21, 454463.
Guess, P.C., Att, W., Strub, J.R., 2012. Zirconia in fixed implant
prosthodontics. Clin. Implant Dent. Relat. Res. 14, 633645.
Guo, X., 2004. Property degradation of tetragonal zirconia induced
by low-temperature defect reaction with water molecules.
Chem. Mater. 16, 39883994.
International Organization for Standardization. ISO 6872, 1995.
(E): Dental Ceramics, 2nd ed.
International Organization for Standardization. ISO 13356, 2008.
Implants for Surgery-Ceramic Materials Based on Yttriastabilized Tetragonal Zirconia (Y-TZP), Geneva.
Kim, H.T., Han, J.S., Yang, J.H., Lee, J.B., Kim, S.H., 2009. The effect
of low temperature aging on the mechanical property and phase
stability of YTZP ceramics. J. Adv. Prosthodont. 1, 113117.
Kosmac, T., Oblak, C., Jevnikar, P., Funduk, N., Marion, L., 1999.
The effect of surface grinding and sandblasting on flexural
strength and reliability of Y-TZP zirconia ceramic. Dent. Mater.
15, 426433.
Kosmac, T., Oblak, C., Jevnikar, P., Funduk, N., Marion, L., 2000.
Strength and reliability of surface treated Y-TZP dental
ceramics. J. Biomed. Mater. Res. 53, 304313.
Kosmac, T., Oblak, C., Marion, L., 2008. The effects of dental
grinding and sandblasting on ageing and fatigue behavior of
dental zirconia (Y-TZP) ceramics. J. Eur. Ceram. Soc. 28, 10851090.
Larsson, C., Holm, L., Lovgren, N., Kokubo, Y., Von, S., Vult von
Steyern, P., 2007. Fracture strength of four-unit Y-TZP FPD
cores designed with varying connector diameter. An in-vitro
study. J. Oral. Rehabil. 34, 702709.
Lee, T.H., Lee, S.H., Her, S.B., Chang, W.G., Lim, B.S., 2012. Effects
of surface treatments on the susceptibilities of low
temperature degradation by autoclaving in zirconia. J. Biomed.
Mater. Res. Part B 100, 13341343.
Lughi, V., Sergo, V., 2010. Low temperature degradation -aging- of
zirconia: a critical review of the relevant aspects in dentistry.
Dent. Mater. 26, 807820.
Luthardt, R.G., Holzhuter, M.S., Rudolph, H., Herold, V., Walter, M.H.,
2004. CAD/CAM-machining effects on Y-TZP zirconia. Dent.
Mater. 20, 655662.
Ozcan, M., Melo, R.M., Souza, R.O., Machado, J.P., Felipe Valandro, L.,
Botttino, M.A., 2013. Effect of air-particle abrasion protocols on
the biaxial flexural strength, surface characteristics and phase
transformation of zirconia after cyclic loading. J. Mech. Behav.
Biomed. Mater. 20, 1928.
Quinn, J.B., Quinn, G., 2010. A practical and systematic review of
Weibull statistics for reporting strengths of dental materials.
Dent. Mater. 26, 135147.
Piconi, C., Maccauro, G., 1999. Zirconia as a ceramic biomaterial.
Biomaterials 20, 125.
Pittayachawan, P., McDonald, A., Petrie, A., Knowles, J.C., 2007.
The biaxial flexural strength and fatigue property of Lava
Y-TZP dental ceramic. Dent. Mater. 23, 10181029.

journal of the mechanical behavior of biomedical materials 37 (2014) 78 87

Porter, D.L., Heuer, A.H., 1977. Mechanisms of toughening


partially stabilized zirconia (PSZ). J. Am. Ceram. Soc. 60,
183184.
Sato, H., Yamada, K., Pezzotti, G., Nawa, M., Ban, S., 2008.
Mechanical properties of dental zirconia ceramics changed with
sandblasting and heat treatment. Dent. Mater. J. 27, 408414.
Souza, R.O., Valandro, L.F., Melo, R.M., Machado, J.P., Bottino, M.A.,
Ozcan, M., 2013. Air-particle abrasion on zirconia ceramic
using different protocols: effects on biaxial flexural strength
after cyclic loading, phase transformation and surface
topography. J. Mech. Behav. Biomed. Mater. 26, 155163.
Studart, A.R., Filser, F., Kocher, P., Gauckler, L.J., 2007a. In vitro
lifetime of dental ceramics under cyclic loading in water.
Biomaterials 28, 26952705.
Studart, A.R., Filser, F., Kocher, P., Luthy, H., Gauckler, L.J., 2007b.
Cyclic fatigue in water of veneer-framework composites for
all-ceramic dental bridges. Dent. Mater. 23, 177185.

87

Toraya, H., Yoshimura, M., Somiya, S., 1984. Calibration curve for
quantitative analysis of the monoclinic tetragonal ZrO2
system by X ray diffraction. J. Am. Ceram. Soc. 67, 119121.
Wang, H., Aboushelib, M.N., Feilzer, A.J., 2008. Strength
influencing variables on CAD/CAM zirconia frameworks. Dent.
Mater. 24, 633638.
Weibull, W., 1951. A statistical distribution function of wide
applicability. J. Appl. Mech. 18, 293297.
Yoshimura, M., Noma, T., Kawabata, K., Somiya, S., 1987. Role of
H2O on the degradation process of Y-TZP. J. Mater. Sci. Lett. 6,
465467.
Zhang, Y., Lawn, B.R., Malament, K.A., Thompson, V.P., Rekow, D.,
2006. Damage accumulation and fatigue life of particleabraded ceramics. Int. J. Prosthodont. 19, 442448.
Zhang, Y., Pajares, A., Lawn, B.R., 2004. Fatigue and damage
tolerance of Y-TZP ceramics in layered biomechanical
systems. J. Biomed. Mater. Res. B. Appl. Biomater. 15, 166171.

Potrebbero piacerti anche