Sunteți pe pagina 1din 12

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.intl.elsevierhealth.com/journals/dema

Inuence of surface treatment on the


resin-bonding of zirconia
b , Erhan Cmlekoglu
b,
Seda Sanl a , Mine Dndar Cmlekoglu
Mehmet Sonugelen b , Tijen Pamir c , B.W. Darvell d,
a

Ulukent Dental Clinic, Menemen, Izmir, Turkey


Department of Prosthodontics, School of Dentistry, Ege University, Izmir, Turkey
c Department of Restorative Dentistry, School of Dentistry, Ege University, Izmir, Turkey
d Dental Materials Science, Faculty of Dentistry, Kuwait University, Kuwait
b

a r t i c l e

i n f o

a b s t r a c t

Article history:

Objective. To compare the effect of various surface treatments on the bonding of luting resin

Received 29 September 2014

cements to zirconia under four-point bending.

Received in revised form

Methods. Bar specimens (n = 200) (2 mm 5 mm 25 mm) were prepared from zirconia blocks

8 January 2015

(VITA In-Ceram YZ, Vita Zahnfabrik) with the cementation surface (2 mm 5 mm) of

Accepted 13 March 2015

groups of 40 treated in one of ve ways: airborne particle abrasion with 50 m Al2 O3


(GB), zirconia primer (Z-Prime Plus, Bisco) (Z), glaze ceramic (Crystall.Glaze spray, Ivoclar
Vivadent) + hydrouoric acid (GHF), fusion glass-ceramic (Crystall.Connect, Ivoclar Vivadent)

Keywords:

(CC), or left untreated as control (C). Within each treatment, bars were cleaned ultrasoni-

Zirconia surface treatment

cally for 15 min in ethanol and then deionized water before bonding in pairs with one of

Adhesive cementation

two luting resins: Panavia F 2.0, (Kuraray) (P); RelyX U-200 (3 M/Espe) (R), to form 10 test

Four-point bending test

specimens for each treatment and lute combination. Mechanical tests were performed
and bond strengths (MPa) were subject, after log transformation, to analysis of variance,
ShapiroWilk and HolmSidak tests; also log-linear contingency analysis of failure mode
distribution; all with = 0.05. Fracture surfaces were examined under light and scanning
electron microscopy.
Results. While the effect of surface treatment was signicant (p = 1.27 109 ), there was no
detected effect due to resin (p = 0.829). All treatments except CC (30.1 MPa / 1.44)* were
signicantly better than the untreated control (24.8 MPa / 1.35) (p = 3.28 109 ). While the
effect of GB which gave the highest mean strength (50.5 MPa / 1.29) was not distinguishable from that of GHF (39.9 MPa / 1.29) (p = 0.082), it was signicantly better than
treatment with either CC or Z (33.1 MPa / 1.48) (p < 0.05). (* After log transformation for
analysis and back; asymmetric error bounds as s.d. in log values.)
Signicance. The novel test method design, which has good discriminatory power, conrmed
the value of gritblasting as a simple and effective treatment with low operator hazard. It gave
the highest bond strengths regardless of the cement type. Glaze layer application followed

Corresponding author. Tel.: +44 1225 81 06 71.


E-mail address: b.w.darvell@hku.hk (B.W. Darvell).

http://dx.doi.org/10.1016/j.dental.2015.03.004
0109-5641/ 2015 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

658

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

by hydrouoric acid-etching on zirconia before cementation might be viable for adhesive


zirconia cementation, but represents a much greater hazard as well as having problems
with thickness control.
2015 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

Zirconia, or rather yttria-partially stabilized tetragonal zirconia (YTZP), is increasingly used in dentistry in view of its
remarkable strength [1,2]. However, a major impediment to
its effective use is its lack of reactivity since it is a non-polar
and more or less inert material which becomes very dense and
homogeneous on sintering [3]. This means that all the usual
means of bonding (in the broad chemical sense of established
a covalent bond between adhesive material and the substrate),
effective enough for other ceramic systems, do not work [3],
and although phosphate-based systems have shown some
improvement [3], nothing is yet considered reliable enough.
This also includes the mechanical retention afforded by gritblasting, where particle size has to be limited to minimize
subsurface damage and hence the risk of cracking [4]. Thus, it
is of great interest to develop an effective means of bonding
to zirconia to enable long-term retention. For this purpose, a
great number of surface treatment methods have been tried,
for example to roughen the surface with rotary systems, laser
irradiation, or selective inltration etching, or to modify the
surface through silica coating, silane application, hydrouoric
acid etching after fused glass-ceramic application, or nanostructured alumina coating [3,57].
The nature of dentistry, in terms of economy, processing
time, and the fact that all restorative devices are effectively
one-off prototypes, means that the clinical and laboratory procedures to be used for zirconia-based polycrystalline ceramic
restorations to achieve adhesive cementation should be practical and easily handled. We may note that while hydrouoric
acid (HF) etching is an easy if hazardous surface treatment
for silicate-containing ceramics [8], some oxide ceramics, and
in particular zirconia, which contain less than 15 mass% silica
and little or no glass phase [9], cannot be so treated. However, the application of a fused glass-ceramic on the intaglio
surfaces of zirconia restorations, which glaze is then etchable
and so capable of adhesive cementation, has been described
[5,10,11]. The effectiveness of this bonding depends on intimate contact and micromechanical key with grain boundary
topography, rather than chemical interaction as has sometimes been suggested [12], there being no chemical reaction
or elemental migration or diffusion at the interface [13]. Some
studies suggest that compressive stresses arising from a difference in coefcient of thermal expansion (CTE) between a
ceramic and zirconia have a positive correlation with the bond
strength [14]. This is a general principle that should apply
in the present context. Ordinary silane chemistry also may
not improve the afnity of resin cements for zirconia since
it is more stable then silica-based ceramics and cannot be
hydrolysed easily [3]. Zirconate coupling agents have been
introduced with the aim of improving the bonding of zirconia

to resin cements [3]. It has been reported that the phosphate


groups of one of these primers, which is organophosphatecarboxylic acid monomer based, bond to metal oxides such as
zirconia, while the organic moiety can be co-polymerized with
the monomers of lled resins and therefore might be feasible
as a surface treatment for zirconia [15].
The great variability in the data obtained from some studies
of dental materials stems from a variety of reasons: inappropriate set-ups for the relevant clinical problem, unsuitable
test methods, inappropriate specimen geometries, as well as
several variables that cannot be standardized [16,17]. Bond
strength is commonly treated in direct tension since brittle
dental materials are more sensitive to tensile stresses than
compression [18], but there are great difculties in doing this
with ceramics because of that brittleness: risk of introduction
of surface aws during the machining of the specimens, and
parasitic stresses introduced in mounting small misalignments are enough to be problematic [17,19]. Tensile strength
as such is then often determined by a bending (or exural)
test [20], which is less sensitive to such problems, although
alignment of the supports remains critical. Shear testing is
associated with much greater difculties [17,21]. Simple shear
test designs, as commonly used in dental research, do not
give a uniform stress at the interface, especially with elastic
modulus mismatch, and with stress concentrations at contact
points [22].
Flexural strength can be determined using three- and
four-point loading. However, the two modes commonly yield
different results [23]. Although the equations for the two
approaches consider the differences in load application
and stress distribution, this difference between results is
explained by the volume of the test piece affected: in a threepoint test the effective volume of peak stress (or location for
a critical aw) is very small and lies under the center loading point, while in a four-point test the region between the
inner span loading points (which may be half the outer span
length) is uniformly stressed in pure bending such that the
risk of a large critical aw being present in the critical zone
is much greater. However, if a bending test is to be used
instead of direct tension for a bonding effectiveness assessment, logically the bonded surface must be central. However,
in three-point bending this then requires great (and, routinely, essentially unobtainable) precision with the alignment
of the surface directly under the central load roller, but if the
substrate-superstructure model is used, whether in a threeor four-point test, the specimen is then asymmetrical in modulus of elasticity and so in deection, and thus the stress state
is not accurately represented by the usual equations [24,25].
Furthermore, to place a load singularity on the adhesive layer
is itself problematic. Despite this, such asymmetric systems
have been used in four-point bending of zirconia-resin-dentin
and zirconia-resin [24,26]. Thus, although it could be argued

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

that concentrated loading at the interface (i.e. in tension on


the lower surface) is relevant there would be practical difculties. This is borne out perhaps by the large coefcients of
variation found with this approach, some 20 28% [24].
Elementary beam theory is based on the idea that the
geometry of the system is unaffected under load. In addition,
the shear stress in the specimen between the outer and inner
support(s) contributes to the deection, and if this is large it
must be taken into account [20], especially for deection calculations [27]. Even slight malalignment in the three-point case
will thus superimpose a shear stress on the bonded region
which will compromise the interpretation of the results, and
this applies especially when the bonded region has a denite
thickness (as it must), and especially with intermediary layers:
the stress causing failure is unknown. In any case, a uniform
stress state is required in order to identify the weakest link,
and this cannot be attained in three-point bending. This does
not apply in the four-point case. To control shear deformation,
and thus without correction, a span to depth ratio (S/d) of 16
is considered acceptable for most materials, although some
require S/d = 3264 in four-point tests [27]. For dental ceramics,
the recommended ratio is 10:1 or greater to provide consistent
strength values [28]. Four-point bending is used for porous [29]
materials such as spinel, zirconia and graphite, according to
ASTM C167411, a standard exural test method for advanced
ceramics with engineered porosity (e.g. honeycomb cellular
channels) at ambient temperatures.
The goal of laboratory testing ought to be to provide interpretable data through a mode of loading relevant to the service
conditions. Certainly, it is difcult to envisage a pure tensile stress state in any oro-dental system (likewise for pure
shear, a fortiori). On the other hand, bending, especially at the
pontic of a xed dental prosthesis (FDP), is much more common, and occlusal loads there cause traction forces that have
been reported to cause a lever effect on the terminal edges of
the restorations [30] rather similar to the stress state in the
load axis plane of a exural test specimen. This might therefore provide an appropriate model for simulating (or at least
representing) intraoral loading conditions. Indeed, a exural
strength test is widely used and proposed for strength testing
for ceramics [27], although the symmetrical four-point mode
has not previously been used for dental ceramic-resin bond
strength testing.
In microtensile tests, since the resin, treated surfaces and
zirconia have differing elastic moduli, Poisson ratios and
strengths, the materials differing behaviors mutually interact and a pure, uniform stress eld is unattainable [25]. This
situation cannot be controlled. In the usual so-called shear
tests, uncontrolled and unmeasurable parasitic stresses occur
since the bending of the material cannot be prevented despite
all the precautions taken in test designs [17,21,31]. In contrast,
three-point bending and microtensile tests are recommended
for homogeneous materials, while four-point bending tests
should be preferred for heterogeneous materials, where feasible [25].
The primary motivation for the present work was to be
able to study the behavior of the interface between zirconia and resin cement or surface treatments (and, secondarily,
between resin cement and surface treatment). The union
between adhesive and tooth tissue was not of concern and

659

thus could be eliminated from the test system. This suggested


that a symmetrical zirconia-surface treatment-resin cementsurface treatment-zirconia test system could be used. Given
then that all interfaces should be loaded identically, and also
given the problems of direct tension and shear, four-point
bending appears to provide all necessary conditions.
Thus, with a view to identifying an effective procedure
for bonding zirconia, and a suitable test method free of common problems, the effect of various surface treatments on the
strength of the union to luting resins was investigated using
a symmetrical specimen, bonding zirconia to zirconia, using
four-point bending, enabling easier fabrication and simpler
interpretation of results.

2.

Materials and methods

2.1.

Specimen preparation

Presintered zirconia blocks (Vita In-Ceram YZ-40/19; Vita Zahnfabrik, Bad Sckingen, Germany) were cut with a low-speed
(400 rpm) diamond saw under water cooling (Isomet 1000;
Buehler, Lake Bluff, IL, USA) to give bars with the dimensions
of 31.25 mm 6.25 mm 2.50 mm (each aspect measured at
three places with a digital caliper) using an acrylic jig. These
were then sintered at 1530 C for 7.5 h in a furnace (MOS
160/1; Protherm, Ankara, Turkey) giving nal dimensions
of 25.0 mm 5.0 mm 2.0 mm after shrinkage. These dimensions were decided according to previous remarks on S/d ratio
[27,28] for this test method [29], given the size of the zirconia
blocks.
The zirconia specimens were then ultrasonically cleaned
(Sonorex RK102 Transistor; Bandelin, Walldorf, Germany) for
15 min in 96% ethanol (Smyras, Bornova, Izmir, Turkey) and
15 min in distilled water to ensure the absence of particulate debris immediately before surface treatment. Calculation
indicated that 10 specimens per subgroup would give a power
>90%.

2.2.

Surface treatments

The following treatments were used, with 20 pieces per group


for n = 10 luted assemblies.

2.2.1.

Control (C)

No surface treatment.

2.2.2.

Gritblasting (GB)

Gritblasting (EasyBlast; Bego, Bremen, Germany) was performed normal to the luting surface with 50 m Al2 O3 (Shera
Werkstoff Technologie, Hannover, Germany) for 13 s at a pressure of 2.8 bar from a distance of 10 mm [32].

2.2.3.

Glaze and hydrouoric acid etching (GHF)

The luting surface received two coats of sprayed glaze (Crystall


Glaze spray; Ivoclar Vivadent, Schaan, Liechtenstein) yielding a continuous, thin layer. Glaze ring was conducted with
a ceramic furnace (Programat P90; Ivoclar Vivadent), following the manufacturers detailed program, but broadly this
involved starting at 403 C, the temperature then being raised

660

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

at 60 C/min to 725 C, which was held for 1 h before cooling. Glaze that extended beyond the cementation surface was
removed, by grinding at on 600-grit SiC paper under water,
to avoid any supportive effect during the fracture test.
The glazed surfaces were then treated at room temperature with 9.5% hydrouoric acid gel (Porcelain Etch; Ultradent
Products, South Jordan, UT, USA) for 60 s, rinsed with deionized water for 90 s. They were then neutralized with a slurry
of neutralizing powder (CaCO3 , Na2 CO3 ) (IPS Ceramic Neutralizing Powder, Ivoclar Vivadent) for 5 min, washed thoroughly
for 20 s with distilled water, and air dried [5].

2.2.4.

Fusion glass-ceramic (CC)

To mix the material (Crystall.Connect; Ivoclar Vivadent), the


supplied closed capsule was lightly pressed onto a vibrating
plate (Ivomix; Ivoclar Vivadent) for 10 s for agitation. The mixture was applied with a brush tip on the cementation surface,
the specimen held against the vibrating plate for 5 s, and then
allowed to dry for 60 s. After cleaning away the excess material
with a clean brush tip, ring was conducted using a ceramic
furnace (Programat P90; Ivoclar Vivadent), again following the
manufacturers detailed program; broadly, this involved starting at 403 C, the temperature then being raised at 30 C/min
to 820 C for 2 min, then to 840 C for 7 min before cooling. The
prepared luting surface was then etched, washed, and dried
as above.

2.2.5.

Zirconia primer (Z)

After cleaning the cementation surfaces by rinsing with deionized water and air drying, two coats of primer (Z-Prime Plus;
Bisco, Schaumburg, IL, USA) were applied uniformly, wetting
the bondable surface, and dried with an air syringe for 35 s to
remove solvent, according to the manufacturers instructions.

2.3.

Luting

The zirconia bars were luted end-to-end on the prepared


surfaces with one of two chemically distinct luting resins
(Panavia F 2.0: Kuraray, Tokyo, Japan [P]; Rely-X U 200: 3 M
ESPE, St. Paul, USA [R]) (Table 1). A custom-made stainless steel
mold was used to standardize the luting material thickness
to 0.10 0.05 mm. Specimens were placed in the mold end to
end having applied the cement to one end of each using a
cotton pellet, the molds screw was tightened to approximate
the pieces and decrease the cement gap thickness to the target value, as indicated by the separation of the mold faces. A
4 mm gap between the molds approximating faces was left to
allow for the removal of excess cement. Prior to the cementation the length of each half-specimen (25 mm) was measured
before placing in the mold. The gap between the molds edges
was measured using a digital caliper to calculate the closure
required. After closure, excess cement was removed and the
mold face separation rechecked.
For both resins, catalyst and base pastes in equal
amounts were mixed according to the manufacturers instructions, applied to both luting surfaces of the zirconia bars,
which were then brought together. Under magnication
(loupe, 20), excess cement was removed using microapplicators (Disposable micro applicators, ne, Premium Plus
International, Hong Kong, China) to avoid dragging the resin

from the interface. For resin P, the material was then covered
with an oxygen-inhibiting material (Oxyguard; Kuraray, Tokyo,
Japan), and irradiated for 20 s each from the top and the bottom
(LED Bluephase C5; Ivoclar Vivadent),. Resin R was similarly
irradiated from top and bottom for 10 s each, to prevent movement, then left undisturbed for 5 min (25 C, 5065% relative
humidity) to complete self-curing.

2.4.

Four-point bending tests

Specimens were mounted with the luted interface centralized


between the upper load points. The load was applied through
rods of radius 2.0 mm at a crosshead speed of 0.5 mm/min on
a computer-controlled universal testing machine (Autograph
Model AG-5 kN; Shimadzu, Kyoto, Japan). The load at fracture
was recorded. Four-point exural strength ( 4 ) was calculated
from:
4 =

3F(L Li )
2wh2

where F is the load at fracture, Li is the distance between


the centers of the loading rollers (20.0 mm), L is the distance
between the centers of the supporting rollers (40.0 mm), w is
the specimen width, and h the mean thickness (2.0 mm) [29],
giving S/d = 20.

2.5.

Fractographic analysis

After fracture, the surfaces of the specimens were evaluated


under light microscopy (LM) (Eclipse ME600, Nikon, Melville,
NY, USA) at 10 magnication. Failure was classied as: adhesive (a) more than 85% between the framework and the
cement; cohesive (c) more than 85% debonding within the
zirconia, lute, or both; and mixed (m) other combination.
This was on the grounds that neither adhesive nor cohesive
failure was seen pure, and the arbitrary cut was taken to try
to assess the dominant behavior.
Two specimens for each group were taken for further
analysis of the fractured surfaces using Environmental Scanning Electron Microscope-Energy Dispersive Spectroscopy
(ESEM-EDS) (Quanta 250 FEG SEM, FEI, Hillsboro, OR, USA).
Low-vacuum imaging mode was used, with no need for a conductive specimen coating. Specimens were evaluated under
56 and 10,000 magnications and analysed with energydispersive X-ray spectroscopy (EDS) at 43 magnication and
10 keV.

2.6.

Statistical analysis

Flexural strengths were subjected to two-way analysis of


variance (2 AoV), Resin Surface Treatment, subsequently
reducing to 1 AoV on Surface Treatment, with checks
for homogeneity of variance and normality (ShapiroWilk)
(SigmaPlot 12.5, Systat Software, San Jose, CA, USA). Comparisons were then made using the HolmSidak method.
Fracture-type distribution was examined by means of loglinear analysis of the three-way contingency table (3 CT) [33],
then by exact 2 (StatXact v.9, Cytel Software, Cambridge, MA,
USA) on the then-indicated collapsed 2 CT.

661

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Table 1 Manufacturers information on products used.


Product

Manufacturer
a

Panavia F 2.0

Kuraray Medical, Tokyo,


Japan

Composition description/mass%

Batch

A paste: BPEDMA/MDP/DMA
B paste:
*
Ba-* B-Si-glass/silica-containing
composite
Glycerin gel

00037A
00020A

3 M/Espe
Seefeld, Germany

Glass powder, silica, calcium


hydroxide, pigment, substituted
pyrimidine, peroxy compound,
initiator
Methacrylated, phosphoric esters,
dimethacrylates, acetate,
stabilizers, self-cure initiators,
light-cure initiators

315751

Vita In-ceram
YZ-40/19
Zirconia Block

Vita Zahnfabrik, Bad


Sckingen, Germany

9194%* ZrO2 , 46%* Y2 O,


24%* HfO2 , <0.1% Al2 O3 , <0.1%
SiO2 , <0.1% Na2 O

ECYZ40192

IPS e-max. Crystall/Connect


(Fusion glass-ceramic)

Ivoclar, Vivadent, Schaan,


Liechtenstein

SiO2 60.065.0%, Al2 O3 6.010.5%,


Na2 O 6.011.0%, * K2 O 15.019.0%,
ZnO 1.03.0%, other oxides
5.530.0, Pigments 0.10.5%,
Powder-liquid system

P56200

IPS e-max Crystall.Glaze


spray glaze

Ivoclar, Vivadent, Schaan,


Liechtenstein

4060% powder (6065%


SiO2 ,1519% K2 O, 610, 5% Al2 O3 ,
5.530% other oxides and
pigments), 1520% propanol,
2040% isobutane (propellant),
butanediole

P56200

Z-Prime Plus
(Zirconia Primer)

Bisco, Schaumburg, IL, USA

<10% biphenyl dimethacrylate,


<20% hydroxyethyl
dimethacrylate, <90% ethanol

B-6001P

Ultradent
Porcelain Etch

Ultradent Products; South


Jordan, UT, USA

9.5% hydrouoric acid

C123

Oxyguard II
b

Rely-X U200
Base
(Clicker Dispenser)
Rely-X U200
Catalyst
(Clicker Dispenser)

00471A

BPEDMA: Bisphenol-A-polyethoxydimethacrylate; MDP: 10-methacryloyloxy-decyl-dihydrogenphosphate; DMA: aliphatic dimethacrylate.


Elements used for EDS analysis.
a
Equal amounts of A and B pastes mixed, applied, excess cement cleaned, covered with Oxyguard II and light-cured 20 s each from the top and
the bottom.
b
Equal amounts of base and catalyst pastes mixed, applied, excess cement cleaned and light-cured for 10 s each from the top and the bottom;
wait for self-curing.

3.

Results

3.1.

Four-point bending

For the 2 AoV, although the data passed the homogeneity of


variance test (p = 0.712), they failed the normality test (p < 0.05).
This was possibly due to the correlation between variance
and mean (r2 = 0.3528, n = 10, F = 4.36073, p = 0.0702), a common
result for strength measures. Accordingly, the data were transformed by taking logarithms and the analysis repeated. Both
homogeneity of variance (p = 0.804) and normality (p = 0.180)
tests were then passed.
While the effects of Surface Treatment were highly significantly different, there was no detected effect due to Resin
(Table 2). The apparently signicant interaction can therefore be ignored as uninterpretable. Accordingly, the results
were pooled by Resin and a one-way analysis conducted for
Surface Treatment, again on the transformed data. Homogeneity of variance (p = 0.202) and normality (p = 0.180) tests

were again passed. The power of the test was reported as


0.999999. The effect of Surface Treatment remained highly
signicant (Table 3).
The comparisons are set out in Table 4 and illustrated in
Fig. 1. We may note that all treatments except CC were signicantly better than the untreated control. While the effect
of GB which gave the highest mean strength was not distinguishable from that of GHF, it was signicantly better than
treatment with either CC or Z.

3.2.
Failure mode and environmental scanning
electron microscopy
The distribution of failures modes is shown in Table 5. In
the 3 CT analysis (Table 6) the interaction (or lack thereof)
between Resin and Surface Treatment has been ignored as the
design was balanced (i.e. not a random distribution of specimens by Resin and Surface Treatment). Thus, the effect of

662

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Table 2 Analysis of variance of four-point bending strength: resin surface treatment (log-transformed data).
Source of variation

DF

SS

MS

Resin
Treatment
Resin treatment
Residual

1
4
4
90

0.000841263
1.108007
0.209951
1.620414

0.000841263
0.277002
0.0524878
0.0180046

0.0467
15.385
2.915

Total

99

0.829
1.27 109
0.0256

Table 3 Analysis of variance of four-point bending strength: surface treatment (log-transformed data).
Source of variation

DF

SS

MS

Treatment
Residual

4
95

1.108007
1.831206

0.277002
0.0192759

14.370

Total

99

2.939213

p
3.28 109

Table 4 HolmSidak comparisons of means of four-point bending strength by surface treatment.


Comparison

Diff. of means

GB vs. C
Z vs. C
GHF vs. C
CC vs. C
GB vs. GHF
GB vs. CC
GB vs. Z
GHF vs. CC
GHF vs. Z
Z vs. CC

0.308124
0.124909
0.205230
0.0838744
0.102894
0.224250
0.183215
0.121355
0.0803203
0.0410349

7.018
2.845
4.674
1.910
2.344
5.108
4.173
2.764
1.829
0.935

<0.000001
0.0322
0.000078
0.167
0.0821
0.000015
0.000466
0.0338
0.136
0.352

p < 0.05
Yes
Yes
Yes
No
No
Yes
Yes
Yes
No
No

Table 5 Distribution of failure modes: resin surface treatment mode.


Resin:

Treatment

Total

Total

C
GB
Z
GHF
CC

7
0
1
1
0

3
6
8
9
7

0
4
1
0
3

10
10
10
10
10

6
1
5
5
4

4
9
5
5
2

0
0
0
0
4

10
10
10
10
10

Total

33

50

21

25

50

Surface treatment mode.


within P: 2 = 30.8, 8 df, exact p = 8.729 105 .
within R: 2 = 24.7, 8 df, exact p = 9.989 104 .

Surface Treatment on failure Mode is highly signicant, while


that of Resin is weaker but still signicant.
Environmental scanning electron microscopy (ESEM)
images are shown in Fig. 2. For both resins in the control groups (C), there was a sharp, distinct boundary
between residual cement and the visible zirconia, while for

GB the boundary was ragged with clear evidence of an


interpenetrating interface. In contrast to C, for GB, EDS showed
the presence of elements from the cements (Na, Si, Ba) on
the apparent zirconia surface, indicating a degree of cohesive failure of the resin visually undetectable. For GHF, a fused
glassy surface was present with patches of zirconia showing,

Table 6 Three-way contingency table analysis (3 CT) of failure mode distribution: log-linear analysis results (G2 )
compared with exact 2 (Treatment = surface treatment).
Treatment mode
Resin mode
Treatment mode|resin
Resin mode|treatment
a
b

i.e. pooled resins.


i.e. pooled treatments.

G2

df

2

p (exact)

34.92
7.42
52.94
25.44

8
2
16
10

<1 104
0.0245
<1 104
0.0046

32.57a
7.24b
(Not possible)
(Not possible)

5.47 105
0.0254

663

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Table 7 Failure type discrepancy evidence. Y means: (under A) in an area that appears to be simply an adhesive
failure, there is evidence of remnants of cement; (under C) in an area that appears to be simply a cohesive failure, there is
evidence of the underlying zirconia showing through. N means: not detected in the elds examined. E: as determined
by high magnication ESEM examination. S: as determined by EDX spectrum.
Resin

Treatment
C
GB
Z
GHF
CC

A
Y (E, S)
N
Y (E)
Y (E,S)
N

suggesting an interpenetrating network of lute, glaze and zirconia for both resins. For CC, sharply distinct areas of the
lute, glaze (with visible voids), and the zirconia surface were
found, with apparently relatively clean separation at these
interfaces, while for Z remnants of primer and cement were
found, EDS showing Si from the cement, indicative again of
a degree of cohesive failure of the cement as well as adhesive failure between cement and primer. The results of the
ESEM and EDS checks of the categorization of failure mode
are shown in Table 7, indicating where contrary evidence was
found.

4.

Discussion

The symmetrical four-point test was found to be a straightforward and effective means of testing bond strength, with the
coefcient of variation for the log-transformed data (8.6 2.0)
being acceptable for this class of test (Fig. 1, top) and indicating useful discriminatory power. Technically, the set-up
could be improved to control more precisely the lute or cement

Fig. 1 (bottom) Four-point bending test fracture stress


results for the two resins, with means and standard
deviation error bars calculated from log values for the
pooled resins. (top) Coefcient of variation for log data by
resin and treatment.

R
C
Y (E)
Y (E,S)
Y (E,S)
N
Y (E,S)

A
Y (E,S)
Y (E,S)
Y (E)
Y (E,S)
Y (E,S)

C
Y (E)
Y (E)
Y (E,S)
Y (E,S)
Y (E)

thickness, as would be true elsewhere [24,26], although it is


noted that in clinical practice such control is also problematic
[24,26] so that some variation is realistic. It is also a conservative test in that two interfaces are loaded, with the weakest
controlling the outcome.
It would seem self-evident that bond-strength testing
should use a system that properly stresses the adhesive interface [34], yet while shear has been imagined to represent a
relevant failure mode, it has not yet been shown, to our knowledge, that this occurs in practice. Even so, shear tests have a
number of problems, including inhomogeneity of the stress
eld and blatant stress concentrations and parasitic stresses
[18,22,25,34,35]. In contrast, it has been said that interfacial
tension is preferable [18]. However, direct tension also has difculties in specimen preparation, alignment and xation of
the specimens, and so again a tendency for an inhomogeneous
stress distribution [25,34]. Flexural strength tests are, in contrast, relatively easy to set up (complex specimen preparation
increases the likelihood of damage), and specimen xation
problems are avoided [3537]. Nevertheless, it needs to be recognized that the stress eld in the load axis is a gradient, with
the tensile maximum on the convex surface. We would argue
now that this is more likely to represent the clinical situation
than the improbable direct (uniform) tension and the frankly
impossible pure shear. As we argue above, three-point testing
is problematic because of the small critical volume in relation to the (intentionally) narrow interfacial region. The pure
bending zone of a four-point test [35] reduces the criticality of
(lateral) alignment and removes the stress singularity caused
by the central load-point of the three-point test [36]. That is
not to say that the need for care is removed [28]. Specimen
and testing apparatus designs vary in existing standards (e.g.
DIN ENV 843, ISO 6872, ASTM C167411) for this test [29,37].
Therefore, the properties of the material to be tested should
be very well known. Here, ASTM C167411 was preferred [29].
Although zirconia for dentistry may contain from 0 to 15%
silica [38] or other glass phase, becoming very dense and
homogeneous on sintering [9], conventional adhesive procedures cannot work [3] because being very unreactive it is not
etched by phosphoric or hydrouoric acids and so cannot
provide any roughness for simple micromechanical adhesion
[3,911], unlike gritblasting. A silica coating, however, is reactive and etchable, while zirconia primer (see below) has
been said to increase chemical adhesion [3]. Although the
usual resin cements provide good mechanical retention on
rough surfaces, those containing phosphate monomer are
preferred [3] for zirconia since they are said to be hydrolytically stable in the long term because of the strength of the

664

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Fig. 2 ESEM images for typical fracture surfaces, low and high magnications.

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

chemical interaction between the phosphate ester monomers


of 10-methacryloyloxydecyl dihydrogen phosphate (MDP) and
hydroxyl groups on the zirconia surface [3,39]. On the other
hand, the lm thickness of MDP-containing resin cements is
twice that of other resin cements and this is a disadvantage
for adaptation [40,41]. It was thought, therefore, that the resins
tested here, Panavia F 2.0 (with MDP), and Rely-X U2001 (with
an alternative phosphoric ester methacrylate monomer), were
most appropriate for the present study.
In a study on the effect of metal primers on the microtensile bond strength to zirconia of phosphate-containing
resin cements (Panavia F: 8.8 5.1 MPa; Rely-X Unicem:
7.2 3.2 MPa), the difference was not signicant [39]. Likewise,
an asymmetric four-point bend test of lled-resin composite
to gritblasted (125 m alumina) zirconia found no significant difference between three resin cements (Panavia F:
43.9 12.2; RelyX Unicem: 41.3 8.1 MPa; Multilink Automix:
39.0 8.9 MPa) [24]. These results are echoed now (in terms
of relative magntude) in symmetric four-point bend, but with
rather smaller coefcients of variation.
From the results, we may say that simple gritblasting (GB)
was the most effective treatment because of its simplicity
(given the complexity and hazard of GHF), while the fusion
glass-ceramic coating (CC) failed to achieve any detectable
improvement. Gritblasting with alumina has been widely
studied. It has given good results with MDP-containing cement
[42], has been reported to increase signicantly resin-zirconia
bond strength in a shear test, whether before or after the
sintering stage (4.5 to 6.5 MPa) [43], and with priming also
to increase tensile bond strength, and to improve long-term
bonding, using both high- and low-pressure blasting [44]. Gritblasting is intended to remove contamination and increase
the roughness and hence the specic surface area, thereby
decreasing the effective contact angle and increasing the wettability of the luted material [32,45,46]. While GB was found to
give the best bond strength here, there is a caveat. It is known
that this can induce the tetragonal monoclinic phase transformation, despite stabilization [47]; the affected layer may be
some 10 m thick [48]. While the conditions now used are not
far removed from the optimum indentied then [47], it will
require further work to ascertain what the implications are
for bond strength. It would appear that the present method
would provide the necessary discriminatory power.
If surface hydroxyl groups are important for bonding with
phosphate-containing cements, the increased specic area
of a blasted surface should also be benecial, even without
appreciable micromechanical key [49].
However, gritblasting has been reported to cause impactinduced aws which may affect the longevity of zirconia
restorations [24,50], although this may be controlled to some
extent by decreasing the blasting pressure and grit size [34,41].
Even so, shear-bond strength has been reported to increase
[51], while biaxial exural strength has been found to increase
along with the monoclinic phase induced by gritblasting, and
decrease likewise when the monoclinic phase was removed
by heat treatment [48]. However, while these initial cracks
may have little or no effect, under cyclic loading such cracks

This product is known in other regions as RelyX Unicem 2.

665

propagate and have been found to decrease bond strength by


2030% [49,52,53].
For zirconia-supported xed partial dentures (FPDs), phosphate monomer-containing primers, bonding agents and
zirconate binding agents have been proposed, intended to give
a similar effect as silanes on silica-based ceramics [15,49].
A copolymerizable functional group is supposed to establish
continuity with the resin system. Magne et al. investigated an
experimental primer of this type in a shear test and reported
that with Panavia F it gave the highest strength [15]. A similar
treatment here (Z) gave a marginal effect.
A simple fused ceramic glaze layer is meant to inltrate
roughness and porosity, but its adhesion to zirconia depends
only on van der Waals and electrostatic forces [54]. Etching this
glass with hydrouoric acid (to which zirconia is not susceptible) then gives a micromechanically retentive surface, said
to yield a durable and stable resin bond [55,56], which surface may also be treated with silane. This has been reported
to decrease low-temperature degradation and damage [56],
hence the inclusion now of such a process. Derand et al.
reported signicantly increased bond strength compared with
the non-treated surface for such a treatment [54], a result now
echoed (GHF). However, while Ntala et al. found that a lithium
disilicate glaze was better than gritblasting [57], here there
was no signicant difference between GB and GHF, albeit with
a feldspathic ceramic. The differing results might stem from
either test mode, glaze chemistry, or that fact that the glaze is
two-phase, or all three.
The thickness of the glaze layer has ranged between 20
and 40 m [58,59], although a new approach reduced this to
some 79 m [5]. For comparison, the cement thickness for
FPDs with current CAD-CAM systems has been reported to
be 3050 m, but 39502 m for zirconia-supported devices,
with marginal openings of 8272 m [60]. While the glaze
layer for GHF would not affect internal adaptation, the thickness should be carefully controlled for zirconia prostheses
(CAD-CAM may offer a means of doing this in practice
[61]). Here, control of the glaze for CC was compromised
since the wet mixture showed shear-thinning under vibration. A reformulation or a method better to control this is
required.
Considering the distribution of fracture types, that mode
was affected by treatment can be expected, but the signicant
interaction Resin Mode|Treatment is noteworthy (Table 6). It
can be seen from Table 5 that the proportion of adhesive failures for resin R is appreciably higher. Whether this conrms
the value of MDP (as in resin P), or identies some deciency
in resin R, requires further investigation.
Thus, using LM to estimate percentage area coverage in
order to classify the failure type is a simple enough procedure for this kind of relatively crude analysis of the behavior of
cements. However, it is clear from the detection by EDS of remnants of cement in areas that were apparently plain adhesive
failure (that is, seeming to be clean zirconia substrate), when
viewed by LM or low-magnication ESEM, that such simple
categorization may be misleading (Table 7). Thus, for example, in the case of P-Z, remnants of cement were found in areas
that had apparently failed adhesively, while in the case of RGHF zirconia was found cleanly exposed when cohesive failure
of the cement had been the LM and low-magnication ESEM

666

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

Table 8 Comparison of t of failure data to Normal and


Weibull Probability plots (r2 values).
Ordinate:
abscissa:
C
CC
GB
GHF
Z

Normal
linear

Normal
log

Weibull
log

0.9334
0.8815
0.9299
0.9415
0.9255

0.9812
0.9613
0.9838
0.9709
0.9497

0.9071
0.8839
0.9076
0.9263
0.8884

conclusion, i.e. a more mixed condition. Thus, LM alone may


be unreliable in this respect.
The thickness of the glass-ceramic layer (CC) is clear while
that for the glazed specimens, GHF was both thinner and
adherent to both the substrate and cement (Fig. 2). Both of
these treatments involve a high-temperature process, GHF:
725 C max, CC: 840 C max. The thermal properties of the
coatings are similar [62] but it has been suggested that such
heat treatment could decrease the strength of the zirconia [63].
This might account for the rather more disrupted appearance
of the fracture surfaces in these cases, which effect would be
worth further study.
The distribution of failures regarding the surface treatment procedures support the bond strength test ndings (C:
lowest  4 : 26 MPa and highest adhesive failure: 65%; GB: highest  4 : 52 MPa and lowest adhesive failure: 5%; Z, GHF and
CC gave similar  4 values and adhesive and mixed failure
rates) with the exception that cohesive failures for fusion
glass-ceramic (CC) were more frequent (35%). This nding
points to this material being a potential alternative for zirconia surface treatment for enhancing adhesive cementation
if the thickness concern can be solved and its strength
improved.
The need to use the log transformation to stabilize the variance is, as indicated above, often necessary for strength data.
This is shown quite clearly in Fig. 3 (top), where a sensible
approach to linearity is found (with the possible exception
of CC), which linearity is lost on an untransformed abscissa.
Further, it is commonly assumed that that the Weibull distribution applies to such strength data in general; it is plain
from Fig. 3 (bottom) that this is not the case here the curvature is pronounced for each surface treatment. Comparison of
the coefcients of determination in each case bears out both
points (Table 8) where the Normal distribution for log data
is uniformly the best of the three, and the Weibull plot the
worst except for CC. It has been said that it is often difcult
to discriminate between the distributions for small sample
sizes (i.e. <50) [64], but here it is apparent at n = 20. Thus,
in two respects, it is important always to check and validate
distributional assumptions. In that light, inspection of Fig. 3
(top) shows the behavior of both CC and Z to be odd: whilst
overall both normality and homogeneity of variance tests
were passed, it would be appropriate to investigate whether
these data represent mixed failure mechanisms (which seems
to be the case), were the treatments indicated to be of any
value here, which they have not been, as this could lead to
approaches to eliminate the controlling defects.

Fig. 3 Comparison of the t of the log(fracture strength)


data to (top) Normal and (bottom) Weibull distributions
probability scales, using (i 0.5) to calculate the proportion,
where i is the value rank.

5.

Conclusions

The symmetric four-point bending test for adhesion testing


was found to be simple to implement and effective, providing good discriminatory power, lending itself to the problem
of adhesion testing in general.
It is possible improve appreciably the adhesion of resin
luting cements to zirconia by surface treatment, and that

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

simple gritblasting alone (GB) may be the current optimized


procedure.
Although the fusion glass-ceramic application (CC) proved
difcult to control, the concept could be viable if the formulation can be improved.
Zirconia primer (Z) and etched glaze (GHF) gave some
improvement in adhesion over the control (C), but a protocol
for controlling the thickness of the glaze layer is required.
Light and scanning electron microscopy might not provide
full information on fractured surfaces, but EDS analysis can
be recommended as a complementary method for evaluation of the chemistry of fracture surfaces in more detail.

Acknowledgements
This work was submitted in partial fulllment of the requirements for the degree of Ph D for SS. It is dedicated to the
memory of Professor Dr. Cenk Cura (19652011), under whose
tutelage this work was commenced. We gratefully acknowledge the assistance of the Izmir Institute of Technology, Izmir,
Turkey, in the ESEM-EDS analysis.

references

[1] Kelly JR, Denry I. Stabilized zirconia as a structural ceramic:


an overview. Dent Mater 2008;24:28998.
[2] Denry I, Kelly JR. State of the art of zirconia for dental
applications. Dent Mater 2008;24:299307.
[3] Thompson JY, Stoner BR, Piascik JR, Smith R.
Adhesion/cementation to zirconia and other non-silicate
ceramics: where are we now? Dent Mater 2011;27:7182.
[4] Vagkopoulou T, Koutayas SO, Koidis P, Strub JR. Zirconia in
dentistry: Part 1. Discovering the nature of an upcoming
bioceramic. Eur J Esthet Dent 2009;4:13051.
[5] Cura C, zcan M, Isik G, Saracoglu A. Comparison of
alternative adhesive cementation concepts for zirconia
ceramic: glaze layer vs. zirconia primer. J Adhes Dent
2012;14:7582.
[6] Zhang S, Kocjan A, Lehmann F, Kosmac T, Kern M. Inuence
of contamination on resin bond strength to nano-structured
alumina-coated zirconia ceramic. Eur J Oral Sci
2010;118:396403.
[7] Foxton RM, Cavalcanti AN, Nakajima M, Sherriff M, Melo L,
Watson TF. Durability of resin cement bond to aluminium
oxide and zirconia ceramics after air abrasion and laser
treatment. J Prosthodont 2011;20:8492.
[8] zcan M, Allahbeickaraghi A, Dndar M. Possible hazardous
effects of hydrouoric acid and recommendations for
treatment approach: a review. Clin Oral Investig
2012;16:1523.
ME. Adhesion concepts in
[9] zcan M, Dndar M, Cmlekoglu
dentistry: tooth and material aspects. J Adhes Sci Technol
2012;26:2281661.
[10] Chen JH, Matsumura H, Atsuta M. Effect of etchant, etching
period, and silane priming on bond strength to porcelain of
composite resin. Oper Dent 1998;23:2507.
[11] Bailey LF, Bennet RJ. DICOR surface treatments for enhanced
bonding. J Dent Res 1988;67:92531.
[12] Tada K, Sato T, Yoshinari M. Inuence of surface treatment
on bond strength of veneering ceramics fused to zirconia.
Dent Mater 2012;31:28796.

667

[13] Kwon JE, Lee SH, Lim HN, Kim HS. Bonding characteristics
between zirconia core and veneering porcelain. Dent Mater
2009;25(5):e42.
[14] Gstemeyer G, Jendras M, Dittmer MP, Bach FW, Stiesch M,
Kohorst P. Inuence of cooling rate on zirconia/veneer
interfacial adhesion. Acta Biomater 2010;6:45328.
[15] Magne P, Paranhos MPG, Burnett LH. New zirconia primer
improves bond strength of resin-based cements. Dent Mater
2010;26:34552.
[16] Anusavice KJ, Kakar K, Ferree N. Which mechanical and
physical testing methods are relevant for predicting the
clinical performance of ceramic-based dental prostheses?
Clin Oral Implants Res 2007;18:21831.
[17] Scherrer SS, Cesar PF, Swain MV. Direct comparison of the
bond strength results of the different test methods: a critical
literature review. Dent Mater 2010;26(2):e7893.
[18] Ban S, Anusavice KJ. Inuence of test method on failure
stress of brittle dental materials. J Dent Res 1990;69:17919.
[19] Armstrong S, Geraldeli S, Maia R, Raposo LHA, Soares CJ,
Yamagawa J. Adhesion to tooth structure: a critical review of
micro bond strength test methods. Dent Mater
2010;26:e5062.
[20] Timoshenko S. Strength of materials. Part I. Elementary
theory and problems. New York: D. Van Nostrand; 1940. p.
137.
[21] Placido E, Meira JB, Lima RG, Muench A, de Souza RM,
Ballester RY. Shear versus micro-shear bond strength test: a
nite element stress analysis. Dent Mater 2007;23:108692.
[22] Darvell BW. Adhesion strength testingtime to fail or a
waste of time? J Adhes Sci Technol 2009;23:93544.
[23] Della Bona A, Donassollo TA, Demarco FF, Barrett AA,
Mecholsky JJ. Characterization and surface treatment effects
on topography of a glass-inltrated
alumina/zirconia-reinforced ceramic. Dent Mater
2007;23:76975.
[24] Mirmohammadi H, Aboushelib MN, Kleverlaan CJ, Jager N,
Feilzer AJ. The inuence of rotating fatigue on the bond
strength of zirconia-composite interfaces. Dent Mater
2010;26:62733.
[25] Granjon H. Fundamentals of welding metallurgy. Cambridge,
UK: Abington Publishing; 1991. p. 192206.
[26] Staninec M, Kim P, Marshall GW, Ritchie RO, Marshall SJ.
Fatigue of dentincomposite interfaces with four-point
bend. Dent Mater 2008;24:799803.
[27] Carter CB, Norton MG. Ceramic materials: science and
engineering. Springer; 2007. p. 2978.
[28] Jones DW, Jones PA, Wilson HJ. The relationship between
transverse strength and testing methods for dental
ceramics. J Dent 1972;1:8591.
[29] ASTM International. Designation: C1674-11. Standard test
method for exural strength of advanced ceramics with
engineered porosity (Honeycomb Cellular Channels) at
ambient temperatures. West Conshohocken, PA, USA: ASTM
International; 2011.
[30] Song HY, Yi YJ, Cho LR, Park DY. Effects of two preparation
designs and pontic distance on bending and fracture
strength of ber-reinforced composite inlay xed partial
dentures. J Prosthet Dent 2003;90:34753.
[31] Van Noort R. An introduction to dental materials. 4th ed.
London: Elsevier Mosby; 2013. p. 317.
[32] Amaral R, zcan M, Valandro LF, Balducci I, Bottino MA.
Effect of conditioning methods on the microtensile bond
strength of phosphate monomer-based cement on zirconia
ceramic in dry and aged conditions. J Biomed Mater Res B
Appl Biomater 2008;85:19.
[33] Lowry R. Log-linear analysis for an A B C contingency
table; 2001. http://vassarstats.net/abc.html [accessed
01.03.14].

668

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 657668

[34] Valandro LF, zcan M, Amaral R, Vanderlei A, Bottino MA.


Effect of testing methods on the bond strength of resin to
zirconia-alumina ceramic: microtensile versus shear test.
Dent Mater 2008;27:84955.
[35] Jin J, Takahashi H, Iwasaki N. Effect of test method on
exural strength of recent dental ceramics. Dent Mater
2004;23:4906.
[36] Berenbaum R, Brodie I. Measurement of the tensile strength
of brittle materials. Brit J Appl Phys 1959;10:2817.
[37] Zhai T, Xu YG, Martin JW, Wilkinson AJ, Briggs GAD. A
self-aligning four-point bend testing rig and sample
geometry effect in four-point bend fatigue. Int J Fatigue
1999;21:88994.
[38] Kelly JR. Dental ceramics: what is this stuff anyway? J Am
Dent Assoc 2008;139:4S7S.
[39] De Souza GMD, Thompson VP, Braga RR. Effect of metal
primers on microtensile bond strength between zirconia
and resin cements. J Prosthet Dent 2011;105:296303.
[40] Osman SA, McCabe JF, Walls AW. Film thickness and
rheological properties of luting agents for crown
cementation. Eur J Prosthodont Restor Dent 2006;14:237.
[41] Casucci A, Osorio E, Osorio R, Monticelli F, Toledano M,
Mazzitelli C, et al. Inuence of different surface treatments
on surface zirconia frameworks. J Dent 2009;37:8917.
[42] Aboushelib MN, Feilzer AJ, Kleverlaan CJ. Bonding to zirconia
using a new surface treatment. J Prosthodont 2010;19:3406.
[43] Monaco C, Cardelli P, Scotti R, Valandro LF. Pilot evaluation
of four experimental conditioning treatments to improve
the bond strength between resin cement and Y-TZP ceramic.
J Prosthodont 2011;20:97100.
[44] Kern M, Barloi A, Yang B. Surface conditioning inuences
zirconia ceramic bonding. J Dent Res 2009;88(9):81722.
[45] Kern M, Wegner SM. Bonding to zirconia ceramic: adhesion
methods and their durability. Dent Mater 1998;14:6471.
[46] Amaral R, Ozcan M, Bottino MA. Microtensile bond strength
of a resin cement to glass inltrated zirconia-reinforced
ceramic: the effect of surface conditioning. Dent Mater
2006;22:28390.
[47] Hallmann L, Ulmer P, Reusser E, Hmmerle CHF. Effect of
blasting pressure, abrasive particle size and grade on phase
transformation and morphological change of dental zirconia
surface. Surf Coat Technol 2012;206:4293302.
[48] Sato H, Yamada K, Pezzotti G, Nawa M, Ban S. Mechanical
properties of dental zirconia ceramics changed with
gritblasting and heat treatment. Dent Mater 2008;27:40814.
[49] Tsuo Y, Yoshida K, Atsuta M. Effects of alumina-blasting and
adhesive primers on bonding between resin luting agent
and zirconia ceramics. Dent Mater 2006;25:66974.

[50] Qeblawi DM, Munoz


CA, Brewer JD, Monaco EA. The effect of
zirconia surface treatment on exural strength and shear

[51]

[52]

[53]

[54]

[55]

[56]

[57]

[58]

[59]

[60]

[61]

[62]
[63]

[64]

bond strength to a resin cement. J Prosthet Dent


2010;103:21020.
Re D, Augusti D, Augusti G, Giovannetti A. Early bond
strength to low-pressure sandblasted zirconia: evaluation of
a self-adhesive cement. Eur J Esthet Dent 2012;7:16475.
Zhang Y, Lawn BR, Malament K, Thompson VP, Rekow ED.
Damage accumulation and fatigue life of particle-abraded
ceramics. Int J Prosthodont 2006;19:4428.
Gomes AL, Castillo-Oyage R, Lynch CD, Montero J,
Albaladejo A. Inuence of sandblasting granulometry and
resin cement composition on microtensile bond strength to
zirconia ceramic for dental prosthetic frameworks. J Dent
2013;41:3141.
Derand T, Molin M, Kvam K. Bond strength of composite
luting cement to zirconia ceramic surfaces. Dent Mater
2005;21:115862.
Cattell MJ, Chadwick TC, Knowles JC, Clarke RL. The
development and testing of glaze materials for application
to the t surface of dental ceramic restorations. Dent Mater
2009;25:43141.
Aboushelib MN, Kleverlaan CJ, Feilzer AJ. Effect of zirconia
type on its bond strength with different veneer ceramics. J
Prosthodont 2008;17:4018.
Ntala P, Chen X, Niggli J, Cattell M. Development and testing
of multi-phase glazes for adhesive bonding to zirconia
substrates. J Dent 2010;38:77381.
Douglas HB, Moon PC, Eshleman JR, Lutins ND. The occlusal
dimensional change upon glazing porcelain. J Dent Res
1981;60:8289.
Edelhoff D, zcan M. To what extent does the longevity of
xed dental prostheses depend on the function of the
cement? Working Group 4 materials: cementation. Clin Oral
Implants Res 2007;18(3):193204.
Reich S, Wichmann M, Nkenke E, Proeschel P. Clinical t of
all-ceramic three-unit xed partial dentures generated with
three different CAD/CAM systems. Eur J Oral Sci
2005;113:1749.
Mously HA, Finkelman M, Zandparsa R, Hirayama H.
Marginal and internal adaptation of ceramic crown
restorations fabricated with CAD/CAM technology and the
heat-press technique. J Prosthet Dent 2014;112(2):
24956.
Shenoy A, Shenoy N. Dental ceramics: an update. J Conserv
Dent 2010;13(4):195203.
Guazzato M, Albakry M, Quach L, Swain MV. Inuence of
surface and heat treatments on the exural strength of a
glass-inltrated alumina/zirconia-reinforced dental
ceramic. Dent Mater 2005;21:45463.
Darvell BW. Development of strength in dental silver
amalgam. Dent Mater 2012;28:20717.

S-ar putea să vă placă și