You are on page 1of 6

Ceramics International 49 (2023) 21716–21721

Contents lists available at ScienceDirect

Ceramics International
journal homepage: www.elsevier.com/locate/ceramint

Analysis of zirconia mechanical properties after application of a protocol to


simulate clinical aging
Beshr Hajhamid a, Laurent Bozec a, Howard Tenenbaum a, Eszter Somogyi-Ganss a,
Grace M. De Souza b, *
a
Faculty of Dentistry, University of Toronto, 124 Edward St, Toronto, ON, M5G1G6, Canada
b
Department of Comprehensive Dentistry, School of Dentistry, University of Louisville, 501 S. Preston St., Louisville, KY, 40202, United States

A R T I C L E I N F O A B S T R A C T

Handling EditorDr P. Vincenzini In order to test the effects of hydrothermal and clinically related aging on zirconia, monolithic disc-shaped
samples were milled, sintered and polished from two high translucency zirconia , 3 mol % (HT) and 5 mol %
Keywords: (XT). Samples were divided into three groups: non-aged: control (CT); hydrothermal aging (HA - autoclave aging
Hydrothermal aging for 12.5 h at 134 ◦ C, 2 bar); in vitro clinically-related aging (CRA - chewing simulation for 1.2 million cycles
In vitro aging
followed by 50,000 thermocycles and acidic exposure in HCl, pH 1.2, for 15 h). Mechanical properties (flexural
Fatigue strength
strength, fatigue, hardness and elastic modulus) were analyzed and compared using the analysis of variance (at a
Clinically-related aging
level of 5% significance). In vitro clinically-related aging significantly decreased fatigue strength of XT zirconia
with no effects on HT zirconia. Surface hardness and elastic modulus were not affected (p = 0.591 and 0.392
respectively). Hydrothermal aging increased fatigue strength for both materials and decreased the surface
hardness and modulus of HT zirconia (p ≤ 0.001). Hydrothermal aging and in vitro clinically-related aging have
different effects on the mechanical properties of zirconia , when used to simulate five years of clinical service.

1. Introduction tetragonal zirconia has low translucency due to the non-uniform


three-dimensional shape of tetragonal crystals, and the distribution of
Zirconia restorations are used widely in restorative dentistry due to Al2O3 along the tetragonal crystal boundaries [6,7] which has a negative
their excellent mechanical and aesthetic properties as well as their impact on its aesthetic properties. Thus, a new generation of high
known biocompatibility [1]. Dental zirconia, made of zirconium dioxide translucency tetragonal zirconia was developed with lower Al2O3 con­
(ZrO2), is composed of three temperature-dependent crystals: mono­ tent, known as ‘2nd generation zirconia’ [8]. Another approach to
clinic (m), tetragonal (t) and cubic (c) [2]. Yttria, or Yttrium oxide improve the translucency of zirconia-based materials is to increase the
(Y2O3), is used commonly to stabilize cubic and tetragonal zirconia amount of isotropic cubic crystals [9]. Increasing yttria content from 3
crystals at room temperature. Three mol-% of Y2O3 is used to stabilize to 4 or 5 mol % allows more than 50% of cubic crystals to be stabilized at
the tetragonal crystals at room temperature to produce conventional room temperature and results in improved optical properties [8]. These
tetragonal zirconia, and alumina (Al2O3) is added (0.25 wt %) as a materials are often referred to as ‘3rd generation zirconia’ but have
sintering aid [3]. The tetragonal crystals have the highest density among reduced mechanical properties as compared to conventional and 2nd
zirconia crystals, but are metastable under certain stimuli and can revert generation zirconia due to the higher concentration of cubic crystals and
back to a monoclinic crystal form (t-m transformation) [4]. Trans­ the lack of transformation toughening [10].
formation toughening is used to provide a positive impact on tetragonal A drawback of t-m transformation is known as low-temperature
crystals’ metastability since the compressive stress associated with the degradation (LTD) which occurs when the metastable tetragonal crys­
transformation can seal a crack from propagating and increase zirconia’s tals are exposed to humidity and transform back to m spontaneously
toughness and flexural strength, something that is particularly impor­ over time [11,12]. This process leads to increased surface roughness
tant when under occlusal function [4,5]. However, conventional with the formation of microcracks through which water molecules can

* Corresponding author.
E-mail addresses: Beshr.hajhamid@utoronto.ca (B. Hajhamid), l.bozec@dentistry.utoronto.ca (L. Bozec), Howard.Tenenbaum@sinaihealth.ca (H. Tenenbaum), e.
s.ganss@utoronto.ca (E. Somogyi-Ganss), grace.desouza@louisville.edu (G.M. De Souza).

https://doi.org/10.1016/j.ceramint.2023.03.311
Received 17 November 2022; Received in revised form 27 March 2023; Accepted 30 March 2023
Available online 1 April 2023
0272-8842/© 2023 Elsevier Ltd and Techna Group S.r.l. All rights reserved.
B. Hajhamid et al. Ceramics International 49 (2023) 21716–21721

penetrate thus leading to t-m transformation in the sub-surface of zir­ GMBH) to simulate an opposing tooth. Samples were exposed to 1.2
conia [13]. million cycles in the chewing simulator with 0.5 mm vertical and hori­
Clearly, more understanding of issues that regulate and/or partici­ zontal paths (0–50 N, 60 mm/min, artificial saliva, 37 ◦ C) [20]. After
pate in the changes noted above needs to be investigated further. In this, samples were exposed to thermocycling between temperatures of
relation to this, hydrothermal aging in an autoclave has been introduced 5 ◦ C and 55 ◦ C for 50,000 cycles (30 s in each bath, 15 s transfer time; SD
to simulate the accelerated aging of zirconia to estimate the stability of Mechatronik Thermocycler GMBH). Then, the acidic challenge was
the zirconia crystals [14] that might correlate to longer term aging/use applied by immersion of samples in simulated gastric fluid (pepsin-free
in the mouth. A previous study suggested that 5 h of autoclave accel­ 0.2% w/v sodium chloride in 0.7% v/v hydrochloric acid, pH = 1.2) at
erated aging (134 ◦ C, 2 bar in steam) corresponds to two years of 25 ◦ C for 15 h (simulated gastric fluid, Ricca chemical company, Lot
intraoral aging [15]. However, intraoral aging of a material encom­ #4004D83). Samples were rinsed with distilled water for 30 s and stored
passes numerous other factors that may account for additional chal­ in fresh distilled water at room temperature until further tests were
lenges to the stability of zirconia, including but not necessarily limited to applied.
humidity and temperature. For example, mastication produces cyclic
loads on teeth and restorative materials [16]. In addition, eating and 2.3. Mechanical properties
drinking episodes result in pH and temperature variations [17] that
could play an essential role in the stability of zirconia crystals, affecting 2.3.1. Flexural strength
the longevity of the restorations. Biaxial flexural strength was measured in 10 samples from each
Many previous studies evaluated the effects of hydrothermal aging group using the piston-on-three-ball configuration (6872–2015) in a
on zirconia-based materials. Still, there is no evidence of the effect of universal testing machine (Instron Model 8501). Disc-shaped samples
more clinically-related factors such as mastication, temperature and pH were placed on 3 mm supporting balls, positioned 120◦ from each other
variations on zirconia-based materials. Therefore, the objective of this in a 10 mm diameter circle (ISO Standard 6872:2015). The load was
study was to compare the effect of in vitro clinically-related aging applied perpendicular to the center of the sample with a speed of 1.0
including hydrothermal aging on the mechanical properties of zirconia- mm/min. The load-to-failure was recorded for each sample and con­
based materials. We hypothesize that the different methods used for verted into MPa.
simulation of aging decrease the mechanical properties of zirconia-
based materials, and that mechanical properties will be affected by 2.3.2. Fatigue strength
material composition. Fatigue strength based on the staircase approach was analyzed using
the same fixture applied for the biaxial flexural strength test [21].
2. Materials and methods Samples were immersed in a container filled with distilled water at room
temperature (Instron Model 8501) and fatigue loading was applied for
2.1. Sample preparation 100,000 cycles, at a frequency of 6 Hz. The parameters for the test were
set based on piston-on-three-ball flexural strength mean values: initial
Two high translucency forms of zirconia: 2nd generation (3 mol % - load – 70% of the mean flexural strength; step size – 5% of the mean
Cercon HT Dentsply Sirona Prosthetics- HT) and 3rd generation (5 mol flexural strength. If the sample failed before completing the 100,000
% - Cercon XT- Dentsply Sirona Prosthetics- XT) were milled using cycles of fatigue, the load applied to the next specimen was decreased by
computer-aided design/computer-aided manufacturing (CAD/CAM) one step size (5% of the mean flexural strength). If the sample survived
technique to obtain 90 disc-shaped samples from each material. Samples 100,000 cycles, the load to the next specimen was increased by one step
were sintered according to the manufacturer’s recommendations and size. This procedure was repeated until 15 samples were tested after the
the final dimensions of the discs created had a diameter of 16 mm and a first inversion of events.
thickness of 1.2 ± 0.2. All samples had one side polished using silicon
carbide papers (400, 600, 800 and 1200 grit, Buehler, Illinois, US) fol­ 2.3.3. Surface hardness and modulus
lowed by 0.3–0.6 μm diamond paste suspension using a rotary grinder Following ultrasonic cleansing in acetone for a period of 10 min, the
and polisher (Metaserv 250 Grinder/Polisher, Buehler) at a speed of 300 surface hardness and modulus of samples were measured using a
rpm. After being cleaned ultrasonically in acetone for 5 min and air- nanoindenter (NHT3 – Anton Paar GMBH). A 50 nm radius Berkovich
dried, samples were annealed in a laboratory chamber furnace diamond tip indenter was used to perform seven progressive multicycles
(1200 ◦ C/2 h, CWF1300, Carbolite) to remove compressive stresses of loading and partial unloading at ten different locations on each
occurring during sample preparation [18]. specimen (10 Hz, speed of penetration depth 2000 nm/min). Hardness
and elastic modulus values were determined using the Oliver-Pharr
2.2. Aging method [22].

Samples from each material were distributed randomly into three 2.4. Statistical analysis
different groups (n = 30 per group) as follows: no treatment (Control,
CO), hydrothermal aging (HA), and in vitro clinically-related aging After checking for normal distribution of datapoints between the
(CRA). samples for the values of flexural strength, fatigue resistance, hardness
CO–No aging was applied. Samples were stored in distilled water at and modulus, two-way analysis of variance (ANOVA) and Tukey post-
room temperature until tests were applied. hoc test (SPSS Inc) were used to evaluate the effects of aging on me­
HA – zirconia samples were aged for 12.5 h in the autoclave (Ritter chanical and surface properties. The independent factors included zir­
M9 Midmark Sterilizer) at 134 ◦ C, with 2 bars of pressure [19]. Then, conia composition and aging. The Dixon-Mood method was used to
samples were stored in distilled water at room temperature until further determine the mean fatigue limits and standard deviation. A Kaplan-
testing. Meier survival analysis was performed to determine the survival rate
CRA – this model was developed to simulate mechanical, thermal of specimens after fatigue. An overall significance level of 5% was set for
and acidic challenges occurring in the oral cavity. Zirconia samples were all tests.
embedded with polymethylmethacrylate (PMMA, Ivoclar Vivadent) in a
metallic holder and stabilized to the base of a chewing simulator
(Mechatronik CS 4.4, SD Mechatronik GMBH). The antagonist surface
was a 6 mm diameter spherical steatite indenter (SD Mechatronik

21717
B. Hajhamid et al. Ceramics International 49 (2023) 21716–21721

3. Results Table 2
Surface hardness and elastic modulus of zirconia-based materials before and
3.1. Flexural strength after aging. Different letters in each column represent 5% significance among
the groups.
A significant effect of aging (p ≤ 0.001), material (p ≤ 0.001) and Material Aging Surface hardness GPa (±SD) Surface modulus GPa (±SD)
interaction between material and aging (p ≤ 0.001) was observed for HT CO 20.1 (2.0)a
273.6 (18.9)a
flexural strength, as shown in Table 1. The lowest flexural strength was HA 13.0 (0.9)b 208.6 (34.6)b
observed for XT zirconia after CRA (315.4 ± 79.6 MPa) and the highest CRA 20.0 (4.3)a 261.8 (18.4)a
flexural strength among groups was recorded for HT zirconia after HA XT CO 18.8 (2.6)a 255.7 (18.0)a
(865.3 ± 179.3 MPa). HA 19.2 (2.0)a 259.7 (40.7)a
CRA 18.7 (2.8)a 251.3 (34.33)a

3.2. Fatigue strength There was a significant effect of aging (p ≤ 0.001) and the interaction between
material and aging on hardness (p ≤ 0.001) and modulus (p ≤ 0.001). In
Aging (p ≤ 0.001), material (p ≤ 0.001) and interaction (p = 0.003) contrast, the material composition was not significant on hardness (p = 0.094)
had a significant effect on fatigue resistance (Table 1). and modulus (p = 0.327). The lowest surface hardness (13 ± 0.9 GPa) and
The staircase results showed that there was a decrease in the strength modulus (208.6 ± 34.6 GPa) were observed for HT zirconia after hydrothermal
aging, as reported in Table 2.
of zirconia, ranging between 25.6% and 38.6% (Table 1). The number of
survival-failure and the range of cycles for failure are reported in
Table 1, and Fig. 1 shows the staircase profile for each zirconia after which taken together could represent more closely the environment to
aging. The fatigue strength of XT zirconia significantly decreased after which zirconia restorations would be exposed clinically. To reiterate,
CRA (211.6 ± 13.6 MPa) compared to CO group (310.5 ± 4.2 MPa). The this in vitro clinically-related model replicated some of the oral chal­
fatigue strength of HT zirconia was unaffected by CRA. Hydrothermal lenges, various intraoral chewing forces, temperature changes and
aging increased fatigue strength for both types of zirconia compared to acidic challenge faced by dental restorations in the mouth. The extent of
the control group, as observed in Table 1. each of those regimens was determined to simulate 5 years of clinical
There were no significant differences in the fatigue survival rates aging based on information published previously [15,27–29].
among the groups (Kaplan-Meier and the Long Rank test (Mantel-Cox) p The study hypothesis, that the different methods used for aging
= 0.924, Fig. 2). decreased the mechanical properties of zirconia-based materials, was
accepted. The results showed that hydrothermal aging for 12.5 h had no
significant effect on the flexural strength of zirconia, as previously re­
3.3. Surface hardness and modulus
ported in a study that investigated hydrothermal aging for up to 20 h
[12].
However, hydrothermal aging significantly increased the strength of
zirconia after 100,000 fatigue cycles, as based on the use of the staircase
4. Discussion
method. This effect of hydrothermal aging on HT zirconia might be
related to the volumetric expansion that occurs during fatigue cycling
This study investigated the effect of in vitro aging on the mechanical
associated with the transformation toughening mechanism of tetragonal
properties of two different high translucency forms of zirconia: 2nd
crystals to monoclinic at the superficial layer, resulting in the concen­
generation zirconia (3 mol % - Cercon HT); and 3rd generation (5 mol %
tration of the compressive stresses that prevent the initial propagation of
- Cercon XT). These in vitro methods are used often to estimate the
cracks. In support of these study results, a previous study by this team
degradation of dental zirconia is hydrothermal aging [14] in ways that
showed a transformation of 40% from tetragonal to monoclinic for HT
might yield information relating to their predicted functionality in the
zirconia after autoclave aging for 12.5 h [30]. The same study showed
oral cavity as restorations. At the same time, one study estimated that 5
0% of monoclinic crystals on HT and XT zirconia after sintering, pol­
h of autoclave aging represents 15–20 years of clinical service of
ishing, and annealing when the same protocol reported in this study was
biomedical zirconia [23], another study estimated that 5 h of autoclave
used [30].
aging simulates two years of clinical service for dental zirconia [15]. It
An increase in fatigue strength of 2nd generation zirconia as a
seems that in general, the correlation between hydrothermal aging and
consequence of hydrothermal aging has been reported previously [31].
clinical aging of zirconia has not been clarified. In relation to this, other
In in vitro clinically-related aging, the results showed a significant
studies have shown the lack of correlation between hydrothermal aging
decrease in flexural strength and fatigue resistance of XT zirconia only.
and chewing simulation [24–26], which is just one of the factors ac­
Both zirconia formulations used in this study contained different levels
counting for the clinical degradation of zirconia. Therefore, this study
of metastable tetragonal crystals. Based on the manufacturer’s descrip­
applied, what we believe to be a more in vitro clinically-related model of
tion,5 and 9% of yttria are used to stabilize tetragonal and cubic crystals
stressors to which zirconia restorations might be exposed in order to
in HT, and XT zirconia, respectively. HT zirconia is composed solely of
simulate challenges occurring in the mouth with the use of the methods.
tetragonal crystals after sintering (i.e., 100%), while XT contains 50%
These included the use of mechanical, chemical and thermal stresses,

Table 1
Flexural strength and fatigue resistance results for zirconia-based materials after hydrothermal aging (HA) and clinically-related aging (CRA). Different letters in each
column represent 5% significance among groups.
Material Aging Flexural strength MPa (±SD) Fatigue resistance MPa (±SD) Strength decrease (%) Numbers of Cycles to failure

Failure Survival

HT CO 778.8 (94.4)a 548.7 (113.9)b 29.6% 6 9 314 - 61,496


HA 865.3 (179.3)a 661.8 (65.3)a 25.6% 7 8 149 - 50,924
CRA 721.9 (160.6)a 512.8 (80.9)b 29% 7 8 1422–50,106

XT CO 505.4 (85.2)b 310.5 (4.2)d 38.6% 8 7 827 - 26,132


HA 484.6 (67.4)b 347 (64.1)c 28.4% 6 9 3665–71,065
CRA 315.4 (79.6)c 211.6 (13.6)e 32.9% 6 9 250 - 15,089

21718
B. Hajhamid et al. Ceramics International 49 (2023) 21716–21721

Fig. 1. Staircase profile for all tested groups. The orange line represents the mean fatigue limit. The black shaded triangle indicates failure, whereas the white shaded
triangle indicates the survival of the sample. The arrow represents the first sample prior to the inversion of events. (For interpretation of the references to colour in
this figure legend, the reader is referred to the Web version of this article.)

protocols on the strength of the HT zirconia. A previous study showed


only 5% of tetragonal to monoclinic transformation for HT zirconia after
aging using the same clinically-related aging model used in this study
[30]. The results of the current study are in concert with findings re­
ported previously where the flexural strength of 2nd generation zirconia
(IPS e.-max ZirCAD) could not be demonstrated even after being
exposed to one million cycles of chewing simulation [26]. This was also
demonstrated in another study where there were no changes in the
flexural strength of 2nd generation zirconia following 5000 cycles of
thermocycling [33]. It might also be postulated that the results reported
here regarding a lower rate of transformation toughening in XT are not
related solely to the content/concentration of tetragonal crystals as
compared to HT zirconia but also because cubic crystals are larger in size
than tetragonal crystals which could decrease the mechanical properties
of the material during aging. A previous study reported an average grain
size of 0.51 μm for HT zirconia and 1.49 μm for XT zirconia [30]. Smaller
crystal sizes are more resistant to mechanical stresses and thus to crack
propagation, which makes HT zirconia more resistant to fracture than a
Fig. 2. Staircase survival analysis plot obtained by Kaplan-Meier and the Long material with larger grain size, like XT zirconia [34]. Furthermore,
Rank test for number of cycles before failure (MPa). smaller crystals size limits the dislocation of crystalline boundaries,
resulting in increased mechanical properties of zirconia [35]. This might
tetragonal and 50% of cubic crystals due to the higher yttria content. help explain the observed reduction in strength of XT zirconia after in
Metastable tetragonal crystals are dense and provide better mechanical vitro clinically-related aging. Another explanation might be related to
properties compared to stabilized cubic crystals [32]. In fact, the the combination effect of chewing simulation with 50 N of loading force
compositional differences and crystalline structure of both in artificial saliva with the temperature and acidity changes during the
zirconia-based materials indicate that transformation toughening is aging therefore, cracks may propagate into the bulk of XT zirconia under
possibly only occurring in HT zirconia due to the higher presence of the stress of the chewing simulator, with a low rate of transformation
tetragonal crystals in the bulk of the material, and no significant trans­ toughening, and the presence of temperature and acidic changes.
formation toughening occurs after CRA aging. This could also explain Nanoindentation was used to detect the effects of aging on the sur­
why there was no significant effect of the in vitro clinically-related face hardness and modulus of the zirconia sub-types. The indenter

21719
B. Hajhamid et al. Ceramics International 49 (2023) 21716–21721

reading was obtained from non-defective regions by chewing simulation • The clinically-related aging model developed caused a decrease in
for each sample due to the machine sensitivity. The findings suggested both flexural strength and fatigue resistance of 3rd generation zir­
that there are similar hardness values as well as similar and modulus of conia, with no changes in the 2nd generation. In terms of mechanical
elasticity measures for both forms of zirconia after being exposed to the properties, the 2nd generation seems to be a better alternative as a
in vitro clinically-related aging model. There was no effect on the surface permanent restorative material.
hardness and modulus of extra high translucency in the different forms • Hydrothermal aging increased flexural strength and decreased the
of zirconia following hydrothermal aging. On the other hand, hydro­ surface hardness of 2nd generation zirconia. Hydrothermal aging
thermal aging decreased the surface hardness and modulus of HT zir­ also increased the fatigue resistance of both zirconia-based materials.
conia. Hydrothermal aging leads to volume expansion associated with t- • Hydrothermal aging failed to reproduce the changes imposed by
m transformation and lowers the hardness of the material due to low clinically-related aging failed to zirconia-based materials.
local atomic density [36]. A previous study reported increased surface
roughness of tetragonal zirconia after hydrothermal associated with Funding
grain pull-out, which may affect the indenter reading during loading and
unloading cycles [25]. In agreement with this study, a previous inves­ This work was funded by the National Sciences and Engineering
tigation demonstrated a significant decrease on the surface hardness and Research council of Canada (NSERC discovery grant #2018-04979).
modulus of 2nd generation zirconia following 15 h of autoclave hy­
drothermal aging and similarly there were no changes for 3rd generation
zirconia [37]. Declaration of competing interest
The study also hypothesized that mechanical properties will be
affected by material composition, which was partially accepted. The The authors declare that they have no known competing financial
results showed that HT zirconia has significantly higher strength interests or personal relationships that could have appeared to influence
compared to XT zirconia. This is mainly related predominantly to the the work reported in this paper.
crystalline composition of the different forms of zirconia as discussed
above. XT zirconia contains fewer tetragonal crystals than HT zirconia Acknowledgment
and this difference is likely to be playing a role in our findings suggesting
that XT zirconia had evidently inferior mechanical properties as The authors are thankful to XXXXX for their help in specimens
compared to HT zirconia. Previous studies showed that 3rd generation milling and sintering.
zirconia has lower strength than 2nd and 1st generations [10,37]. On the
other hand, surface hardness and elastic modulus were not affected by References
the material composition. There was no significant difference in surface
hardness and modulus between XT and HT zirconia. Even though the [1] C. Gautam, J. Joyner, A. Gautam, J. Rao, R. Vajtai, Zirconia based dental ceramics:
previous study reports correlation between the surface hardness and structure, mechanical properties, biocompatibility and applications, Dalton Trans.
45 (2016) 19194–19215. https://www.ncbi.nlm.nih.gov/pubmed/27892564.
crystalline size and the hardness, in which zirconia with smaller crys­ [2] R.C. Garvie, R.H. Hannink, R.T. Pascoe, Ceramic steel? Nature 258 (1975)
talline size exhibited higher surface hardness [38]. The crystalline 703–704, https://doi.org/10.1038/258703a0.
structure size and the indenter size may affect the result reported. The [3] S. Fabris, A stabilization mechanism of zirconia based on oxygen vacancies only,
Acta Mater. 50 (2002) 5171–5178, https://doi.org/10.1016/s1359-6454(02)
average size cubic and tetragonal crystals is 1 μm and the radius of the 00385-3.
indenter tip is 50 nm. The comparatively small size of the indenter may [4] F. Zhang, M. Inokoshi, K. Vanmeensel, B. Van Meerbeek, I. Naert, J. Vleugels,
result in testing the hardness and modulus at the grain boundary and Lifetime estimation of zirconia ceramics by linear ageing kinetics, Acta Mater. 92
(2015) 290–298, https://doi.org/10.1016/j.actamat.2015.04.001.
creating variability in the reported data.
[5] R.C. Souza, C.d. Santos, M.J.R. Barboza, C.A.R.P. Baptista, K. Strecker, C.N. Elias,
The current study investigated the effect of hydrothermal aging and Performance of 3Y-TZP bioceramics under cyclic fatigue loading, Mater. Res. 11
in vitro clinically-related aging simulating five years of clinical service (2008) 89–92, https://doi.org/10.1590/s1516-14392008000100017.
[6] H. Zhang, Z. Li, B.-N. Kim, K. Morita, H. Yoshida, K. Hiraga, Y. Sakka, Effect of
on two high translucency zirconia. This study aimed to replicate some of
alumina dopant on transparency of tetragonal zirconia, J. Nanomater. 2012 (2012)
the mechanical, thermal and acidic challenges occurring in the oral 1.
cavity. However, it did not reproduce real oral challenges as close as one [7] M.M. Manziuc, C. Gasparik, A.V. Burde, H.A. Colosi, M. Negucioiu, D. Dudea,
would like, since oral challenges such as biofilm formation, saliva Effect of glazing on translucency, color, and surface roughness of monolithic
zirconia materials, J. Esthetic Restor. Dent. 31 (2019) 478–485, https://doi.org/
buffering effect, and different dietary effects were not included. 10.1111/jerd.12493.
Furthermore, more effort should be made to simulate the clinical sce­ [8] E. Camposilvan, R. Leone, L. Gremillard, R. Sorrentino, F. Zarone, M. Ferrari,
nario in a laboratory setting such as using zirconia crowns instead of J. Chevalier, Aging resistance, mechanical properties and translucency of different
yttria-stabilized zirconia ceramics for monolithic dental crown applications, Dental
disc-shaped samples. By using zirconia in the shape of crowns, it might materials, Off. Publ. Acad. Dent. Mater. 34 (2018) 879–890. http://ovidsp.ovid.
be more relevant when considering the application of forces, particu­ com/ovidweb.cgi?T=JS&PAGE=reference&D=med15&NEWS=N&A
larly those that might be tangential to the unidirectionally applied forces N=29598882.
[9] J. Klimke, M. Trunec, A. Krell, Transparent tetragonal yttria-stabilized zirconia
used in this model system. Future studies might evaluate the effect of the ceramics: influence of scattering caused by birefringence, J. Am. Ceram. Soc. 94
aging model on additional physical and chemical properties of zirconia (2011) 1850–1858, https://doi.org/10.1111/j.1551-2916.2010.04322.x.
including but not limited to crystalline structure analysis, roughness and [10] A.A. Alshamrani, G.M. De Souza, Effect of ionizing radiation on mechanical
properties and translucency of monolithic zirconia, J. Biomed. Mater. Res. B Appl.
optical properties. Additionally, this study did not investigate the effects
Biomater. (2019). https://www.ncbi.nlm.nih.gov/pubmed/31410974.
of glazing or any other surface alterations required for provision of [11] J. Chevalier, What future for zirconia as a biomaterial? Biomaterials 27 (2006)
clinically acceptable crown restorations on the zirconia properties. 535–543. https://www.ncbi.nlm.nih.gov/pubmed/16143387.
[12] G.K.R. Pereira, S. Fraga, A.F. Montagner, F.Z.M. Soares, C.J. Kleverlaan, L.
Finally, the current study tested only two high translucency zirconia and
F. Valandro, The effect of grinding on the mechanical behavior of Y-TZP ceramics:
did not evaluate the effects of the aging, for example, on conventional a systematic review and meta-analyses, J. Mech. Behav. Biomed. Mater. 63 (2016)
and currently in use 1st generation zirconia. We also suggest that future 417–442. http://ovidsp.ovid.com/ovidweb.cgi?T=JS&PA
studies should be done to investigate the effects of the aging model on GE=reference&D=med13&NEWS=N&AN=27469603.
[13] A.M. Dal Piva, J.P. Tribst, L.D. Gondim, I.L. Ribeiro, F. Campos, A. Arata, R.
the zirconia bonding to human enamel. O. Souza, Y-TZP surface behavior under two different milling systems and three
different accelerated aging protocols, Minerva Stomatol. 67 (2018) 234–245. http
5. Conclusion s://www.minervamedica.it/en/getpdf/tOjzVmA0oUCj1aeYmKtPWaVBi5VjsXu
wcuZezuDI9H%252FZaVPGs7nLgJhqlWZgsOY64mfhb38wfOJMKGnUa88T9Q%
253D%253D/R18Y2018N06A0237.pdf http://ovidsp.ovid.com/ovidweb.cgi?
Within the limitations of this study. T=JS&PAGE=reference&D=emexa&NEWS=N&AN=625314145.

21720
B. Hajhamid et al. Ceramics International 49 (2023) 21716–21721

[14] S. Wille, P. Zumstrull, V. Kaidas, L.K. Jessen, M. Kern, Low temperature spectroscopic analysis, Materials 12 (2019). https://www.ncbi.nlm.nih.gov/pub
degradation of single layers of multilayered zirconia in comparison to conventional med/30875849.
unshaded zirconia: phase transformation and flexural strength, J. Mech. Behav. [28] C. Peampring, S. Sanohkan, Effect of thermocycling on flexural strength and
Biomed. Mater. 77 (2018) 171–175. http://www.elsevier.com/wps/find/journalde weibull statistics of machinable glass–ceramic and composite resin, J. Indian
scription.cws_home/711005/description#description. http://ovidsp.ovid.com/ov Prosthodont. Soc. 14 (2013) 376–380, https://doi.org/10.1007/s13191-013-0335-
idweb.cgi?T=JS&PAGE=reference&D=emed19&NEWS=N&AN=618260791. x.
[15] M. Cattani-Lorente, S. Durual, M. Amez-Droz, H.W.A. Wiskott, S.S. Scherrer, [29] A.D. Backer, E.A. Münchow, G.J. Eckert, A.T. Hara, J.A. Platt, M.C. Bottino, Effects
Hydrothermal degradation of a 3Y-TZP translucent dental ceramic: a comparison of simulated gastric juice on CAD/CAM resin composites-morphological and
of numerical predictions with experimental data after 2 years of aging, Dent. mechanical evaluations, J. Prosthodont. 26 (2017) 424–431. https://www.ncbi.
Mater. 32 (2016) 394–402, https://doi.org/10.1016/j.dental.2015.12.015. nlm.nih.gov/pubmed/26682954.
[16] M. Rosentritt, M. Behr, R. Gebhard, G. Handel, Influence of stress simulation [30] B. Hajhamid, L. Bozec, H. Tenenbaum, G. De Souza, E. Somogyi-Ganss, Effect of
parameters on the fracture strength of all-ceramic fixed-partial dentures, Dent. artificial aging on optical properties and crystalline structure of high translucency
Mater. 22 (2006) 176–182. https://www.ncbi.nlm.nih.gov/pubmed/16039706. zirconia, J. Prosthodont. (2023), https://doi.org/10.1111/jopr.13648.
[17] T.A. Sulaiman, A.A. Abdulmajeed, K. Shahramian, L. Hupa, T.E. Donovan, [31] P.H.C.O. Prado, K.S. Dapieve, T.M.B. Campos, L.F. Valandro, R.M. Melo, Effect of
P. Vallittu, T.O. Närhi, Impact of gastric acidic challenge on surface topography hydrothermal and mechanical aging on the fatigue performance of high-
and optical properties of monolithic zirconia, Dent. Mater. 31 (2015) 1445–1452. translucency zirconias, Dent. Mater. 38 (2022) 1060–1071. https://www.ncbi.nlm.
https://www.ncbi.nlm.nih.gov/pubmed/26494266. nih.gov/pubmed/35527035.
[18] H.T. Kim, J.S. Han, J.H. Yang, J.B. Lee, S.H. Kim, The effect of low temperature [32] F. Zhang, M. Inokoshi, M. Batuk, J. Hadermann, I. Naert, B. Van Meerbeek,
aging on the mechanical property & phase stability of Y-TZP ceramics, J Adv J. Vleugels, Strength, toughness and aging stability of highly-translucent Y-TZP
Prosthodont 1 (2009) 113–117. ceramics for dental restorations, Dent. Mater. : Off. Publ. Acad. Dent. Mater. 32
[19] Iso, Implants for Surgery—Ceramic Materials Based on Yttria-Stabilized Tetragonal (2016) e327–e337. http://ovidsp.ovid.com/ovidweb.cgi?T=JS&PAGE=reference
Zirconia, Y-TZP), 2008. &D=emed17&NEWS=N&AN=620374303.
[20] M.A. Ablal, J.S. Kaur, L. Cooper, F.D. Jarad, A. Milosevic, S.M. Higham, A. [33] R. Durkan, G. Deste Gökay, H. Şimşek, B. Yilmaz, Biaxial flexural strength and
J. Preston, The erosive potential of some alcopops using bovine enamel: an in vitro phase transformation characteristics of dental monolithic zirconia ceramics with
study, J. Dent. 37 (2009) 835–839. https://www.ncbi.nlm.nih.gov/pubmed different sintering durations: an in vitro study, J. Prosthet. Dent (2021). https
/19616357. ://www.ncbi.nlm.nih.gov/pubmed/34059297.
[21] J.A. Collins, Failure of Materials in Mechanical Design: Analysis, Prediction, [34] C.D. Holman, W. Lien, F.F. Gallardo, K.S. Vandewalle, Assessing flexural strength
Prevention, John Wiley & Sons, 1993. degradation of new cubic containing zirconia materials, J. Contemp. Dent. Pract.
[22] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and 21 (2020) 114–118. https://www.ncbi.nlm.nih.gov/pubmed/32381812.
elastic modulus using load and displacement sensing indentation experiments, [35] G.K.R. Pereira, L.F. Guilardi, K.S. Dapieve, C.J. Kleverlaan, M.P. Rippe, L.
J. Mater. Res. 7 (1992) 1564–1583, https://doi.org/10.1557/jmr.1992.1564. F. Valandro, Mechanical reliability, fatigue strength and survival analysis of new
[23] J. Chevalier, L. Gremillard, S. Deville, Low-temperature degradation of zirconia polycrystalline translucent zirconia ceramics for monolithic restorations, J. Mech.
and implications for biomedical implants, Annu. Rev. Mater. Res. 37 (2007) 1–32. Behav. Biomed. Mater. 85 (2018) 57–65. http://www.elsevier.
[24] M. Amaral, L.F. Valandro, M.A. Bottino, R.O. Souza, Low-temperature degradation com/wps/find/journaldescription.cws_home/711005/description#description.
of a Y-TZP ceramic after surface treatments, J. Biomed. Mater. Res. B Appl. http://ovidsp.ovid.com/ovidweb.cgi?T=JS&PAGE=reference&D=emexa&NE
Biomater. 101 (2013) 1387–1392. WS=N&AN=2000806638.
[25] G.M. De Souza, A. Zykus, R.R. Ghahnavyeh, S.K. Lawrence, D.F. Bahr, Effect of [36] S.A. Catledge, M. Cook, Y.K. Vohra, E.M. Santos, M.D. McClenny, K. David Moore,
accelerated aging on dental zirconia-based materials, J. Mech. Behav. Biomed. Surface crystalline phases and nanoindentation hardness of explanted zirconia
Mater. 65 (2017) 256–263. http://www.elsevier.com/wps/find/journaldescriptio femoral heads, J. Mater. Sci. Mater. Med. 14 (2003) 863–867. https://www.ncbi.
n.cws_home/711005/description#description http://ovidsp.ovid.com/ovidweb. nlm.nih.gov/pubmed/15348523.
cgi?T=JS&PAGE=reference&D=emed18&NEWS=N&AN=611995064. [37] N.M. Alfrisany, G.M. De Souza, Surface and bulk properties of zirconia as a
[26] P.A. Pinto, G. Colas, T. Filleter, G.M. De Souza, Surface and mechanical function of composition and aging, J. Mech. Behav. Biomed. Mater. 126 (2022),
characterization of dental yttria-stabilized tetragonal zirconia polycrystals (3Y- 104994. https://www.ncbi.nlm.nih.gov/pubmed/34864575.
TZP) after different aging processes, Microsc. Microanal. 22 (2016) 1179–1188. [38] Y. Gaillard, M. Anglada, E. Jiménez-Piqué, Nanoindentation of yttria-doped
[27] R. Sorrentino, C.O. Navarra, R. Di Lenarda, L. Breschi, F. Zarone, M. Cadenaro, zirconia: effect of crystallographic structure on deformation mechanisms, J. Mater.
G. Spagnuolo, Effects of finish line design and fatigue cyclic loading on phase Res. 24 (2009) 719–727, https://doi.org/10.1557/jmr.2009.0091.
transformation of zirconia dental ceramics: a qualitative micro-Raman

21721

You might also like