Вы находитесь на странице: 1из 71

Preparation Guidelines for Full Zirconia Crowns

 Posted on May 21, 2018 by Backer Dental Lab


Zirconia restorations, since its debut in the dental industry, have become the better
choice for dentists who aim to provide their patients with the most technologically-
advanced metal-free dental restorations. It has proven to be a great improvement since
there is a wider array of available milling pucks that allow for greater shade
translucencies and variation that closely resemble natural teeth. In addition to improved
esthetics, the physical properties of Zirconia also allow for strength, durability, and
precision-fitting restorations.

If dental professionals want to maximise the success of Zirconia restorations as well as


to have minimal chair time, it is of utmost importance that the proper preparation and
guidelines are followed. Know more about the different preparations here:

PREPARATION GUIDELINES FOR ANTERIOR ZIRCONIA CROWNS


When the dentist is preparing a tooth for an interior Zirconia crown, there should be
enough room for the wall thickness – there should be a required minimum of 0.3mm and
at least 1.0mm and 1.5mm, or 1.8 to 2.0mm incisal reduction. There should also be a
visible and continuous circumferential chamfer with at least 0.5mm reduction at the
gingival margin. It is also important for the vertical and horizontal preparation of the
tooth should have an angle of approximately 5 degrees and a bevel is certainly not
advisable. All incisal edges should also be rounded.

PREPARATION GUIDELINES FOR POSTERIOR ZIRCONIA CROWNS


When preparing a tooth for posterior Zirconia crowns, on the other hand, you need to
make sure that there is enough room for the wall thickness to have a minimum of
0.5mm and at least between 1mm and 1.5mm or 1.5 to 2mm occlusal reduction. The
preparation should be tapered between an angle of 4 and 8 degrees. There is also a
need for visible and continuous circumferential chamfer and reduction-wise, there
should be at least 0.5mm at the gingival margin. Similar to that of anterior crown
preparation, a bevel is should also not be used. All occlusal edges should be rounded.

Factors That Make Crown Preparations Unacceptable for Zirconia Restorations


In order for a crown prep to be acceptable for a Zirconia crown restoration, the prep
should not have any undercuts nor a gutter preparation. A 90-degree shoulder is also
unacceptable as well as parallel wall preparations. Occlusal edges or sharp incisal are
also not suitable for Zirconia restorations.
 

Cementing and Finishing a Zirconia Restoration


Oftentimes, marginal finishing is required when fitting and cementing a Zirconia
restoration. Excess cement, if there’s any, must be removed to avoid plaque formation
that can lead to tooth sensitivity and periodontal disease. Although there are some
adjustments that might slightly roughen the surface of Zirconia restorations, it should
still be easy to polish to create an exceptionally smooth surface.

If you wish to discuss a dental implant case in more detail, know that our experienced
technical team is always ready to assist you. Reach out to us for more details!

Monolithic Vs. Layered Restorations: Considerations


for Achieving the Optimum Result
Newton Fahl, Jr., DDS, MS; Edward A. McLaren, DDS; and Robert Margeas, DDS

Q: What are your parameters when considering placing monolithic versus


layered restorations?
Dr. Fahl

The primary goal of either monolithic or layered restorations is to reintegrate form, function, and
esthetics with minimal damage and maximum longevity to the remaining natural dentition. Today’s
state-of-the-art technology available in both realms is capable of yielding from above average to
excellent esthetic results. The clinical choice between one or the other can depend on several
factors that include strength and esthetics and whether restoring the anterior or posterior segments.
The layering porcelain that is stacked over the core of all restorations is the weakest link that gives
“under shear” or flexural loads between 90 and 140 MPa. Because of high flexural strength (380-
1,000 MPa), monolithic restorations are ideally indicated for stress-bearing areas and can be used
as a single bulk material without the need for the weaker outer layer of stacked porcelain, especially
in the posterior zone or in the form of short-span anterior or posterior bridges.

Most powder/liquid porcelains will present color and optical properties that most closely match that of
natural dentin and enamel, which is an advantage over monolithic ceramics. The greatest challenge
with monolithic restorations has been optimizing esthetic outcomes. Newer blocks and ingots that
present improved color and optical properties minimize the use of surface stains. CAD/CAM
technology now allows the milling of blocks that have a dentin-like color bulk (eg, CEREC  Block
®

PC), with more pronounced hue and chroma, topped with an enamel-like, more translucent layer.
The milling can be adjusted to achieve the desired color result while final staining is still an option for
further customization. This option is convenient in the posterior segment where esthetic challenges
are not as great. Other CAM/CAM systems (eg, Lava  DVS) allow characterization to be applied

internally, rendering the restoration more polychromatic and natural looking. Other monolithic
systems now present a high-translucency coping material that precludes the use of a veneering
layer due to improved optical characteristics (eg, e.max  HT). This restorative option is particularly
®

important in cases of anterior veneers on bruxers or where occlusion poses a challenge.

There is a wide range of all-ceramic choices. The outer esthetic layer can be accomplished with
conventional powder and liquid porcelain or pressed over the ceramic coping. The latter seems to be
gaining greater acceptance due to the ease of fabrication and precision of the marginal fit. Layered
all-ceramic restorations comprise veneers, inlays/onlays/overlays, full crowns, and bridges. The
principal difference in layered all-ceramic restorations lies in the ceramic used for the coping, which
include zirconia, alumina, and lithium disilicate.

Dr. McLaren
The good news for clinicians is that more products are entering the marketplace. While monolithic
lithium disilicate (e.max ) and full-contour zirconia (BruxZir ) have been the dominant materials and
® ™

have performed excellently, manufacturers are increasingly developing competitive materials, which
is paving the way for more innovation. In the zirconia area, translucency is being improved. Zirconia
is a wonderful material to work with, however translucency can be an issue. My recommendation
today would be to use full-contour zirconia mostly on the posterior teeth. In clinical situations where
cement is needed, conventional cements can be used such as phosphate or glass ionomer. This is
because the strength of the restoration is not enhanced by the bonding procedure.
One of the premier benefits of lithium disilicate is that it is easy to bond. It is an etchable ceramic. I
differentiate those two when I have a partial preparation or a conservative preparation when doing
essentially non-retentive preparations. This is clearly an indication for e.max or other similar
products. For a more “normal” crown situation, involving a retentive-type preparation or subgingival
margins that can’t be bonded, or if not in the anterior, full-contour zirconia would be recommended—
ie, one of the more translucent versions and one on which some surface color could be added. In the
anterior, for monolithic restorations a high-translucency material is needed for esthetics purposes,
which would be lithium disilicate or e.max. A new version coming onto the market from VITA, which
is already on the market in Europe, is called Suprinity , a zirconium-reinforced glass ceramic. Also,
®

DENTSPLY has introduced new Celtra  and a machinable version, Celtra  Duo. Like e.max,
™ ™

Suprinity and Celtra require machine crystallization. The Duo version does not, however it offers
about half the strength, making it well suited in situations where time is an issue and excessive
strength is not needed. In the author’s experience testing these newer materials, translucency is not
an issue. They can be made opaque or translucent based on the version of the material—therein lies
the advance.

Relative to anterior teeth and monolithic, my preference continues to be to layer, whether using
e.max, veneer, or zirconia coping. If there is space to layer and if strength is not an issue, 3-
dimensional color can be created. Monolithic systems can provide good esthetics if the correct
translucency is chosen and surface color is applied effectively.

Dr. Margeas

Monolithic restorations are being promoted by dental laboratories heavily for their strength and
reasonable cost, with the most popular being lithium disilicate and zirconia. They are both good
restorations, but they should be used in different areas of the mouth for maximum strength and
esthetics. In my practice, I use full-zirconia restorations sparingly in the posterior region of the mouth
for full-coverage molar crowns, mostly when the patient does not want a gold restoration and there is
not enough room for a porcelain-fused-to-metal crown. The esthetics can range from bad to good
depending on the laboratory that fabricates the restorations. These restorations can be
conventionally cemented and are very cost effective. However, they are surface stained, so if the
occlusion needs adjustment, the restorations may need to be reglazed. The other drawback is if the
crown needs to be removed or endodontics performed, it is very hard to drill through. As
manufacturers continue to create more translucent zirconia, technicians will be able to create
anterior restorations that are monolithic with minimal to no addition of layered porcelain.
Lithium-disilicate monolithic restorations are more esthetic, but the material is not as strong. They
have about 400 MPa of strength and are used in the molar region and some second bicuspids. The
material is more translucent than zirconium, but the final esthetics still depends on the laboratory
that fabricates them. They also are cost effective because they are easily fabricated, and are popular
with many laboratories. For maximum esthetics the lithium-disilicate material should be cut back and
a layer of porcelain fired over the core. This would be necessary in the anterior region.

ZIRCONIA CROWNS PREPARATION GUIDELINES


Posted by Bill Warner | July 02, 2019| 
Dental Technology, Tips & Tricks, Zirconia

Since its emergence into the dental arena, Zirconia has increasingly


become the material of choice for clinicians who wish to provide their
patients with the most technologically advanced metal-free
restorations. Zirconia has improved significantly with the introduction of a
wider array of available milling pucks which allow for greater shade
variation and translucencies that closely resemble natural dentition. In
addition to improved esthetics, the Zirconia physical properties allow for
durability, strength, and precision-fitting restorations. To maximize on the
success of seating Zirconia restorations, and minimizing chair time, it is
essential to ensure that proper preparation guidelines are being followed.

Download Crowns & Bridges Comparison Chart »

Why Tooth Preparation Design Is Important 


 Smooth edges result in lower stress on the crown. This lower stress
decreases the percentage of fractures occurring. 
 Ceramic restorations require a passive fit. 
 Uniform reduction results in ideal ceramic strength.  
 Sufficient reduction leads to the best esthetic results. 
 Digital scanners read smoother preparations with more accuracy. 

Preparation Guidelines for an Anterior Zirconia Crown


When prepping a tooth for an anterior Zirconia crown, you will need to
ensure that there is sufficient room for the wall thickness to have a
minimum of 0.3 mm and ideally between 1.0 mm and 1.5 mm, or 1.8 to
2.0 mm incisal reduction. There should be a clearly visible and continuous
circumferential chamfer with a reduction of at least 0.5 mm at the gingival
margin. The horizontal and vertical preparation of the tooth should have
an angle of approximately 5° and a bevel is not advisable. All the incisal
edges should be rounded and you want to reduce the linguals of anteriors
with football diamond to create a concave lingual. 

Preparation Guidelines for a Posterior Zirconia Crown


When prepping a tooth for a posterior Zirconia crown, you will need to
ensure that there is sufficient room for the wall thickness to have a
minimum of 0.5 mm and ideally between 1 mm and 1.5 mm or 1.5 to 2
mm occlusal reduction. The prep should be tapered between 4°and 8°. It
will also need to have a clearly visible and continuous circumferential
chamfer and a reduction of at least 0.5 mm is required at the gingival
margin. Just as with the preparation for an anterior crown, a bevel is not
recommended. Ensure that all occlusal edges should be rounded.

Although the Shoulder and Chamfer preparations are the most ideal,
Feather edge preparations are typically not recommended, but can be
acceptable for full-Zirconia crowns.  Check with your dental laboratory to
see if their fabrication process will allow for this form of prep, as different
types of Zirconia require different guidelines.

Factors That May Make a Crown Preparation


Unacceptable for a Zirconia Restoration
To be acceptable for a Zirconia crown restoration, the preparation should
not have any undercuts and it should not have a gutter preparation. Both
90° shoulder and parallel wall preparation are unacceptable. Sharp incisal
or occlusal edges are not suitable for a zirconia restoration.
 

Layered or Monolithic Restorations


For restorations within the esthetic zone, a Zirconia crown layered on the
facial, or full layering, will provide optimal results. Over the past few years
layering porcelain techniques have improved significantly so a Zirconia
substructure that has been layered with porcelain is unlikely to chip or
fracture on the occlusal or incisal. However, if maximum strength is
required because a patient has bruxism, a heavy bite or where there is
only limited occlusal clearance, a monolithic crown may be a better
posterior solution. Thanks to newer monolithic zirconia (single, solid
blocks of zirconia), this type of crown is nearly unbreakable and built to
withstand the harsh environment the back of the mouth endures better
than all-porcelain crowns.  Monolithic restorations are providing an
increasingly aesthetic result with the introduction of High Translucency
Zirconia or the DDS Lab product reference, Zirconia HT.

For a full-contour monolithic Zirconia crown, there should be a 1.0 to 1.5


mm occlusal depth cut to achieve appropriate occlusal anatomy. You will
need to ensure a 1.0 to 1.5 mm functional cusp tip reduction, a 0.5 mm
gingival chamfer reduction, a 6 to 8 degree taper to the axial walls, and a
1.0 mm occlusal 1/3 reduction of the functional cusp. For the marginal
design, prepare a 0.3 to 0.5 mm chamger to allow for a more accurate mill
of the pre-sintered zirconia. 

Cementing and Finishing Zirconia Restorations


Quite often marginal finishing is required when fitting and cementing
Zirconia restorations. Excess cement must be removed to avoid plaque
formation which can lead to tooth sensitivity and periodontal disease.
Gingival margins can be finished using non-cutting, safe-end finishing burs
that will protect the soft tissues. Although these adjustments may slightly
roughen the surface of a Zirconia restoration, it should be easy to polish,
creating an exceptionally smooth surface.
If adjustments are required it is important to use a fine diamond bur
suitable for a Zirconia restoration. Too much heat or sparking from an
aggressive reduction will lead to micro-fractures in the Zirconia. It is
important to use as little pressure as possible to reduce the amount of
heat that is produced. Using water while adjusting is also a helpful way to
keep the restoration cool.

Please be reminded that should you wish to discuss a


Zirconia implant case in more detail, our experienced technical team is
here to assist you.

Click here to schedule a consultation with our technical team  » 

The Zirconia Ceramic: Strengths and Weaknesses


Elie E. Daou*
Author information Article notes Copyright and License information Disclaimer
This article has been cited by other articles in PMC.

Abstract
Go to:

INTRODUCTION
When restoring a tooth, the clinician faces the dilemma: Which material should he use? [1]. The
major factors that may influence the final choice are esthetics and strength of the prostheses.
Metal ceramic fixed partial dentures (FPDs) are considered the gold standard, as reliable
materials. However, the request for esthetic dentistry as well as the rising question regarding
biocompatibility of dental alloys, support the commercialization of new products [2]. Nowadays,
all-ceramic prostheses are replacing, more and more, metal-based restorations [3]. A variety of
ceramic systems are developed for single crowns or fixed dental prostheses (FDPs) with an
excellent esthetic outcome [4].
Traditional ceramics (glass- glass-reinforced, and feldspathic ceramics) and also Al2O3-
reinforced ceramics encountered some problems, especially in the molar region [4]. Ceramic
material used, core-veneered bond-strength, and crown thickness are some factors essential to
withstand occlusal forces [5]. The reliability of industrially prefabricated ceramics block appears
to be more consistent than laboratory manually processed ceramics [6, 7].
Transformation-toughened zirconia is prone to be a successful alternative in the different clinical
situations compared to other all-ceramic systems [8]. Their mechanical and optical properties
allowed them to be used as a framework material. In vitro studies demonstrated a flexural
strength of 900–1200MPa [9, 10], and a fracture toughness of 9–10MP am 1/2 [2, 11]. The
restorations are processed either by soft machining of presintered blanks followed by sintering at
high temperature, or by hard machining of fully sintered blanks [12, 13].
This review article describes the current status of zirconia-based fixed restorations, including
results of current in vitro studies and the clinical performance of these restorations [14].
Since its development in 2001 [12], direct ceramic soft- machining of pre-sintered 3Y-TZP is
now on the market. First, the die or the wax pattern is scanned, the computer software (CAD)
designs an enlarged restoration and computer aided machining mill a pre-sintered ceramic blank.
The restoration is then sintered at high temperature [13].
Hard-machining Y-TZP blanks consist of milling restorations in very high density blocks,
presintered at 99% of the theoretical density [13]. The milling system has to be particularly
robust due to the high hardness and low machinability of fully sintered Y-TZP [13].
Same 5-years survival rates have been reported for all-ceramic crowns and metal-ceramic for
anterior teeth. When utilized for premolars and molars the success decreased to 90.4% and
84.4% respectively for In: ceram crowns and glass–ceramic crowns [15-17]. The force, contact
area and duration were greater for molar chewing cycles than incisor ones [18]. The most
common complication encountered with all-ceramic crowns was crown fracture [19].
Up to 3-years published controlled clinical studies of zirconia-based crowns reported lower
complication rates [20, 21]. The authors concluded that Y-TZP could sufficiently withstand
functional load in the posterior zone [22]. However, as mentioned by Conrad and others,
following traditional preparation guidelines will better distribute stress during dynamic loading
of the restoration [5, 23].
An electronic search has been conducted, during July 2013, via PubMed and Elsevier. Peer-
reviewed articles were targeted. The following key-words have been used: Zirconia, zirconia
restorations, allceram, zirconia crowns, zirconia FPDs, zirconia bonding, and zirconia strength.
Available full-text articles were read. Related articles were also scrutinized. No hand search was
driven.
Go to:

FRACTURE RESISTANCE
The mechanical properties of zirconia allowed them to be used in posterior FPDs and permit
substantial reduction in core thickness [13].
Under ambient pressure, the temperature will influence the crystallographic form of unalloyed
zirconia. At room temperature and upon heating up to 1170 °C, the structure is monoclinic. Then
it is tetragonal between 1170 and 2370 °C and cubic above 2370 °C and up to the melting point
[24]. Upon cooling, transformation from the tetragonal (t) phase to the monoclinic (m) phase will
induce a substantial increase in volume (~4.5%). This will lead to catastrophic failure. Adding
CaO, MgO, Y2O3 or CeO2 to zirconia-alloys allows the retention of the tetragonal structure at
room temperature. This will control the stress-inducing t→m transformation. Compressive
stresses developed in the vicinity of a crack tip, arrest crack propagation and lead to high
toughness [13, 25, 26].
Composition, grain size, shape of the zirconia particles, type and amount of the stabilizing
oxides, interaction of zirconia with other phases and processing are also factors that have impact
on the metastability of the transformation [26].
However, grinding or sandblasting are responsible to cause the t→m transformation altering the
phase integrity of the material and increasing the susceptibility to aging [27, 28]. The presence of
water will exacerbate this well-documented "Low Temperature Degradation" (LTD) [29, 30].
The Y2O3 can react with the aqueous environment producing yttrium hydroxide (Y[OH]3H2O)
[31, 27]. Grain pullout and microcracking as well as decrease in strength are reported
consequences of this aging process [13, 32]. Frameworks or parts of a framework that are not
veneered as well as zirconia implants and abutments exposed to the oral environment, are
subjected to this phenomenon. This is why during framework design, non-veneered zirconia
should be avoided [27].
Innovative bioceramics such as zirconia magnesia (Mg-PSZ with bioactive glass coating) [33]
and alumina composites TZP stabilized [34] are recently reported as degradation-free materials
[35].
Both in vitro and in vivo studies demonstrated that fracture of the connectors was the exclusive
mode of failure in all-ceramic FPDs [36, 37]. Connector fracture was initiated at the gingival
embrasure. Concentration of tensile stresses can be reduced by larger radius of curvature at the
gingival embrasure [38]. Whereas, sharp occlusal embrasures did not affect FPDs' fracture
resistance [39, 40].
An occlusogingival height of 2.5 mm and a buccolingual width of 2.5 mm of the connectors (a
connector surface area of 6.25 mm2) are sufficient to ensure long-term success of metal-ceramic
FPDs [41]. These dimensions are achievable both in the anterior and in the posterior segments.
Mechanical strength of zirconia frameworks is up to three-times higher than other allceram. It
can withstand physiological occlusal forces applied in the posterior region [4, 14, 42, 43] Even
rare, framework fractures in all-ceram FPDs were reported in the connector region [4, 44-46].
Therefore, connector dimensions are crucial for fracture resistance [40].
Fracture propagated from the gingival surface of the connector toward the pontic [47]. A
connector dimension of 3×3 mm increased the fracture strength of zirconia-based FPDs by 20%
[44, 48, 49]. Required dimensions for the connector could still be smaller than for other all-
ceramic core materials [40]. Even though, some authors recommended a connector dimension of
4×4 mm and that the framework must support the veneering porcelain, which should not include
more than 2.0 mm of unsupported veneering material [14, 27, 50-52]. Worth noting that bulk
fracture appears to be quite uncommon [13].
The major problem encountered is porcelain cracking. The difficulties are material-specific with
an incidence from 8 to 50% [53, 54]. Thickness ratios or framework design also play a role. For
comparison, porcelain problems on metal–ceramic prosthesis over a 10 years observation period
was reported to be, no higher than 6% for most alternative alloys, [55]. 98% completely intact
porcelain at 5 years was reported for goldbased alloy [56, 69]. Thus, porcelain–zirconia
compatibility is to be considered [13].
Zirconia–porcelain interface may be involved in crazing and chipping during function. Stresses
could be related to surface property, as bulk thermal expansion/contraction mismatches does not
appear to be the cause [13]. The aggressivity of silicate glasses as solvents of refractory materials
at high temperature is known [57]. Under firing conditions, aluminum oxide is soluble in dental
porcelains [58]. Cerium and zirconium diffuse into a glass used to infiltrate a partially sintered
Ce-TZP powder [59]. Lessening of stabilizing dopants (e.g., Y and Ce) might induce local
changes in zirconia surface [60] resulting in destabilization of the t-phase [61] with quite high
local associated strains [62]. Liquid silicate can penetrate the grain boundaries perhaps analogous
to water penetration of Y-TZP [13, 63].
Go to:

CHIPPING AND FAILURES


Chipping is defined as "a typical failure of contact loadings, normally produced when a crack
generated or propagated by contact loads deflects due to the presence of a free surface nearby"
[64, 65]. Tensile stress induces fracture of the brittle ceramic usually perpendicular to the applied
force [66].
Thermal coefficient mismatches, processing (porosity, impurity inclusion) and inherent material
defects (large grains, residual scratches) will increase the probability of crack propagation under
loading [67]. Brittle fracture of ceramics will be triggered adjacent to these zones [66].
In the case of metal-ceramic prostheses, an adherent layer of oxide is essential to achieve a
strong bond. This will enhance the wettability and the adherence of the ceramic. When the
temperature attains a certain level, part of this oxide will be dissolved into the glass. In the case
of nickel-chromium alloys, excess oxide production will cause weak bonding [68]. High-gold
alloys will produce an adequate oxide layer for a solid bond with the porcelain [68].
The zirconia core–veneer bond should also be strong enough to gain profit from exceptional
properties of the framework. However, according to Aboushelib, this bond strength is lower than
for other all-ceram systems [69]. This can induce chipping and delamination under friction.
Framework surface treatment, the surface finish, the type and method of application of the
veneer ceramic will affect this bonding [70].
If no fractures of the zirconia framework have been reported [2], chip-off fracture rates up to
20% was observed at 5-years follow-up period [4, 5]. In the case of metal framework FPDs, a
review of the literature revealed either no fracture of the veneering ceramic [71] or substantially
lower fracture rates ranging from 2.7% up to 5.5% for observation periods from 10 to 15 years
[2, 72, 73].
Adequate framework design, proper veneering ceramic support and thickness are factors
implicated in the ceramic survival [74]. Moreover, occlusal forces, such as direction, magnitude
and frequency are to be taken into consideration [2, 75]. Roughness of the veneer that results
from occlusal contacts or grinding may induct chipping. Fractography analysis revealed that
crack propagation originated from wear area and occlusal adjustments [27, 76, 77]. Studies have
shown that sandblasting and sharp indentations even at very low loads are very harmful to long-
term longevity of zirconia [78-80].
Marchack et al.  demonstrated that scanning of full contour waxing will provide an optimal
porcelain thickness on appropriate coping design [81]. This will decrease porcelain fracture
[14, 82]. It has been recommended that the veneer thickness does not exceed two-fold of the core
thickness. The pontic framework must have an anatomical shape to support veneers' cusps [83].
However, a completely suitable veneer system has not yet been found. Differences in
microtensile bond strength between several veneering porcelains have been shown [84]. Strong
veneering systems are recommended to avoid chipping [4].
For others, the thermal expansion coefficient (TEC) plays a major role far before the zirconia-
veneer bond strength [85, 86, 77]. Most manufacturers provide veneering porcelains having a
slight mismatch between their porcelain and zirconia, with the porcelain having approximately
TEC lower than the zirconia TEC [84]. Desirable residual compressive stress in the veneering
ceramic is present when a framework material with a slightly higher TEC is used [87]. In
contrast, when Zirconia's TEC is lower than ceramic's, veneer delamination and microcracks
occurred [69, 88]. This approach is used for most metal-ceramic systems and non-zirconia all-
ceramic systems [89, 90, 13]. Therefore, if a compatibility issue is occurring with Y-TZP, it is
likely not only due to a simple thermal expansion coefficient mismatch between the bulk
materials [13]. Low fusing veneering ceramics with similar TEC have been developed. Grain
size may also play a role [2]. The wide range of sintering temperature has an incidence on the
particle size and later on the phase stability of zirconia-yittria [13].
Some recent studies described a layering method of indirect composite onto a zirconia
framework [61-66]. A short term in-vitro study reported a superior bond strength by using a
priming agent containing the functional monomer MDP [61]. Plastic and viscoelastic effects, as
well as susceptibility to creep And recovery [67, 68] are advantageous properties of using
composite, especially in areas of high occlusal stress [69, 87].
Zirconia presents a thermoconductivity much lower than that of other framework materials
[6, 9]. This low thermal conductivity retards the ceramic cooling rate at the interface. This
generates thermal residual stress [91, 92]. It may induce thermal cycling delamination of the
veneering porcelain [2]. The effect of different cooling rates (rapid and slow) on the bond
strength between layering porcelain and zirconia ceramics has been assessed [93, 94]. Prolonged
cooling phases have been proposed to reduce this stress and veneer chipping [83, 95, 96]. Slow
cooling time ameliorated the resistance of the veneered zirconia restorations [76, 97], and
enhanced the shear bond strength [93]. However, Gostemeyer et al. argued that adding 5 min
cooling in the furnace lowered the bond strength [94]. Komine noticed that these conflicting
findings are the result of different cooling and testing methods [87].
Applying a liner material has been proposed to mask the opacity of the zirconia-core.
Unfortunately, this reduced the core-veneer bond-strength and increased the percentage of
interfacial failure [70, 84]. Kim and Fischer agreed on the negative effect of liner application
[86, 98, 99]. Aboushelid contraindicated their use with Press-on ceramics [84]. The lower
strength of liners compared to dentine ceramic may play a role in these negative results. Still,
others found that liner materials enhances the bond strength between zirconia and some layering
ceramic [87].
Go to:

LAYERED /PRESSED VENEERING CERAMICS


Cohesive and adhesive failures of the veneering are recurrent complications of veneered zirconia
frameworks [76].
To counteract this tendency, the "overpressing technique" has been proposed. A specific ceramic
is pressed onto the zirconia framework [100]. According to Beuer et al. [101] this technique is
reliable since no chipping has been detected [14].
Fabrication of conventional dental porcelains consists of a frit condensation followed by a
sintering process. Sintering may introduce thermally induced residual stresses [102]. This can
modify the measured biaxial flexure strength [103-105]. The moisture content of the veneering
material during sintering might induce changes in the zirconia/veneering interface and provoke
transformation from the tetragonal phase to the monoclinic phase [106]. Swain et al. [95]
preconized that residual stresses and contact-induced cracking will develop chip-fracture.
Beuer et al. [100] reported higher strength of CAD/CAM veneering ceramic compared to the
layered veneering technique. Using of the pressed ceramic may reduce the chipping incidence
[4], since the heatpressing fabrication method would reduce the formation of large flaws and
minimize the thermally induces residual stresses [102, 105]. The manufactured blanks are
reported flawless. Greater porosities are entrapped during fabrication stages in dental laboratory,
added to human errors [107, 108]. The shrinkage level of the porcelain may be related to the
ratio of the mixed powder/liquid veneering ceramic. Minimal three firing cycles are needed.
Catastrophic failures can be induced by the incorporation of small impurities like
inhomogenities, pores, since cracks cannot be healed but slow growth may occur under oral
conditions [108, 109]. The pressing technique permit the creation of desired tooth anatomy while
minimizing firing shrinkage [110].
The manufactured CAD/CAM veneer will be joined to the zirconia framework by fusion glass
ceramic or by using resin cement [111]. Lithium disilicate has been proposed to be connected to
zirconia framework by glass fusion ceramic [76, 112]. Higher tensile strength of press-on
veneers and the superior quality of the interface can prevent porcelain chipping [84]. These
materials exhibited better fracture strength and fatigue behavior when compared to the hand-
layered ceramics. The latter show early veneer failures under mouth-motion cyclic loading
[76, 113].
In one recent study, overpressed zirconia three-unit posterior prostheses had significantly less
fractures and chippings compared with layered ones [114]. In another study no chipping was
observed [101, 108]. Ishibe and Aboushelid recommended the application of press-on veneer
ceramics directly onto air-borne-particle-abraded surfaces [70, 84, 115, 116]. However, other
studies found no difference in fracture incidence between the pressed and layered techniques
[100, 108, 117].
Connector design had an incidence on milled ceramic fracture resistance, but not the pressed
ceramic [118].
Go to:

SHEAR BOND STRENGTH AND INFLUENCE OF ARTIFICIAL AGEING


In metal-ceramic prostheses, as determined by the International Standards Organization (ISO)
[119], the minimum required bond strength between metal and layering porcelain is 25 MPa. No
such estimate has been yet determined for all-ceram [87]. Comparison between zirconia- and
metal-ceramic restorations found a similar bond strength [88, 115, 120, 121]. Other studies
reported greater bond strength between porcelain and zirconia than for metal-zirconia [2, 122].
Results are conflicting [87].
Guess found metal-ceramic shear bond strength (SBS) higher than zirconia-ceramic one's.
Thermocycling has no effect on zirconia-ceramic bond [2]. Yet, Silva noticed that in contrast to
Y-TZP systems where failures were accelerated by fatigue, metal-ceramic restorations failures
occurred as a function of load and not fatigue [123].
The difference in the results between metal- and zirconia-ceramic SBS could be reported to the
dissimilar adhesion mechanisms. If mechanical interlocking and chemical bond resulting from
suitable metal oxidation and interdiffusion of ions are essential in the metal ceramic interface,
the Y-TZP-ceramic bonding mechanisms are still unclear [124, 125]. For the latter, some
micromechanical interactions can just be assumed based on the wettability of zirconia-core by
veneering ceramic [2].
When Ishibe and Aboushelib compared zirconia-layered ceramic shear bond strength to zirconia-
pressed one, they found equivalent results [70, 84, 115].
Oral fluids are known to facilitate stress corrosion of ceramic materials. Water molecule will
diffuse into the glass and provoke a corrosion mechanism [126]. Dissolution of ceramic can
happen through 2 ways: by ionic exchange during exposure to an acidic solution, or by
breakdown of Si-O network in a basic solution [127]. The intensity of chemical deterioration is
related to the glass matrix composition and the ratio of crystal incorporation [66]. This will result
in slow crack growth and may lead to failure of ceramic restorations in the oral cavity complex
situation [2, 128, 129]. So some concerns are assessed regarding zirconium dioxide structural
stability when it's exposed to oral environment [6, 27] .
Different zirconia systems were tested in vitro, using artificial aging, through dynamic loading
and thermal cycling. No significant effects on the fracture load were observed for 3-unit FDPs,
and no failures occurred [4, 130, 131]. Thermocycling had no incidence on the zirconia-layered-
ceramic bond [2, 88]. The bond strength stability is equivalent to results found about bonding of
porcelain-metal framework [87, 88, 132].
Schmitter noticed that artificial aging has no incidence on CAD/CAM ceramics, in contrast with
the manually veneered crowns [76]. Another study found no difference between the two
veneering methods after aging [108]. No difference in the fatigue properties of the Zirconia
Everest® core material following sintering or heat pressing of the veneering material was
detected [133].
Analyses of the fracture surfaces for the pressed ceramic, revealed a combined adhesive and
cohesive failure scheme independent of the ageing [76]. Even on polished zirconia, the failure
was mostly cohesive within the veneering ceramic [69, 86]. The flexure strength varied between
70 and 100 MPa, depending of the product [108, 134]. The flexural strength of the zirconia
veneering porcelain, similar to metal-ceramics, will block the propagation of the crack due to the
tetragonal phase [69, 108].
Stawarczyk concluded that overpressed veneering porcelains for zirconia single crown
frameworks exhibited similar or better fracture load compared with layered ones [108]. Guess
noticed that hand-layer-veneered zirconia crowns revealed a high susceptibility to mouth-motion
cyclic loading in early veneer failures [113]. Other factors such as grain size and shape and
porosity should also be considered [26]. Grain size strongly affect the mechanical properties of
3Y-TZP [14, 20, 21]. On the other hand, sintering temperatures will influence the grain size and
the phase stability of 3Y-TZP [13].
Soft machining restorations are sintered at a later stage. This will prevent the stress-induced
transformation from tetragonal to monoclinic. The final surface will virtually be free of
monoclinic phase unless grinding adjustments are needed or sandblasting is performed [13]. In
contrast, hardmachined restorations of fully sintered 3Y-TZP blocks contain a significant amount
of monoclinic zirconia [26]. This may result in surface microcracking, higher susceptibility to
LTD and lower reliability [27].
Several searches studied the fatigue behavior of 3Y-TZP [28-31]. When tested in cyclic loading,
both sandblasting and sharp indentations even at very low loads are harmful to the long-term
performance of 3Y-TZP [13, 29-31]. The presence of residual stresses was detrimental in
promoting LTD [13].
Worth to be noticed that the pressable ceramic materials showed significantly less change in
marginal opening than metal ceramic and copy milled ceramic crowns [16].
Go to:

COLOR AND ESTHETICS


Tooth Enamel, constituted of 97% hydroxyapatite mineral, is very translucent and can transmit
up to 70% of light. Dentin is also able to transmit up to 30% of light. The esthetic dilemma for
metal-ceramic restorations is that opaque porcelain is needed to mask a metal substrate. It will
reflect light and decrease translucency. Consequently, they will often appear brighter intraorally
[5, 135]. In-Ceram Spinnell had higher levels of translucency than In-Ceram Alumina (VITA
Zahnfabrik), followed by In-Ceram Zirconia (VITA Zahnfabrik), which was comparable to a
metal alloy [5].
Zirconia framework is esthetically better accepted than metallic framework, but it remains
clinically too white and opaque. Therefore, manufacturers introduce colored zirconia framework
to ameliorate the overall matching color [136]. Different techniques have been proposed: adding
pigments to the initial zirconia ceramic powder, dipping zirconia milled frameworks in dissolved
coloring agents, applying liner material to sintered framework [69, 137]. Thinner veneer is then
required to mask the underlying framework [138].
The ability to control the shade of the core may also eliminate the need to veneer the lingual and
gingival aspects of the connectors in those situations where the interocclusal-distance is limited
and the required connector dimensions are minimally achieved. In addition, the palatal aspect of
anterior crowns and FPDs may be fabricated of the core material exclusively in situations of
extensive vertical overlap and lack of space for lingual veneering porcelain [40, 139].
Individualized colored overpressed ceramics have been also proposed as a quick and easy
technique [108]. Excellent esthetic and perfect matching are difficult to attain, as appearance will
rely on precolored ingots. To enhance esthetic, a layering ceramic can also be applied over an
pressed-on veneer [140].
Lava system (3M ESPE Dental Products), which is relatively translucent but may still mask
colored abutment, is proposed in 7 shades, permitting shading from the intaglio surface to the
outer [139].
The increase of the concentration of the coloring pigments at grain boundaries could be at the
expense of the stabilizing elements. This may result in higher percentage of tetragonal-
monoclinic transformation. If this transformation occurs on the surface of the framework, this
will provoke grain pullout and surface lifts [70]. This phenomenon results from competitive
displacement of the stabilizing elements by the metallic pigments in the liquid state. The latter
have a melting point lower than the yetrium oxide one [70]. A minor alteration of the location or
the concentration of the stabilizing elements can alter the mechanical properties of the zirconia
framework [141]. A fatigue process started on individual surface areas will lead to monoclinic
spots and results in surface microcracking and lifts. The color pigments at grain boundaries
replacing the reduction of yttrium will affect the slowly extend of this process toward the bulk of
the material [142, 143].
One study showed that the bond strength of colored zirconia is significantly lower compared to
non-colored zirconia [70]. When the framework is colored by dipping in pigment solution, the
pigments will concentrate on the outer surface. Those surface pigments tend to crystallize on the
surface and weaken the bond with the veneer ceramic [70].
Go to:

DISCUSSION
Variable study conditions and plethora of materials available made the comparison of the results
from relevant literature, a challenging issue [5]. Usually, a failure in any clinical trial results
from a combination of causes or events [1]. Worth noting, there is a remarkable emphasis on
clinical examination of the zirconia product [13], even though some of these studies lack
scientific support [144]. Reproducing intra-oral conditions, during the in vitro studies, is quite
difficult. An effort was made to create artificial oral environments by applying cyclic forces in
artificial saliva, under fluctuating temperature [145]. Long-term clinical studies are still needed
to make conclusions [5]. In the era of evidence-based dentistry, reinforcing standardization of
clinical cohort studies will permit more efficient conclusions [4]. It has been noted that some
granted research centers may be reluctant to publish bad results [146].
Concerns regarding metal-ceramic restorations biocompatibility limitations and optical qualities
provoked the shift to all-ceramic restorations placement. While achieving marginal accuracies
equal to that of metal-ceramic crowns, All-ceramic crowns provide superior gingival response
[147].
Glass ceramic crowns, even those with a densely sintered alumina core, showed brittle fracture
in the posterior region [148]. Patient selection may be critical and the technique remains
sensitive [149]. Poor oral hygiene, high caries incidence, moderate gingival inflammation and
severe parafunction are some of the exclusion criteria cited [150]. A coping design allowing
optimal ceramic layering thickness, a uniform cement film, and an adequate TEC matching
between the laminate and the core may reduce the stresses [148].
Studies reported that zirconia ceramic flexure strength and fracture toughness are twice that of
alumina ceramics [151]. The partially stabilized tetragonal modification of zirconia to a
monoclinic phase, induce by a tensile stress, exhibits 4% volume expansion. To propagate, the
crack must overcome the compressive stresses generated at the crack tip [152, 153].
The aim of this review was not to evaluate survival and failure of different restorations. Authors
agreed that Y-TZP can withstand physiologic functional loading forces and are comparable to
metal-ceramic FPDs [27, 154]. Strength and marginal fit of zirconia ceramic has been confirmed
by extensive laboratory testing [155, 156]. Still 5 to 10 year clinical studies are needed to
determine primary mode of failure and success rate [157].
The major complication reported is chipping of the veneer with a rate that will increase from 6%
to 10% between 3 and 5 years, whereas these values are reported on a 10 years observation
period for metal-ceramic restorations [27, 55]. Fracture of the zirconia framework is not probable
[27]. Long-term success is essentially dependant of the performance of the veneering [74]. Minor
chip-off of the layering ceramic has been reported as the most frequent complication [2]. Short-
span posterior frameworks are reliable, whereas data is lacking for long span and cantilevers [4].
If bond failure has been pointed as chipping reason [158], differences in thermal coefficients
[159], liner material and poor core wetting [84], veneer firing shrinkage [85, 86], phase
transformation [160], loading stresses, flaw formation [161], coloring pigments [70] and surface
properties [33] have been reported as potential causes. Upon fracture, similar to porcelain-alloys
[162], a thin porcelain layer remained attached to the zirconia surface, showing that cohesive
strength was lower than adhesive bond strength [27]. Even scientific evidence was lacking,
Fischer assumed that bond between zirconia and ceramic was chemical [86]. Others go for
mechanical interlocking added to cooling compressive stresses [163]. The ability of zirconia to
counteract crack propagation will result in a crack deflection [164]. Framework design must
provide uniform veneer support [14, 165, 166]. Pressable materials with an increase of the
crystalline content generally improved the mechanical properties [26].
Ceramic crowns made only of zirconia, monolithic zirconia crowns, are not used widely in
clinical practice because of the absence of a sound standard and the possibility of wear of the
opposing teeth due to the hardness of zirconia [65].
Even if zirconia frameworks are preferred in posterior situations, compared to other allceram
materials [5], some limitations still exist and proper diagnosis is critical for success [167] .
The quantity, size, and chemical properties of the crystals within the ceramic matrix will
determine the opacity of a ceramic material [168]. In-ceram Zirconia (VITA Zahnfabrik, Bad
Sackingen, Germany) is reported the least translucent when compared to other ceramics
[169, 170].
While success rate for 35% partially stabilized zirconia has been evaluated promising [171],
long-term clinical data remain rare [172]. The mechanical [173], esthetic [174], biocompatible
[175], and metal-like radiopaque [176] properties allow the zirconia ceramics to be versatile,
even though the opaque core limits their use in the anterior sextant [170]. Careful patient
selection and operating technique are essential. Bruxers, periodontally involved teeth exhibiting
increased mobility, and cantilever prostheses are to be avoided [172]. Fracture, located in the
area between the retainer and pontic is the primary mode of failure. Under high tensile stress, it
emanates from the gingival surface of the connectors, leading to catastrophic loss [177].
A framework design allowing for a uniform thickness and support of veneering porcelain has
been shown to optimize the strength of bilayered specimens [178]. Radial surface cracks can be
generated by Intaglio wall adjustments of the framework, either with a 50 micron or coarser
diamond rotary cutting instrument, and under dry or water cooling. This will compromise the
strength of the zirconia core [179]. Marginal fit has been reported similar to that of metal-
ceramic restorations [180]. Cementation of zirconia-based FPDs with either composite resin,
glass ionomer, or resin-modified glass ionomer cements have been proposed, even long-term
data are lacking [157, 174, 181].
Go to:

CONCLUSION
Zirconia restorative material is well-placed to satisfy esthetic requirements and to fulfill
functional requirements. Further studies should be conducted to resolve the complications that
may reduce restorations longevity.
Within its limitations, this review has pointed some of the strengths and weaknesses of this
promising material.
1. Zirconia is able to withstand physiological posterior forces.
2. Zirconia-veneer bonding is not yet well understood.
3. Studies to reduce veneer chipping should be conducted.
4. Ageing process, coloring pigments and liner materials have negative impact on the
veneer- zirconia bond strength.
5. Pressed veneer porcelain exhibit reduced fracture incidence compared to layered veneer.
6. New compatible high strength ceramic veneers would reduce chip-off incidence.
7. Framework design must provide anatomical support to the layer veneering ceramic.
Understanding each of these mechanisms will enhance the reliability of the zirconia as a
multipurpose material.
Go to:

ACKNOWLEDGEMENTS
None declared.
Go to:

CONFLICT OF INTEREST
The authors confirm that this article content has no conflicts of interest.
Go to:

REFERENCES
1. Bayne SC. Dental restorations for oral rehabilitation – testing of laboratory properties versus
clinical performance for clinical decision making. Review Article. J Oral Rehab. 2007;34:921–
32. [PubMed] [Google Scholar]
2. Guess C, Kulis A, Witkowskia S, Wolkewitz M, Zhang Y, Strub JR. Shear bond strengths
between different zirconia cores and veneering ceramics and their susceptibility to
thermocycling. Dent Mater. 2008;24:1556–67. [PubMed] [Google Scholar]
3. Heintze SD, Cavalleri A, Zellwegera G, Buchler A, Zappinia G. Fracture frequency of all-
ceramic crowns during dynamic loading in a chewing simulator using different loading and
luting protocols. Dent Mater. 2008;24:1352–61. [PubMed] [Google Scholar]
4. Schley JS, Heussen N, Reich S, Fischer J, Haselhuhn K, Wolfart S. Survival probability of
zirconia-based fixed dental prostheses up to 5 yr: a systematic review of the literature. Eur J Oral
Sci. 2010;118:443–50. [PubMed] [Google Scholar]
5. Conrad HJ, Seong WJ, Pesun IJ. Current ceramic materials and systems with clinical
recommendations: a systematic review. J Prosthetic Dent. 2007;98: 389–404. [PubMed] [Google
Scholar]
6. Morrmann WH, Stawarczyk B, Ender A, Sener B, Attin T, Mehl A. Wear characteristics of
current aesthetic dental restorative CAD/CAM materials: two-body wear. gloss
retenion.roughness and Martens hardness. J Mechanic Behav Biomed Mater . 2013; 20:113–
25. [PubMed] [Google Scholar]
7. Wittneben JG, Robert FW, Weber HP, Gallucci GO. A systematic review of the clinical
performance of CAD/CAM single-tooth restorations. Int J Prosthodont. 2009;22:466–
71. [PubMed] [Google Scholar]
8. Della Bona A, Robert Kelly J. The clinical success of all-ceramic
restorations. JADA. 2008;139(Suppl 4 ):8–13. [PubMed] [Google Scholar]
9. Tinschert J, Zwez D, Marx R, Anusavice KJ. Structural reliability of alumina-.
feldsar-.leucite-., mica- and zirconia-based ceramics. J Dent. 2000;28:529–
35. [PubMed] [Google Scholar]
10. Filser F, Kocher P, Weibel F. Reliability and strength of all-ceramic dental restorations
fabricated by direct ceramic machining (DCM). Int J Comput Dent. 2004:89–
106. [PubMed] [Google Scholar]
11. Christel P, Meunier A, Heller M, Torre JP, Peille CN. Mechanical properties and short-term
in vivo evaluation of yttrium-oxide-partially-stabilized zirconia. J Biomed Mater
Res. 1989;23:45–61. [PubMed] [Google Scholar]
12. Filser F, Kocher P, Gauckler LJ. Net-shaping of ceramic components by direct ceramic
machining. Assembly Autom. 2003;23:382–90. [Google Scholar]
13. Denry I, Kelly JR. State of the art of zirconia for dental applications. Dent
Mater. 2008;24:299–307. [PubMed] [Google Scholar]
14. Komine F, Blatz MB, Matsumura H. Current status of Zircobia-based fixed restorations. J
Oral Sci. 2010;52(4):531–9. [PubMed] [Google Scholar]
15. Pjetursson Bjarni E, Sailer I, Zwahlen M, Hammerle CHF. A systematic review of the
survival and complication rates of all-ceramic and metal-ceramic reconstructions after an
observation period of at least 3 years. Part I single crowns. Clin Oral Impl Res. 2007;18(Suppl
3 ):73–85. [PubMed] [Google Scholar]
16. Cho SH, Nagy WW, Goodman JT, Solomon E, Koike M. The effect of multiple firings on
the marginal integrity of pressable ceramic single crowns. J Prosthetic Dent. 2012;107:17–
23. [PubMed] [Google Scholar]
17. Galindo ML, Pedram P, Marinello CP. Estimating long-term survival of densely sintered
alumina crowns: A cohort study over 10 years. J Prosthetic Dent. 2011;106:23–
8. [PubMed] [Google Scholar]
18. Kohayama K, Hatakeyama E, Sasaki E, Azuma T, Karita K. Effect of sample thickness on
bite force studied with a multiple-point sheet sensor. J Oral Rehab. 2004;31:327–
34. [PubMed] [Google Scholar]
19. Goodacre CJ, Bernal G, Rungcharassaeng K, Kan JYK. Clinical complications in fixed
prosthodontics. J Prosthetic Dent. 2003;90:31–41. [PubMed] [Google Scholar]
20. Cehreli MC, Kokat A, Akca K. CAD/CAM Zirconia vs.slipcast glass-infiltrated
Alumina/Zirconia all-ceramic crowns: 2-year results of a randomized controlled clinical trial. J
App Oral Sci. 2009;17:49–55. [PMC free article] [PubMed] [Google Scholar]
21. Encke BS, Heydecke G, Wolkewitz M, Strub JR. Results of a prospective randomized
controlled trial of posterior ZrSiO(4)-ceramic crowns. J Oral Rehab. 2009;36:226–
35. [PubMed] [Google Scholar]
22. Ortorp A, Maria KL, Carlsson GE. A 3-year retrospective and clinical follow-up study of
zirconia single crowns performed in a private practice. J Dent. 2009;37:731–
6. [PubMed] [Google Scholar]
23. Scurria MS, Badder JD, Shugars DA. Meta-analysis of fixed partial denture survival:
prostheses and abutments. J Prosthetic Dent. 1998;79:459–64. [PubMed] [Google Scholar]
24. Kisi E, Howard C. Crystal structures of zirconia phases and their interrelation. Key Eng
Mater. 1998;153/154:1–35. [Google Scholar]
25. Heuer AH, Lange FF, Swain MV, Evans AG. Transformation toughening: an overview. J
Am Ceram Soc. 1986;69:i–iv. [Google Scholar]
26. Guazzato M, Albakry M, Ringer S, Swain M. Strength. fracture toughness and
microstructure of a selection of all-ceramic materials. part II. Zirconia-based dental
ceramics. Dent Mater. 2004;20:449–56. [PubMed] [Google Scholar]
27. Koutayas SO, Vagkopoulos T, Pelekanos S, Koidis P, Strub JR. Zirconia in Dentistry: part
2.evidence-based clinical breakthrough. Eur J Esthet Dent. 2009;4:348–80. [PubMed] [Google
Scholar]
28. Deville S, Chevalier J, Gremillard L. Influence of surface finish and residual stresses on the
ageing sensitivity of biomedical grade zirconia. Biomaterials. 2006;27:2186–
92. [PubMed] [Google Scholar]
29. Chevalier J, Cales B, Drouin JM. Low-temperature aging of Y-TZP ceramics. J Am Ceram
Soc. 1999;82:2150–4. [Google Scholar]
30. Guo X. On the degradation of zirconia ceramics during low-temperature annealing in water
or water vapor. J Phys Chem Solids. 1999;60:539–46. [Google Scholar]
31. Lin JD, Duh JG, Lo CL. Mechanical properties and resistance to hydrothermal aging of
ceria- and yttriadoped tetragonal zirconia ceramics. Mater Chem Phys. 2002;87:808–18. [Google
Scholar]
32. Papanagiotou HP, Morgano S, Giordano RA, Pober R. In vitro evaluation of low-temperature
aging effects and finishing procedures on the flexural strength and structural stability of Y-TZP
dental ceramics. J Prosthetic Dent. 2006;96:154–64. [PubMed] [Google Scholar]
33. Rahaman MN, Li Y, Bal BS, Huang W. Functionally graded bioactive glass coating on
magnesia partially stabilized zirconia for enhanced biocompatibility. J Mater Sci Mater Med
(Mg-PSZ) 2008;19:2325–33. [PubMed] [Google Scholar]
34. Kim DJ, Myung-Hyun L, Lee DY, Han JS. Mechanical properties. phase stabiity.and
biocompatibility of (Y., Nb)-TZP/Al2)O3) composite abutments for dental implant. J Biomed
Mater Res. 2000;53:438–43. [PubMed] [Google Scholar]
35. Heness G, Ben-Nissan B. Innovative bioceramics. Mater Forum. 2004;27:104–14. [Google
Scholar]
36. Suarez MJ, Lozano JF, Paz Salido M, Martinez F. Three-year clinical evaluation of In-Ceram
Zirconia posterior FPDs. Int J Prosthodont. 2004;17:35–8. [PubMed] [Google Scholar]
37. Esquivel-Upshaw JF, Anusavice K, Young H, Jones J, Gibbs C. Clinical performance of
lithia disilicate-based core ceramic for three-unit posterior FPDs. Int J
Prosthodont. 2004;17:469–75. [PubMed] [Google Scholar]
38. Oh WS, Anusavice K. Effect of connector design on the fracture resistance of all-ceramic
fixed partial dentures. J Prosthetic Dent. 2002;87:536–42. [PubMed] [Google Scholar]
39. Oh WS, Anusavice KJ. Effect of connector design on the fracture resistance of all-ceramic
fixed partial dentures. J Prosthetic Dent. 2002;87:536–42. [PubMed] [Google Scholar]
40. Raigrodski AJ. Contemporary materials and technologies for all-ceramic fixed partial
dentures: A review of the literature. J Prosthetic Dent. 2004;92:557–62. [PubMed] [Google
Scholar]
41. Miller LL. Framework design in ceramo-metal restorations. Dent Clin North
Am. 1977;21:699–716. [PubMed] [Google Scholar]
42. Att W, Grigoriadou M, Strub JR. ZrO2 three-unit fixed partial dentures: comparison of
failure load before and after exposure to a mastication simulator. J Oral Rehabil. 2007;34:282–
90. [PubMed] [Google Scholar]
43. Tinschert J, Gerd N, Mautsch W, Augthun M, Spiekermann H. Fracture resistance of
lithium-disilicate-. alumna-.and zirconia-based three-unit fixed partial dentures. Int J
Prosthodont. 2001; 14:231–8. [PubMed] [Google Scholar]
44. Sundh A, Sjogren G. Fracture resistance of all-ceramic zirconia bridges with differing phase
stabilizers and quality of sintering. Dent Mater. 2006;22:778–84. [PubMed] [Google Scholar]
45. Ttinschert J, Natt G, Mautsch W, Augthun M, Spiekermann H. Fracture resistance of lithium
disilicate-. alumna-.and zirconia-based three-unit fixed partial dentures a laboratory study. Int J
Prosthodont . 2001;14: 231–8. [PubMed] [Google Scholar]
46. Ttinschert J, Natt G, Mohrbotter N, Spiekermann H, Schulze KA. Lifetime of alumina- and
zirconia ceramics used for crown and bridge restorations. J Biomed Mater Res B Appl
Biomater. 2007;80:317–21. [PubMed] [Google Scholar]
47. Plengsombut K, Brewer JD, Monaco EA Jr, Davis EL. Effect of two connector designs on
the fracture resistance of all-ceramic core materials for fixed dental prostheses. J Prosthetic
Dent. 2009;101:166–73. [PubMed] [Google Scholar]
48. Bahat Z, Mahmoood DJ, Vult von Steyern P. Fracture strength of three-unit fixed partial
denture cores (Y-TZP) with different connector dimension and design. Swed Dent
J. 2009;33:49–59. [PubMed] [Google Scholar]
49. Vult von Steyrn P. All-ceramic fixed partial dentures.Studies on aluminum oxide- and
zirconium dioxide-based ceramic systems. Swed Dent J. 2005;(Suppl ):1–69. [PubMed] [Google
Scholar]
50. Tsumita M, Kokubo Y, Vult von Steyern P, Fukushima S. Effect of framework shape on the
fracture strength of implant-supported all-ceramic fixed partial dentures in the molar region. J
Prosthodont. 2008;17:274–85. [PubMed] [Google Scholar]
51. Vult von Steyrn P, Carlsson P, Nilner K. All-ceramic fixed partial dentures designed
according to the DC-Zirkon technique.A 2-year clinical study. J Oral Rehabil. 2005;32:180–
7. [PubMed] [Google Scholar]
52. Larsson C, Holm L, Lovgern N, Kokubo Y. Vult von Steyern.Fracture strength of four-unit
Y-TZP core designed with varying connector diameter: an in-vitro study. J Oral
Rehabil. 2007;34:702–9. [PubMed] [Google Scholar]
53. von Steyern PV. All-ceramic fixed partial dentures.Studies on aluminum oxide- and
zirconium dioxide-based ceramic systems. Swed Dent J Suppl. 2005;173:1–
69. [PubMed] [Google Scholar]
54. Larsson C, Volt von Steyern P, Sunzel B, Nilner K. All-ceramic two- and five-unit implant-
supported reconstructions.A randomzed.prospective clinical trial. Swed Dent J . 2006;30:45–
53. [PubMed] [Google Scholar]
55. Anderson RJ, Janes GR, Sabella LR, Morris HF. Comparison of the performance on
prosthodontic criteria of several alternative alloys used for fixed crown and partial denture
restorations: Department of Veterans Affairs Cooperative Studies project 147. J Prosthetic
Dent. 1993;69:1–8. [PubMed] [Google Scholar]
56. Walter M, Reppel P, Boning K, Freesmeyer WB. Six year follow-up of titanium and high-
gold porcelain-fused-to-metal fixed partial dentures. J Oral Rehabil. 1999;26:91–
6. [PubMed] [Google Scholar]
57. Sandhage KH, Yurek GJ. Direct and indirect dissolution of sapphire in calcia-magnesia-
alumina-silica melts: dissolution kinetics. J Am Ceram Soc. 1990;73:3633–42. [Google Scholar]
58. Kelly JR, editor. In Cambridge.: MA: 1989 Harvard University.; Clinical fracture behavior
and colloidal processing of glass-matrix dental ceramics. [Google Scholar]
59. Durschang B, Raether F. Development of a glass-infiltrated ceramic for dental
applications. Fraunhofer ISC Annual Report
http://www.isc.fraunhofer.de/german/improfil/presse/publikationen/media/e60-61.pdf. 2002 
[Google Scholar]
60. Kim DJ. Effect of Ta2O5. N2O5.and HfO2 alloying on the transformability of Y2O3-
stabilized tetragonal ZrO2. J Am Ceram Soc. 1990; 73:115–20. [Google Scholar]
61. Schubert H. Anisotropic thermal expansion coefficients of Y2O3-stabilized tetragonal
zirconia. J Am Ceram Soc. 1986;69:270–1. [Google Scholar]
62. Matsui K, Horikoshi H, Ohmichi N, Ohgai M, Yoshida H, Ikuara Y. Cubic-formation and
grain-growth mechanisms in tetragonal zirconia polycrystal. J Am Ceram Soc. 2003;86:401–
8. [Google Scholar]
63. Kobayashi K, Kuwajima H, Masaki T. Phase change and mechanical properties of ZrO2-
Y2O3 solid electrolyte after aging. Solid State Ionics. 1981;3(4 ):489–95. [Google Scholar]
64. Kou W, Molin M, SjoGren G. Surface roughness of five different dental ceramic core
materials after grinding and polishing. J Oral Rehab. 2006;33:117–24. [PubMed] [Google
Scholar]
65. Kim MJ, Oh SH, Kim JH , et al. Wear evaluation of the human enamel opposing different Y-
TZP dental ceramics and other porcelains. J Dent. 2012;40:979–88. [PubMed] [Google Scholar]
66. Oh WS, Delong R, Anusavice KJ. Factors affecting enamel and ceramic wear: a literature
review. J Prosthetic Dent. 2002;87: 451–9. [PubMed] [Google Scholar]
67. Denry I. How and when does fabrication damage adversely affect the clinical performance of
ceramic restorations?. Dent Mater. 2013;29:85–96. [PubMed] [Google Scholar]
68. McLean JW. Evolution of dental ceramics in the twentieth century. J Prosthetic
Dent. 2001;85:61–6. [PubMed] [Google Scholar]
69. Aboushelib MN, de Jager N, Kleverlaan CJ , et al. Microtensile bond strength of different
components of core veneered all-ceramic restorations. Dent Mater. 2005;21:984–
91. [PubMed] [Google Scholar]
70. Aboushelib MN, Kleverlaan CJ, Feilzer AJ. Effect of zirconia type on its bond strength with
different veneer ceramics. J Prosthodont. 2008;17:401–8. [PubMed] [Google Scholar]
71. Walter M, Reppel PD, Boning K, Freesmeyer WB. Six-year follow-up of titanium and high-
gold porcelain-fused-to-metal fixed partial dentures. J Oral Rehab. 1999;26:91–
6. [PubMed] [Google Scholar]
72. Coornaert J, Adriaens P, De Boever J. Long-term clinical study of porcelain-fused-to-gold
restorations. J Prosthetic Dent. 1984;51:338–42. [PubMed] [Google Scholar]
73. Valderhaug J. A 15-year clinical evaluation of fixed prosthodontics. Acta Odontol
Scand. 1991;49:35–40. [PubMed] [Google Scholar]
74. Sailer I, Feher A, Filser F. Prospective clinical study of zirconia posterior fixed partial
dentures: 3-year follow up. Quint Int. 2006;37:685–93. [PubMed] [Google Scholar]
75. Raigrodski AJ, Gerard JC, Potiket N. The efficacy of posterior three-unit zirconium-oxide-
based ceramic fixed partial dental prostheses: a prospective clinical pilot study. J Prosthetic
Dent. 2006;96:237–44. [PubMed] [Google Scholar]
76. Schmitter M, Mueller D, Rues S. Chipping behaviour of all-ceramic crowns with zirconia
framework and CAD/CAM manufactured veneer. J Dent. 2012;40:154–62. [PubMed] [Google
Scholar]
77. Sailer I, Gottnerb J, Kanel S, Hämmerle CH. Randomized controlled clinical trial of zirconia-
ceramic and metal-ceramic posterior fixed dental prostheses: a 3-year follow-up. Int J
Prosthodont. 2009;22:553–60. [PubMed] [Google Scholar]
78. Zhang Y, Lawn BR. Fatigue sensitivity of Y-TZP to microscale sharp-contact flaws. J
Biomed Mater Res: Appl Biomater. 2005;72B:388–92. [PubMed] [Google Scholar]
79. Zhang Y, Pajares A, Lawn BR. Fatigue and damage tolerance of Y-Y-TZP ceramics in
layered biomechanical systems. J Biomed Y-Mater Res B Appl Biomater. 2004;71B:166–
71. [PubMed] [Google Scholar]
80. Zhang Y, Lawn BR, Rekow ED, Thompson VP. Effect of sandblasting on the long-term
performance of dental ceramics. J Biomed Mater Res B: Appl Biomater. 2004;71B:381–
6. [PubMed] [Google Scholar]
81. Marchack BW, Futatsuki Y, Marchack CB, White SN. Customization of milled zirconia
coping for all-ceramic crowns: a clinical report. J Prosthetic Dent. 2008;99:169–
73. [PubMed] [Google Scholar]
82. Segal BS. Retrospective assessment of 546 all-ceramic anterior and posterior crowns in a
general practice. J Prosthetic Dent. 2001;85:544–50. [PubMed] [Google Scholar]
83. Mitov G, Heintze SD, Walz S, Woll K, Muecklichd F, Pospiecha P. Wear behavior of dental
Y-TZP ceramic against natural enamel after different finishing procedures. Dent
Mater. 2012;28:909–18. [PubMed] [Google Scholar]
84. Aboushelib MN, Kleverlaan CJ, Feilzer AJ. Microtensile bond strength of different
components of core veneered all-ceramic restorations.Part II: zirconia veneering ceramics. Dent
Mater. 2006;22:857–63. [PubMed] [Google Scholar]
85. Fischer J, Stawarczyk B, Tomic M, Strub JR, Hammerle CHF. Effects of thermal misfit
between different veneering ceramics and zirconia frameworks on in vitro fracture load of single
crowns. Dent Mater. 2007;26(6):766–72. [PubMed] [Google Scholar]
86. Fischer J, Grohmann P, Stawarczyk B. Effect of zirconia surface treatments on the shear
strength of zirconia/veneering ceramic composites. Dent Mater J. 2008;27:448–
54. [PubMed] [Google Scholar]
87. Komine F, Strub JR, Matsumura H. Bonding between layering materials and zirconia
frameworks. Jap Dental Sci Rev. 2012;48:153–61. [Google Scholar]
88. Saito A, Komine F, Blatz MB, Matsumura H. A comparison of bond strength of layered
veneering porcelains to zirconia and metal. J Prosthetic Dent. 2010;104:247–
57. [PubMed] [Google Scholar]
89. Shell JS, Nielsen JP. Study of the bond between gold alloys and porcelain. J Dent
Res. 1962;41:1424–37. [PubMed] [Google Scholar]
90. Knap FJ, Ryge G. Study of bond strength of dental porcelain fused to metal. J Dent
Res. 1966;45:1047–51. [PubMed] [Google Scholar]
91. Hermann I, Bhowmick S, Zhang Y, Lawn BR. Competing fracture modes in brittle materials
subject to concentrated cyclic loading in liquid environments: Trilayer structures. J Mater
Res. 2006;21:512–21. [Google Scholar]
92. Mora GP, O'Brien WJ. Thermal shock resistance of core reinforced all-ceramic crown
systems. J Biomed Mater Res. 1994;28:189–94. [PubMed] [Google Scholar]
93. Komine F, Saito A, Kobayashi K, Koizuka M, Koizumi H, Matsumura H. Effect of cooling
rate on shear bond strength of veneering porcelain to a zirconia ceramic material. J Oral
Sci. 2010;52: 647–52. [PubMed] [Google Scholar]
94. Gostemeyer G, Jendras M, Dittmer MP, Bach FW, Stiesch M, Kohorst P. Influence of
cooling rate on zirconia/veneer interfacial adhesion. Acta Biomater. 2010;6:4532–
8. [PubMed] [Google Scholar]
95. Swain MV. Unstable cracking (chipping) of veneering porcelain on all-ceramic dental
crowns and fixed partial dentures. Acta Biomater. 2009;5:1668–77. [PubMed] [Google Scholar]
96. Taskonak B, Borges GA, Mecholsky JJ Jr, Anusavice KJ, Moore BK, Yan J. The effects of
viscoelastic parameters on residual stress development in a zirconia/glass bilayer dental
ceramic. Dent Mater. 2008;24:1149–55. [PubMed] [Google Scholar]
97. Rues S, Kroger E, Muller D, Schmitter M. Effect of firing protocols on cohesive failure of
all-ceramic crowns. J Dent. 2010;38:987–94. [PubMed] [Google Scholar]
98. Fischer J, Stawarczyck B, Sailer I, Ha¨mmerle CH. Shear bond strength between veneering
ceramics and ceria-stabilized zirconia/alumina. J Prosthetic Dent. 2010;103:267–
74. [PubMed] [Google Scholar]
99. Hong JK, Hyum PL, Park YJ, Vang MS. Effect of zirconia surface treatments on the shear
bond strength of veneering ceramic. J Prosthetic Dent. 2011;105:315–22. [PubMed] [Google
Scholar]
100. Beuer F, Schweiger J, Eichberger M, Kappert HF, Gernet W, Edelhoff D. High-strength
CAD/CAM-fabricated veneering material sintered to zirconia copings - a new fabrication mode
for all-ceramic restorations. Dent Mater. 2009;25:121–8. [PubMed] [Google Scholar]
101. Beuer F, Edelhoff D, Gernet W, Sorensen JA. Three-year clinical prospective evaluation of
zirconia-based posterior Foxed Dental Prostheses (FDPs). Clin Oral Investigat. 2009;13:445–
51. [PubMed] [Google Scholar]
102. Coffey JP, Anusavice KJ, DeHoff PH, Lee RB, Hojjatie B. Influence of contraction
mismatch and cooling rate on flexural failure of PFM systems. J Dent Res. 1988;67:61–
5. [PubMed] [Google Scholar]
103. Isgro G, Addison O, Fleming GJP. Transient and residual stresses induced during the
sintering of two dentine ceramics. Dent Mater. 2010;27:379–85. [PubMed] [Google Scholar]
104. McLean JW, Hughes TH. The reinforcement of dental porcelain with ceramic oxides. Br
Dent J. 1965;119:251–67. [PubMed] [Google Scholar]
105. Isgro G, Addison O, Fleming GJP. Transient and residual stresses in a pressable glass-
ceramic before and after resin-cement coating determined using profilometry. J
Dent. 2011;39:368–75. [PubMed] [Google Scholar]
106. Tholey MJ, Swain MV, Thiel N. SEM observations of porcelain Y-TZP interface. Dent
Mater. 2009;25:857–62. [PubMed] [Google Scholar]
107. Albashaireh ZSM, Ghazal M, Kern M. Two-body wear of different ceramic materials
opposed to zirconia ceramic. J Prosthetic Dent. 2010;104:105–13. [PubMed] [Google Scholar]
108. Stawarczyk B, Özcan M, Roos M, Trottmann I, Sailer I, Hämmerle CHF. Load-bearing
capacity and failure types of anterior zirconiacrowns veneered with overpressing and layering
techniques. Dent Mater. 2011;27:1045–53. [PubMed] [Google Scholar]
109. Drummond JL. Ceramic behavior under different environmental and loading
conditions.Dental materials in vivo: aging and related phenomena. Quinte Chicago IL . 2003:35–
45. [Google Scholar]
110. Holden JE, Goldstein GR, Hittelman EL, Clark EA. Comparison of the marginal fit of
pressable ceramic to metal ceramic restorations. J Prosthodont. 2009;18:645–
8. [PubMed] [Google Scholar]
111. Kim MJ, Kim YK, Kim KH, Kwon TY. Shear bond strengths of various luting cements to
zirconia ceramic: surface chemical aspects. J Dent. 2011;39:795–803. [PubMed] [Google
Scholar]
112. Albrecht T, Kirsten A, Kapperta HF, Fischerb H. Fracture load of different crown systems
on zirconia implant abutments. Dent Mater. 2011;27:298–303. [PubMed] [Google Scholar]
113. Guess C, Zavanelli R, Silva N, Bonfante E, Coelho P, Thompson V. Monolithic CAD/CAM
Lithium Disilicate versus Veneered Y-TZP crowns: comparison of failure modes and reliability
after fatigue. Int J Prosthodont. 2010;23:343–442. [PubMed] [Google Scholar]
114. Christensen RP, Eriksson KA, Ploeger BJ. Clinical performance of PFM. zirconia and
alumina three-unit posterior protheses. IADR Toronto. 2008:105962. [Google Scholar]
115. Ishibe M, Raigrodski A, Flinn BD, Chung KH, Spiekerman C, Winter RR. Shear bond
strengths of pressed and layered veneering ceramics to high-noble alloy and zirconia cores. J
Prosthetic Dent. 2011;106:29–37. [PubMed] [Google Scholar]
116. 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:e78–
93. [PubMed] [Google Scholar]
117. Guess PC, Zhang Y, Thompson VP. Effect of veneering techniques on damage and
reliability of Y-TZP trilayers. Eur J Esthet Dent. 2009;4:262–76. [PubMed] [Google Scholar]
118. Plengsombut K, Brewer JD, Monaco EA Jr, Davis EL. Effect of two connector designs on
the fracture resistance of all-ceramic core materials for fixed dental prostheses. J Prosthetic
Dent. 2009;101:166–73. [PubMed] [Google Scholar]
119. International Organization for Standardization I metal-ceramic dental restorative
systems. Organization for Standardization Geneva Int. 1999 [Google Scholar]
120. Al-Dohan HM, Yaman P, Dennison JB, Razzoog ME, Lang BR. Shear strength of core-
veneer interface in bi-layered ceramics. J Prosthetic Dent. 2004;91:349–55. [PubMed] [Google
Scholar]
121. Suese K. Comparison of bond strength of porcelain fused to core materials of metal and
zirconia. J Osaka Dent Univ. 2010;44:41–7. [Google Scholar]
122. Ashkanani HM, Raigrodski A, Flinn BD, Heindl H, Mancl LA. Flexural and shear strengths
of ZrO2 and a high-noble alloy bonded to their corresponding porcelains. J Prosthetic
Dent. 2008;100: 274–84. [PubMed] [Google Scholar]
123. Silva NRFA, Bonfante EA, Zavanelli RA, Thompson VP, Ferencz JL, Coelho GP.
Reliability of metalloceramic and Zirconia-based ceramic crowns. J Dent
Res. 2010;89(10):1051–6. [PMC free article] [PubMed] [Google Scholar]
124. Mackert JrJr, Ringle R, Parry EE, Evans AL, Fairhurst CW. The relationship between oxide
adherence and porcelain-metal bonding. J Dent Res. 1988;67:474–8. [PubMed] [Google Scholar]
125. Schweitzer DM, Goldstein G, Ricci JL, Silva NR, Hittelman EL. Comparison of bond
strength of a pressed ceramic fused to metal versus feldspathic porcelain fused to metal. J
Prosthodont. 2005;14:239–47. [PubMed] [Google Scholar]
126. Ernsberger FM. The role of molecular water in the diffusive transport of protons in
glasses. Physics Chem Glasses. 1980;21:146–9. [Google Scholar]
127. Newton RG. The durability of glass-a review. Glass Technol. 1985;26:21–38. [Google
Scholar]
128. Peterson IM, Wuttiphan S, Lawn BR, Chyung K. Role of microstructure on contact damage
and strength degradation of micaceous glass-ceramics. Dent Mater. 1998;14:80–
9. [PubMed] [Google Scholar]
129. Zhang Y, Song J, Lawn BR. Deep-penetrating conical cracks in brittle layers from
hydraulic cyclic contact. J Biomed Mater Res B Appl Biomater. 2005;73:186–
93. [PubMed] [Google Scholar]
130. Att W, Stamouli K, Gerds T, Strub JR. Fracture resistance of different zirconium dioxide
three-unit all-ceramic fixed partial dentures. Acta Odontol Scand. 2007;65:14–
21. [PubMed] [Google Scholar]
131. Beuer F, Bastian S, Naumann M, Sorensen JA. Load-bearing capacity of all-ceramic three-
unit fixed partial dentures with different computer-aided design (CAD)/computer-aided
manufacturing (CAM) fabricated framework materials. Eur J Oral Sci. 2008;116:381–
6. [PubMed] [Google Scholar]
132. Petridis H, Hirayama H, Kugel G, Habib C, Garefis P. Shear bond strength of techniques
for bonding esthetic veneers to metal. J Prosthetic Dent. 1999;82:608–14. [PubMed] [Google
Scholar]
133. Tsalouchou E, Cattell MJ, Knowles JC, Pittayachawan P, McDonald A. Fatigue and
fracture properties of yttria partially stabilized zirconia crown systems. Dent
Mater. 2008;24:308–18. [PubMed] [Google Scholar]
134. Fischer J, Stawarzcyk B, Hämmerle CH. Flexural strength of veneering ceramics for
zirconia. J Dent. 2008;36(5):316–21. [PubMed] [Google Scholar]
135. Raptis NV, Michalakis K, Hirayama H. Optical behavior of current ceramic systems. Int J
Periodont Restor Dent. 2006;26:31–41. [PubMed] [Google Scholar]
136. Ardlin BI. Transformation-toughened zirconia for dental inlays. crowns and bridges:
chemical stability and effect of low-temperature aging on flexural strength and surface
structure. Dent Mater. 2002;18:590–5. [PubMed] [Google Scholar]
137. Heffernan MJ, Aquilino S, Diaz-Arnold AM , et al. Relative translucency of six all-ceramic
systems.Part II: core and veneer materials. J Prosthetic Dent. 2002;88:10–5. [PubMed] [Google
Scholar]
138. Devigus A, Lombardi G. Shading Vita In-ceram YZ substructures: influence on value and
chroma. part II. Int J Comput Dent. 2004;7:379–88. [PubMed] [Google Scholar]
139. Raigrodski AJ. Contemporary all-ceramic fixed partial dentures a review. Dent Clin North
Am. 2004;48:531–44. [PubMed] [Google Scholar]
140. Aboushelib MN, Kleverlaan C, Feilzer AJ. Microtensile bond strength of different
components of core veneered all-ceramic restorations. part 3: double veneer technique. J
Prosthodont. 2008;17:9–13. [PubMed] [Google Scholar]
141. Chen PL, Chen IW. Grain boundary mobility in Y2O3: defect mechanism and dopant
effects. J Am Ceram Soc. 1996;79:1801–9. [Google Scholar]
142. Deville S, Gremillard L, Chevalier J, Fantozzi G. A critical comparison of methods for the
determination of the aging sensitivity in biomedical grade yttria-stabilized zirconia. J Biomed
Mater Res B Appl Biomater. 2005;72:239–45. [PubMed] [Google Scholar]
143. Pittayachawan P, McDonald A, Petrie A, Knowles J. The biaxial flexural strength and
fatigue property of Lava(TM) Y-TZP dental ceramic. Dent Mater. 2007;23:1018–
29. [PubMed] [Google Scholar]
144. Carlsson GE. Critical review of some dogmas in prosthodontics. J Prosthodont
Res. 2009;53:3–10. [PubMed] [Google Scholar]
145. DeLong R, Douglas W. Development of an artificial oral environment for the testing of
dental restoratives: bi-axial force and movement control. J Dent Res. 1983;62:32–
6. [PubMed] [Google Scholar]
146. Nico H, Creugers J, Arnd F, Kayser Martin A, van 't Hof. A meta-analysis of durability data
on conventional fixed bridges. Commun Dent Oral Epidemiol. 1994;22:448–
52. [PubMed] [Google Scholar]
147. Yeo IS, Yang JH, Lee JB. In vitro marginal fit of three all-ceramic crown systems. J
Prosthetic Dent. 2003;90:459–64. [PubMed] [Google Scholar]
148. De Jager N, Pallav P, Feilzer AJ. The influence of design parameters on FEA-determined
stress distribution in CAD-CAM produced all-ceramic crowns. Dent Mater. 2005;21:242–
51. [PubMed] [Google Scholar]
149. Burke FJ, Fleming GJ, Nathanson D, Marquis PM. Are adhesive technologies needed to
support ceramics?.an assessment of the current evidence. J Adhes Dent. 2002;4:7–
22. [PubMed] [Google Scholar]
150. Marquardt P, Strub JR. Survival rates of IPS Empress 2 all-ceramic crowns and fixed partial
dentures: results of a 5-year prospective clinical study. Quint Int. 2006;37:253–
9. [PubMed] [Google Scholar]
151. Piconi C, Maccauro G. Zirconia as a ceramic material. Biomaterials. 1999;20:1–
25. [PubMed] [Google Scholar]
152. Yanagida H, Koumoto K, Miyayama M, editors. Chichester. UK:: John Wiley & Sons Ltd;
1996. The chemistry of ceramics. pp. 247–9. [Google Scholar]
153. Guazzato M, Quach L, Albakry M, Swain MV. Influence of surface and heat treatments on
the flexural strength of Y-TZP dental ceramic. J Dent. 2005;33:9–18. [PubMed] [Google
Scholar]
154. Sailer I, Feher A, Filser F, Gauckler LJ, Luthy H, Hammerle CH. Five-year clinical results
of zirconia frameworks for posterior fixed partial dentures. Int J Prosthodont. 2007;20:383–
8. [PubMed] [Google Scholar]
155. Chong KH, Chai J, Takahashi Y, Wozniak W. Flexural strength of In-Ceram alumina and
In-Ceram zirconia core materials. Int J Prosthodont. 2002;15:183–8. [PubMed] [Google Scholar]
156. Bindl A, Mormann WH. Marginal and internal fit of all-ceramic CAD/CAM crown copings
on chamfer preparations. J Oral Rehab. 2005;32:441–7. [PubMed] [Google Scholar]
157. Sadowsky SJ. An overview of treatment considerations for esthetic restorations: a review of
the literature. J Prosthetic Dent. 2006;96:433–42. [PubMed] [Google Scholar]
158. Studart AR, Filser F, Kocher P, Luthy H, Gauckler LJ. Mechanical and fracture behavior of
veneer-framework composites for allceramic dental bridges. Dent Mater. 2007;23:115–
23. [PubMed] [Google Scholar]
159. Aboushelib MN, Feilzer A, de Jager N, Kleverlaan CJ. Prestresses in bilayered allceramic
restorations. J Biomed Mater Res B Appl Biomater. 2008;87:139–45. [PubMed] [Google
Scholar]
160. de Kler M, deJager N, Meegdes M, van der Zel JM. Influence of thermal expansion
mismatch and fatigue loading on phase changes in porcelain veneered Y-TZP zirconia discs. J
Oral Rehabil. 2007;34:841–7. [PubMed] [Google Scholar]
161. Wang H, Aboushelib MN, Feilzer AJ. Strength influencing variables on CAD/CAM
zirconia frameworks. Dent Mater. 2008;24:633–8. [PubMed] [Google Scholar]
162. Oilo G, Johanson B, Syverud M. Bond strength of porcelain to dental alloys-an evaluation
of two test methods. Scand J Dent Res. 1981;89:289–96. [PubMed] [Google Scholar]
163. Ban S, Sato H, Suehiro Y, Nakanishi H, Nawa M. Biaxial flexure strength and low
temperature degradation of Ce-TZP/Al2O3 nanocomposite and Y-TZP as dental restoratives. J
Biomed Mater Res B Appl Biomater. 2008;87:492–8. [PubMed] [Google Scholar]
164. Kim B, Zhang Y, Pines M, Thompson VP. Fracture of porcelain-veneered structures in
fatigue. J Dent Res. 2007;86:142–6. [PubMed] [Google Scholar]
165. Donovan TE. Factors essential for successful all-ceramic restorations. J Am Dent
Assoc. 2008;139(Suppl ):14S–8S. [PubMed] [Google Scholar]
166. Hermann I, Bhowmick S, Lawn BR. Role of core support material in veneer failure of
brittle layer structures. J Biomed Mater Res B Appl Biomater. 2007;82:115–
21. [PubMed] [Google Scholar]
167. Aboushelib MN, deJager N, Kleverlaan CJ, Feilzer AJ. Microtensile bond strength of
different components of core veneered all-ceramic restorations. Dent Mater. 2005;21:984–
91. [PubMed] [Google Scholar]
168. Wassermann A, Kaiser M, Strub JR. Clinical long-term results of VITA Inceram classic
crowns and fixed partial dentures: a systematic review. Int J Prosthodont. 2006;19:355–
63. [PubMed] [Google Scholar]
169. Heffernan MJ, Aquilino SA, Diaz-Arnold AM, Haselton DR, Stanford CM, Vargas MA.
Relative translucency of six all-ceramic systems.part I: core materials. J Prosthetic
Dent. 2002;88:4–9. [PubMed] [Google Scholar]
170. Heffernan MJ, Aquilino S, Diaz-Arnold AM, Haselton DR, Stanford CM, Vargas M.
Relative translucency of six all-ceramic systems.Part II: core and veneer materials. J Prosthetic
Dent. 2002;88:10–5. [PubMed] [Google Scholar]
171. Suarez MJ, Lozano J, Paz Salido M, Martinez F. Three-year clinical evaluation of In-Ceram
Zirconia posterior FPDs. Int J Prosthodont. 2004;17:35–8. [PubMed] [Google Scholar]
172. Raigrodski AJ, Chiche G. The safety and efficacy of anterior ceramic fixed partial dentures:
a review of the literature. J Prosthetic Dent. 2001;86:520–5. [PubMed] [Google Scholar]
173. Luthy H, Filser F, Loeffel O, Schumacher M, Gaucker LJ, Hammerle CHF. Strength and
reliability of four-unit all-ceramic posterior bridges. Dent Mater. 2005;21:930–
7. [PubMed] [Google Scholar]
174. Raigrodski AJ. Contemporary materials and technologies for all-ceramic fixed partial
dentures: a review of literature. J Prosthetic Dent. 2004;92:557–62. [PubMed] [Google Scholar]
175. Scarano A, Di Carlo F, Quaranta M, Piattelli A. Bone response to zirconia ceramic
implants: an experimental study in rabbits. J Oral Implantol. 2003;29:8–12. [PubMed] [Google
Scholar]
176. Raigrodski AJ. Contemporary all-ceramic fixed partial dentures: a review. Dent Clin North
A. 200;48:531–44. [PubMed] [Google Scholar]
177. Tinschert J, Natt G, Mautsch W, Augthun M, Spikermann H. Fracture resistance of lithium
disilicate-. alumna-.and zirconia-based three-unit fixed partial den-tures a laboratory study. . Int J
Prosthodont. 2001;14:231–8. [PubMed] [Google Scholar]
178. Guazzato M, Proos K, Quach L, Swain MV. Strength reliability and mode of fracture of
bilayered porcelain/zirconia (Y-TZP) dental ceramics. Biomaterials. 2004;25:5045–
52. [PubMed] [Google Scholar]
179. Kosmac T, Oblak C, Jevnikar P, Funduk N, Marion L. Strength and reliability of surface
treated Y-TZP dental ceramics. J Biomed Mater Res. 2000;53:304–13. [PubMed] [Google
Scholar]
180. Reich S, Wichmann M, Nkenke E, Proeschel P. Clinical fit of all-ceramic three-unit fixed
partial dentures. generated with three different CAD/CAM systems. . Eur J Oral
Sci. 2005;113:174–9. [PubMed] [Google Scholar]
181. Ernest Claus-Peter, Cohnen U, Stender Elmar, Willershausen Brita. In vitro retentive
strength of zirconium oxide ceramic crowns using different luting agents. J Prosthetic
Dent. 2005;93:551–8. [PubMed] [Google Scholar]
Monolithic zirconia: A review of the literature
Zeynep Özkurt-Kayahan*

Department of Prosthodontics, Faculty of Dentistry, Yeditepe University, Istanbul, Turkey

*Corresponding Author:
Zeynep Özkurt-Kayahan
Faculty of Dentistry
Department of Prosthodontics, Ba?dat cad. No: 238
Yeditepe University
34728, Goztepe, Istanbul
Turkey
Accepted date: May 04, 2016

Visit for more related articles at Biomedical Research


Abstract
Veneer cracking or chipping is the major complication of the zirconia based restorations.
Monolithic zirconia has been introduced to overcome this problem, as well as to use in
patients with limited interocclusal space. Many research articles on monolithic zirconia
crowns have been published in the last years. The aim of this review article was to present
data about the wear, surface roughness, fracture strength, optical properties, and marginal
fit of monolithic zirconia. A PubMed search was conducted with the terms of “zirconia” with
“monolithic”, “full-contour”, “solid”, “translucent”, “anatomiccontoured”, “un-veneered”,
“non-veneered”, “full-coverage”. Based on the results of these studies, monolithic zirconia
crowns with polished surfaces have been shown to cause the lowest wear on the
antagonists compared to glazed zirconia. The fracture strength of monolithic zirconia has
been found higher than veneered zirconia. Monolithic zirconia may be a promising future
and long-term follow-up studies are needed to determine whether it may be an alternative
to conventional veneered zirconia.

Keywords
Monolithic zirconia, Full-contour zirconia, Clinical studies, In vitro studies, Review.

Introduction
Zirconia or zirconium dioxide (ZrO2) is a highly attractive ceramic material in prosthodontics
due to its excellent mechanical properties related to transformation toughening, which are
the highest ever reported for any dental ceramic and enhanced natural appearance
compared to metal-ceramics [1-3]. It is widely used to build prosthetic devices because of
its good chemical properties, dimensional stability, high mechanical strength, toughness,
and a Young’s modulus (210 GPa) similar to that of stainless steel alloy (193 GPa) [2,3].

The high initial strength and fracture toughness of zirconia results from a physical property
of partially stabilized zirconia known as “transformation toughening” [2,3]. Zirconia is a
polymorphic material that has 3 crystal phases: monoclinic (m), tetragonal (t), and cubic (c).
At room temperature, zirconia is in monoclinic phase and transforms into tetragonal phase
at 1170°C, followed by a cubic structure at 2370°C [2]. While cooling, the metastable
tetragonal zirconia is transformed into stable monoclinic zirconia. The tetragonal to
monoclinic (t→m) phase transformation is associated with a large volume expansion (3-5%)
that induces compressive stresses opposing crack opening and acts to increase resistance
to crack propagation [3]. In vitro studies of zirconia specimens demonstrate a flexural
strength of 900 to 1200 MPa and a fracture toughness of 9 to 10 MPa/m 2 [4]. It is a bioinert,
not soluble metal oxide [5] that also exhibits a favorable radioopacity and a low corrosion
potential [1].

Zirconia frameworks can be produced according to two different CAD/CAM techniques. In


soft machining technique, CAD/CAM systems shape pre-sintered blocks, which involves
machining enlarged frameworks in a so-called green state. The enlarged pre-sintered
zirconia frameworks are then sintered in a sintering furnace to their full strength that is
accompanied by shrinkage of the milled framework by 25% to the desired dimensions [1].
In hard machining technique, fully sintered blocks are shaped [1]. The framework
coloration is performed either adding metal oxides to the zirconia powder, or embedding
the frameworks in metal salt solutions after machining [6]. Glazing is created by firing a
small coating of transparent glass onto the surface or by heating the framework up to
glazing temperatures for 1 to 2 minutes to get shiny glass surfaces [7].

Although zirconia has superior mechanical properties, its opaque white color and
insufficient translucency require glassy porcelain veneering on the framework to achieve a
natural appearance and acceptable esthetics [8]. However, cracking or chipping of the
porcelain veneer has been reported to be a major complication of these restorations [9-
12]. The possible causes of porcelain veneer cracking are; differences in coefficient of
thermal expansion (CTE) between framework and porcelain, firing shrinkage of porcelain,
porosities, poor wetting of veneering, flaws on veneering, inadequate framework design to
support veneer porcelain, overloading and fatique [8].

There are several solutions to overcome the veneer cracking problem due to its
multifactorial nature: alternative application of techniques for veneering such as CAD/CAM
produced veneer [13], modification of the firing procedures [14], and modification of the
framework design [15]. Another alternative solution was to use non-veneered zirconia
restorations. The translucency of zirconia was increased and full-contoured, monolithic
zirconia restorations without veneering porcelain have become increasingly popular as a
result of advances in CAD/CAM technology [8,16]. The monolithic zirconia has been used in
posterior region, especially for single crowns, in order to eliminate the veneer cracking
[17,18]. It has been suggested for use in patients with limited interocclusal space because
of its ability to resist high loads with only 0.5 mm occlusal thickness [19]. The technicians
can also prepare monolithic zirconia for all-on-4 prosthesis by using CAD/ CAM. Limmer et
al. [20] presented 1 year results of clinical outcomes of 4 implant supported monolithic
zirconia fixed dental prosthesis, and observed a few complications related to restorations.
They concluded that these kinds of restorations might be a therapeutic option in the
edentulous mandible.

There are 2 types monolithic zirconia materials; opaque and translucent zirconia. Opaque
zirconia offers significantly greater flexural strength and indicated in the posterior regions
of the mouth. Translucent zirconia has more natural esthetic properties. Lava plus high
translucency zirconia (3M ESPE) has a unique shading system that gives laboratories many
options for custom shading and characterization. After milling a porous green-state block,
the laboratory can choose from among 18 dyeing liquids that cover the 16 Vita Classical A1-
D4 shades to achieve custom coloring. The dyeing liquid is applied and then, during the
sintering step, the color ions are incorporated into the zirconia. With greater strength and
improved esthetics, this high translucency zirconia has the potential to be used in either
the posterior or anterior regions of the mouth.

The low temperature degradation (LTD) is an aging phenomenon related to monolithic


zirconia. In the presence of moisture and at low temperatures (150-400°C), slow tetragonal
to monoclinic transformations occur on the surface of zirconia, then progress into the bulk
of the material [21]. The growth of the transformation zone results in severe micro-
cracking, grain pullout and surface roughening that leads to decrease in strength [22]. LTD
was found to intensify for rougher zirconia surfaces; therefore, smooth surfaces are
required to prevent LTD [23].

A definitive cementation protocol for zirconia ceramics has not been validated yet. Both the
conventional and adhesive cementation techniques are feasible. For the adhesive
cementation, different air-blasting protocols associated with chemical primers such as
formulations containing MDP monomers or silane coupling agents are the most
recommended conditioning methods for zirconia restorations, followed by dual-cured resin
cements [24,25].

To date, many articles on monolithic zirconia have been published. However, there is still
little general knowledge with regard to their mechanical behavior and reliability, and the
factors that would contribute to their optimal application performance. Therefore, the
purpose of this article is to give a succinct literature review on the material properties of
monolithic zirconia, to summarize research articles conducted on this subject, and provide
information on this alternative restoration type based on the results of original, full-length,
scientific papers published in journals listed in PubMed.

Materials and Methods


A PubMed search was conducted up to May 2015. The terms of “zirconia” or “zirconium
dioxide” or “yttria-stabilized tetragonal zirconia polycrystals (Y-TZP)” with “monolithic”, “full-
contour”, “solid”, “translucent”, “anatomic-contoured”, “un-veneered”, “non-veneered”, “full-
coverage” were used. The literature search covered all years and focused on publications
that contained dental data regarding in vitro studies, case reports, clinical studies and
reviews. The publications that used veneered zirconia, and the studies that did not use
zirconia material as a superstructure were excluded. Full-text of the articles were obtained
from different sources and the abstracts in English were used which were written in a
different language instead of English.

Results
According to PubMed search, the total number of publications that met the inclusion
criteria for this review was 49. Of these, 28 were laboratory studies, 10 were case reports, 4
were clinical studies, 4 were clinical aspects and techniques, 2 were stress analyses, and 1
was a literature review article on a special subject (wear).

Most of the studies were conducted in vitro [17,18,26-51]. Wear properties was
investigated in 19 articles [17,18,26,28-34,37,41,42,44,46-50], surface roughness in 9
articles [26,28,29,31,43,45,46,48,51], fracture strength in 6 articles [35,38,40,43,49,50],
optical properties and color in 4 articles [7,36,39,50], and marginal fit in 1 article [27]. There
were 2 stress analyses [52,53] and 4 clinical aspects and techniques [54-57]. There was only
1 review article about the wear behavior of monolithic zirconia against enamel [58]. Other
published articles were clinical studies [16,20,59,60] and case reports [61-70].

In vitro studies

Wear: Wear means “loss of material from a surface” [44]. Wear of a material is related to
several factors, such as mechanical contact, surface roughness, grain size, fracture
toughness, occlusal load, temperature, chemical reactions, environment and lubrication
[34]. Surface conditions is one of the most crucial factor, therefore, different kinds of
surface treatments should be applied on the restorative materials in order to prevent
damage of natural antagonist teeth [44].

There are two common surface treatment techniques for monolithic zirconia, such as
polishing (manual/machine) or glazing (glass coating/firing) to improve the esthetic
appearance of the restoration and to obtain smooth surface texture. Diamond points,
rubber wheels and abrasive pastes are used in polishing procedures. Glazing is performed
by firing a thin coating of glass on the surface or by firing the restoration up to
temperature required for glazing [7].

The wear ability of monolithic zirconia was evaluated in 19 studies. (Table 1). According
to Table 1, it can be clearly observed that polished zirconia had the lowest wear on the
antagonists compared to glazed zirconia. This result was attributed to the fact that glazed
zirconia loses the thin glaze after a short period of clinical function, with the result of
appearance of the rough and more abrasive surface of zirconia. It was also stated that
glazed layer is easily removed by chairside occlusal adjustments [47]. Only one study by
Beuer et al. [50] reported higher antagonist wear with a polished zirconia than with a
glazed zirconia. This difference was attributed to polishing techniques that created as
smooth as or smoother than glazed surfaces in other studies. They concluded that results
might be different if other polishing techniques would have been applied on zirconia
surfaces.

The Success of Dental Veneers According To Preparation Design


and Material Type
Yousef Alothman1 and Maryam Saleh Bamasoud2,*
Author information Article notes Copyright and License information Disclaimer
This article has been cited by other articles in PMC.

Abstract
Go to:

Introduction
Since 1930s dental veneers have been used to improve the aesthetic and protection of teeth
(Calamia, 1988) [1], the indications of dental veneers include: 1) discoloured teeth due to many
factors such as tetracycline staining, fluorosis, amelogenesis imperfect, age and others 2)
restoring fractured and worn teeth 3) abnormal tooth morphology 4) correction of minor
malposition 5) Intra-oral repair of fractured crown and bridge facings [2], [3], [4]. Unfavourable
conditions of dental veneers include 1) patients with parafunctional habits such as bruxism 2)
edge to edge relation 3) poor oral hygiene 4) insufficient enamel [5], [6]. Many studies reported
positive clinical outcomes veneers, with a survival rate of 91% in 20 years [7] dental veneers are
considered a predictable aesthetic correction of anterior teeth.
The materials of dental veneers have evolved remarkably, early materials that had been used had
many disadvantages such as the materials needed to be too thick to cover any discolouration,
difficulty to polish which can cause abrasion of the opposing dentition and easy to stain [8], [9].
Researchers and dental material manufacturers have aimed to develop new materials with better
aesthetic characteristics through the years. In 1975 laminate veneers were introduced as a better
material of choice to mask the dentition, the restorations were 1 mm in thickness and were made
from a cross-linked polymeric veneer [10]. The use of laminate veneers resulted in a better
aesthetic outcome and less chair time [11]. The progress of developing new materials reached
porcelain in the 1980s when enamel was etched, and the porcelain surface was treated to
improve the bonding [12], [13].
The desire for more durable aesthetic outcomes did not confine to improve the material type
only; new preparation designs were introduced to the field of dental veneers. There are four
different main designs of teeth preparation commonly mentioned in the literature (Figure 1): 1)
window preparation: in which the incisal edge of the tooth is preserved 2) feather preparation: in
which the incisal edge of the tooth is prepared Bucco-palatable, but the incisal length is not
reduced 3) bevel preparation: in which the incisal edge of the tooth is prepared Bucco-palatable,
and the length of the incisal edge is reduced slightly (0.5-1 mm) 4) incisal overlap preparation: in
which the incisal edge of the tooth is prepared Bucco-palatable, and the length is reduced (about
2 mm), so the veneer is extended to the palatal aspect of the tooth [14], [15], [16], [17].
Figure 1
Showing common veneer preparations a) window b) feather c) bevel d) incisal overlap [17]
Go to:

Influence of preparation design on the survival of dental veneers


Different opinions have been reported about superior preparation design over the others. In fact,
due to the great variety in the materials, preparations designs and luting cement, favourable
approaches to restore teeth with veneers have been controversial.
This review aims to compare the survival rate of dental veneers according to different
preparation designs and different material types. The sub-aim is to reach a favourable
preparation design and material based on scientific evidence.
One important aspect to investigate is the tooth preparation of dental veneers and how it might
affect the fracture resistant of the material and reinforcement of the abutment tooth.
Unfortunately, clinical trials that investigate the survival rate of dental veneers according to
preparation designs are few, the criteria of investigation would include more than one factor
which can affect the outcome of the treatment [16], [18]. In contrast, many in vitro studies have
been conducted to evaluate the influence of different preparations design. Although such studies
do not mimic the actual clinical environments and factors, they can provide criteria and
guidelines for the clinician and further clinical investigations [5]. Table 1 illustrates the results of
multiple in vitro studies regarding the influence of preparation design.

Porcelain Veneers: Problems and


Solutions
Category: Aesthetics Created: Thursday, 01 August 2002 00:00
  Print

 
  Email

Porcelain laminate veneers were first introduced to the dental profession in 1983 by Faunce, 1 Horn,2 and
Calamia.3 Since that time, they have become a popular mode of aesthetic dental treatment. They
are generally more durable than composite resin veneers, and, when properly formed and placed,
are indistinguishable from natural teeth. Light passes through and reflects from a tooth restored
with a porcelain veneer in almost the same way as with an unrestored natural tooth (Figure 1).
Problems occur, however, when the veneer is inappropriately selected as the mode of treatment,
or when some aspects of fabrication and/or placement are less than ideal. The problems most
often encountered with porcelain veneers are: failure to meet the patient’s aesthetic expectations,
fracture or loss of the restoration, or adverse periodontal consequences following placement. 
This article reviews the major causes of these problems and offers suggestions for overcoming
them.

Figure 1.  Porcelain veneers mimic


the aesthetics of natural teeth.
MEETING THE PATIENT’S AESTHETIC EXPECTATIONS
Patients often select porcelain veneers to improve the appearance of their teeth because they feel aesthetically
disfigured. They see their teeth as being the wrong color, the wrong shape, or in the wrong position. They are
looking for a conservative method to correct their problems at a reasonable cost and in a reasonably short time.
The first step toward satisfying a patient is to obtain a clear definition and understanding of that patient’s
personal goals and expectations. The dentist should ask simple, open-ended questions at the initial interview,
such as “How can I help you?” or “What would you like me to do for you to improve the appearance of your
teeth?” Patients should be encouraged to express their feelings about their teeth and how they would like to
change them. Having a patient hold a mirror to view his teeth during this discussion is helpful. The patient and
the dentist must understand the limitations of veneers (in comparison to crowns) and assure themselves that the
patient’s expectations can be met with this treatment modality.
CONTRAINDICATIONS FOR VENEERS
Porcelain veneers are not appropriate for aesthetic restoration of all anterior teeth. Careful selection of the teeth
to receive the veneer is necessary to ensure satisfaction. Long-term success is measured by continued aesthetic
satisfaction, durability without becoming dislodged or fractured, the absence of visible surface or peripheral
staining, and functional harmony with the other teeth. Some of the contraindications for veneers include the
following situations.
Little Enamel Available for Bonding

Figure 2.  After 3 years, the veneer


has fractured from the dentin of
the prepared tooth but remains
bonded to enamel.
Retention can be assured by selecting teeth that, after proper preparation, provide a substrate of enamel with
minimal (if any) dentin exposure. Although materials and methods for bonding resins to dentin are currently
popular, neither the strength nor the durability of the resin-dentin bond compares with the bond of resin to
etched enamel, particularly when stress is applied to the restored teeth. Resin-enamel bonds are strong and
enduring for years, whereas the preponderance of published dental research shows conclusively that over time,
under stress, resin-dentin bonds weaken.4-13 In fact, it is difficult to find any research reports that
indicate that dentin-resin bond strength of any dentin adhesive agent improves or even remains
the same over a period of years. If porcelain veneers are bonded primarily to dentin, even with
the best materials and techniques currently available, they become vulnerable to fracture or loss
over time (Figure 2). The risk is increased significantly when tensile and torquing stresses are
applied to the veneer during normal chewing, or when patients exhibit parafunctional habits such
as bruxism.

Figure 3.  Placing veneers on the


teeth of bulimic patients results in
an incisal margin bonded to
dentin. The risk of incisal fracture
of the veneer is high over time as
stresses are applied to bonded
interface.
Contrary to recommendations made frequently by others writing or lecturing on this subject, I choose not to
place porcelain veneers on any tooth that has more than 30% exposed dentin after preparation. Recognizing the
instability of resin-dentin bonding under stress, I also do not use veneers in cases where the incisal margin
terminates in dentin (such as the typical bulimic patient) (Figure 3). In these cases, ceramo-metal or all-
porcelain crowns, rather than veneers, are more predictably durable.
Substantial Shade Changes Are Desired
Figure 4.  Veneers often appear
unnatural when opaque porcelain
is used to mask very dark teeth.
In order to preserve a substrate of enamel, the thickness of the veneer often must be less than 1 mm. When the
patient’s teeth are very dark colored (eg, severe tetracycline staining), it is impossible to use such a very thin
veneer of porcelain to mimic the translucency and general appearance of an extremely light-colored natural
tooth (Figure 4). Opaque porcelains and/or relatively opaque resin cements must be used to block out the
underlying dark color of the tooth.  Opaque veneers lack translucency and appear artificial. It is generally true
that the greater the degree of lightening of the teeth desired by the patient, the less desirable veneers are to
accomplish that result. All-porcelain crowns are generally thicker than 1 mm, and are more suitable than
veneers when major changes in shade are desired.
Mandibular Teeth
Restoring mandibular anterior teeth with porcelain veneers is much more technically difficult than placing
them on maxillary teeth. This is due to the small size and thin natural enamel of lower teeth, coupled with
inconvenient access and isolation. I do not endorse the use of porcelain veneers to restore mandibular teeth. I
prefer crowns instead. The additional circumferential reduction of the lower anterior teeth required by crowns
offers more opportunity for developing proper embrasure form, anatomical contours, more predictable
retention, and less risk of fracture over time.
Severely Rotated Teeth

Figure 5.  Preparing rotated teeth


for veneers results in a large
extension of porcelain bonded to
exposed dentin. This increases the
risk of fracture of the veneer.
Patients with severely rotated anterior teeth often feel that a veneer will provide them with “instant
orthodontics.” When a labially rotated anterior tooth is prepared to receive a veneer, the result is almost always
removal of all the enamel from the labioverted segment of the tooth. When the preparation is complete, the
veneer must fill a large triangular proximoincisal space with a cantilevered segment of porcelain bonded to an
adjacent island of dentin. This is a very unfavorable situation that increases the probability of fracture. I feel
that “instant orthodontics” should be done using crowns rather than porcelain veneers, except in cases of very
minimal rotation where preparation exposes very little dentin or none at all (Figure 5).
Large Class III Restorations
Some patients requesting veneers present with multiple large, defective class III composite or cement
restorations that will not be covered on the lingual surface by the veneer. The dentist and the patient must
assess the relative expenditure of time, money, and effort involved in replacing these prior to veneering, in
comparison with placing a crown that removes or covers them, eliminating these restorations from any further
consideration.
Bruxism
Figure 6.  Placing veneers on this
bruxing patient’s teeth without
concurrent restoration of the lost
posterior support makes fracture
of the veneers highly likely over
time.
Destructive habits such as bruxism place high stresses on restored teeth. Veneers that lengthen the anterior
teeth are at a very high risk for fracture, unless the posterior teeth are similarly elongated to restore the
patient’s lost vertical dimension. When veneers are placed on the teeth of bruxing patients, protective guards
are often indicated to prevent future fracture (Figure 6).
Based on nearly 17 years of clinical experience with porcelain veneers, which includes many successes and
failures, I limit my use of porcelain veneers (in nonbruxing patients) mainly to maxillary teeth of nearly
normal color, minimal rotation or malalignment, and minimal exposure of dentin and/or old composite
restorations after preparation.
PREPARATION DESIGN
A very important determinant of long-term aesthetic and structural success of the porcelain veneer is the
configuration of the prepared tooth, particularly the location of the proximal margins. When preparing a tooth
for a laminate veneer, the dentist has three options to position the proximal margins: labial to the proximal
contact, halfway through the contact, or lingual to the contact.
When the proximal margin of a porcelain veneer is placed only slightly labial to or halfway through the
proximal contact of the adjacent tooth, access for proper finishing is restricted. There is a high probability of
visible stain forming along this margin sometime in the future. Such stain is difficult, if not impossible,
to remove.

Figure 7. Preparations that do not Figure 8. Stain is evident at the


break the proximal contacts midline because the proximal
complicate fabrication and margin is located slightly labial to
cementation of the veneers, and the contact.
often result in visible interproximal
stain over time.
Placing the proximal margin halfway through the contact makes it impossible to section the stone model
without damaging the margins of the dies. Marginal inaccuracy is almost ensured with this preparation design
(Figures 7 and 8).
However, breaking the contact and placing the proximal margins on the lingual surface of the tooth provides
many benefits:
•Any proximal marginal stain is not visible.
•The margins are more accessible for finishing.
•There is a greater surface area for bonding the veneer.
•The veneer exhibits a greater resistance to labiolingual flexion because of curvature.
•It is easier for the laboratory to separate the dies without damaging the margins.
•There is reduced likelihood of tearing the impression during retrieval.
•There is reduced likelihood of bonding the veneer to the adjacent tooth during cementation.
•It facilitates a more positive positioning of the veneer before and during cementation.

•The veneer can have improved proximal contours and aesthetics (Figure 9).

Figure 9.  Placing the proximal


margins lingual to the contact
provides many advantages over
preparations that maintain the
contacts.
It is not always possible to break both proximal contacts when preparing teeth for veneers. When veneers are
placed on bicuspids and molars with broad contacts, at least one of the proximal margins usually must be
placed labial to the contact to allow seating of the veneer from a buccal or facial direction. Any stain that
might form at the interface over time will be minimally visible.
PROVISIONAL VENEERS
Provisional (temporary) veneers can be the key to ultimate success with porcelain veneers. They maintain the
physical and social comfort of the patient during treatment while permitting both the dentist and the patient to
assess the aesthetic, phonetic, and functional consequences of this type of dental treatment.
The major benefits of provisional veneers are reviewed.
Preview of the Future
High-quality provisional veneers afford the dentist and the patient an excellent opportunity to preview the
future teeth, helping ensure that there will be no unpleasant surprises at the end of treatment. Provisional
restorations that resemble the original disfigured teeth (or do not resemble the final intended product) have no
value in predicting the outcome of treatment. Provisional veneers should be used to predict the future; they
should not be a reflection of the past.
Assessment of Occlusal Function
Aesthetic dental treatment often includes changes in length and/or shape of the teeth. These changes have both
functional and phonetic consequences. A patient may want the incisal length of all the maxillary anterior teeth
to be the same. Restoring the teeth in this manner may create occlusal interferences, particularly in latero-
protrusion. This can lead to discomfort, incisal fracture of the veneers, or destructive wear of the opposing
teeth. The patient often cannot foresee or understand this conflict between aesthetics and function.
Provisional veneers constructed according to the patient’s wishes and expectations will reveal this problem as
the occlusal interferences created by this configuration of the teeth are seen and felt by the patient. This can
help convince the patient that their aesthetic preference is not realistic, or that adjustments of opposing teeth or
even extensive restoration of the posterior teeth to increase the vertical dimension may be necessary if
treatment with veneers is to be successful.
Identification of Consequences
Veneers that close large diastemata and/or enlarge or lengthen the teeth may cause disturbances in speech that
may or may not be accommodated by the patient. This, too, should be assessed with  provisional veneers prior
to commitment of time, energy, and money to permanent restorations that otherwise would need to be
recontoured or replaced.
Accomodation of Patients’  Requests for Major Changes in Shade
Occasionally, patients request treatment that deviates from the usual standards of aesthetics. Perhaps the most
familiar example of this is the desire for teeth that are unnaturally white. To some patients this is beauty,
although the majority of the population may perceive the teeth as fake or artificial. Provisional veneers made
as white as the patient desires give the patient an opportunity to think about this and get reactions from other
people before making a commitment to the final outcome. This process will usually clarify the aesthetic
acceptability of such a request. Often, sending  patients home with “white as snow” provisional veneers is the
only way to convince them to modify their expectations and accept more natural-appearing restorations.
FABRICATING PROVISIONAL VENEERS

A variety of techniques for producing provisional veneers are described in the dental literature.
Many of these articles give greater emphasis to speed, efficiency, utility, and protection from
sensitivity than to their predictive value or their role in preserving and maintaining periodontal
health. Splinted provisional veneers frustrate dental hygiene, making the patient vulnerable to
periodontal disease. In my view, provisional veneers should resemble the final porcelain veneers
as closely as possible. They should be individually bonded to the prepared teeth, permitting the
patient to floss and clean the teeth conveniently. The seal of the provisional veneers to the
prepared teeth should be sufficient to prevent ingress of bacteria or stain, and should protect the
patient from sensitivity. The surface finish should be smooth, conveniently hygienic, and kind to
the periodontal tissues.

Figures 10 and 11.  Individual  


provisional veneers can be made
by pressing composite resin onto
the tooth, curing it, and shaping it
with rotary discs.

Figure 12.  A clear template can be Figure 13.  The template and a
made using a wax-up or a cast of silicone model of the prepared
the patient’s teeth. teeth should fit together precisely.
Figure 14.  Composite resin can be Figure 15.  The clear template can
applied directly to the silicone be used to form a provisional
model without separating agents. veneer resembling the preoperative
model.

Figure 16.  Temporary veneers.


High-quality individual provisional veneers (as well as permanent resin veneers) can be made semidirectly by
pressing microfilled composite resin onto the prepared tooth, curing and removing the prototype veneer, then
shaping it with rotary discs (Figures 10 and 11). I often find it convenient to make provisional veneers
indirectly at chairside using a silicone model of the prepared teeth (Figures 12 through 16). This is done in less
than 5 minutes by injecting die silicone (Mach-2, Parkell; Die-Flex, Danville Materials) into an alginate
impression of the prepared teeth. Composite resin can be applied directly to the silicone model, light cured,
and removed without the need for any separating medium. The prototype veneer is then lifted off, contoured,
and polished with rotary discs. This technique (which can be used to make “permanent” composite resin
veneers as well) is fast, accurate, and involves very little interaction with or cooperation from the patient
during the fabrication process.
If a diagnostic wax-up is available, or if the veneers will closely resemble the shape of the original teeth, a
template can be made preoperatively using clear polyvinylsiloxane (PVS) bite-registration material (Clear-
Bite, Danville Materials; Memosil, Bayer) in a nonretentive tray. After the composite resin is built up on the
silicone model, the clear template can be placed over it, forming a mold that confines and shapes the composite
resin to mimic the diagnostic cast or preoperative contours of the teeth. Excess composite resin can be vented
through holes drilled through the template prior to light curing through the clear template. The mold is then
disassembled and the provisional resin veneers are readily removed and individually finished.
The provisional veneers are attached to the teeth by etching a large central spot of facial enamel followed by
rinsing and drying to reveal the characteristic “frosty” surface. They are bonded to the teeth with any unfilled
resin or flowable composite. When the provisional veneers are removed, all remnant hardened resin must be
removed from the tooth prior to try-in of the porcelain veneers.
In order to avoid unpleasant surprises, unhappy patients, and costly remakes, I recommend delaying the
submission of veneer cases to the laboratory until after the patient has had an opportunity to wear the
provisional veneers for several days and give comments, pro and con, to the dentist. These comments are
relayed to the laboratory in the prescription that is accompanied by a silicone or stone cast of the provisional
veneers bonded to the patient’s teeth.
I have been asked on many occasions to serve as a consultant in legal cases related to dental treatment with
porcelain veneers. This experience indicates clearly that a dentist’s failure to provide quality provisional
veneers often leads to a violation of the patient’s expectations, expensive remakes, and ultimately lawsuits.
Short-cutting the provisionalization stage often validates the lament, “We always seem to have time to do it
over, but never have time to do it right.”
TRY-IN, CEMENTATION, AND FINISHING
Figure 17.  This photo of identical Figure 18.  The same veneers after
veneers illustrates the difference in cementation.
appearance of wet (No. 8) and dry
(No. 9) veneers at try-in.
Previewing the final restoration at the delivery appointment is essential to gaining the patient’s
acceptance. If patients first view their veneers outside the mouth prior to try-in, they often feel
that “what they see is what they’ll get.” However, the appearance of any veneer that is viewed in
one’s hand, on a model, or on a dry tooth will change once it is wet and seated on the prepared
tooth. No useful purpose is served if a patient views or passes judgment on a veneer before it is
placed on his own wet tooth. A veneer that is much too light when viewed dry may prove to be a
perfect shade match when cemented with a clear resin cement. Veneers appear yellow when
viewed on yellow stone models; gray when seen on gray, green, or blue stone models; and too
light when viewed on any tooth-colored dry surface. A veneer that looks to be a good match
when viewed dry will be unacceptably dark after cementation. In all cases, patients will be
disappointed at some point in the procedure if they are allowed to inspect the veneers prior to
intraoral try-in. If the dentist wants to avoid disappointment, he/she should “just say no” to eager
patients who want to see their  veneers too soon (Figures 17 and 18).

Figure 19.  Residual salt formed by Figure 20.  Gingival inflammation


etching the porcelain must be due to ill-fitting veneer.
removed by scrubbing with wet
cotton before bonding the veneer.
Optimal adhesion to the tooth is ensured through proper treatment of both the veneer and the prepared tooth.
The bond of resin to etched porcelain is equivalent to that of resin to etched enamel. It is strong and enduring
over time, under stress. That bond is compromised if, after etching the porcelain, adherent salts remain on the
etched surface. Before cementation, the dentist should dry the veneers and carefully inspect the etched surface
for areas of white, frosty powder. If any remains, it should be removed by scrubbing each veneer with a wet
cotton pellet, followed by rinsing and drying (Figure 19).
Veneers should not irritate and compromise the health of periodontal tissues. Adverse periodontal responses
result from:
•placing gingival margins too deeply into the gingival sulcus.
•laceration of gingival tissues during preparation.
•ill-fitting provisional veneers.
•overcontoured or ill-fitting veneers (Figure 20).
•abraded or underglazed porcelain surfaces that contact the gingival tissues.
•incomplete removal of excess resin cement.
•splinted restorations that interfere with good hygiene.
The gingival tissues can be protected from laceration by retracting them with a relatively large,
nonimpregnated cord prior to final preparation of the gingival margin. Placing the margin 0.5 mm incisal to the
cord will ensure an equi- or subgingival margin that is accessible for finishing and does not violate the biologic
width when the veneer is bonded.
Porcelain veneers do not always meet the dentist’s expectations for precise fit. This can be the result of poor-
quality impressions, inaccurate dies, or the method of fabrication. Not all methods of fabricating porcelain
veneers are equally accurate. Fusing porcelain onto burnished platinum foil has the least potential for precision
fit because there is loss of detail that cannot be reproduced by the platinum foil. It is challenging to make
extremely thin veneers (ie, less than 0.5 mm) with pressed porcelain methods. Fusing porcelain powder
directly onto refractory dies produces the most precise fit, and most easily permits the technician to make very
thin veneers with minimal risk of fracture.
Dentists who place porcelain restorations should become comfortable recontouring and polishing them with
rotary wheels at chairside. Using abrasive wheels that are specifically designed for recontouring and polishing
porcelain, the dentist can customize and improve the veneers for an optimal aesthetic and periodontal outcome.
Summary
Porcelain veneer restorations  require close attention to detail from beginning to end. It is often prudent to go
slowly when working with these cases. Patients receiving them have high expectations that go beyond
considerations of function alone. Many problems are encountered when porcelain veneers are used to improve
the appearance of the teeth. Avoiding and overcoming these problems first requires identification of the causes
of the problems, followed by changes in clinical techniques. Success is the result of careful selection of teeth to
receive veneers; preparing teeth in a manner that optimizes the aesthetic potential of the veneer; employing
techniques that maximize the strength of both the veneer and its adhesive bond to the tooth; utilizing high-
quality provisional veneers; insisting on a precision fit; and paying attention to the details of adhesive bonding
protocols.

Veneer Preparations: When to Break


Interproximal Contact
By Greggory Kinzer on March 8, 2017 | 1 comment   PRINT  
SHARE
 

SHARE

 tweet

 share
In general when preparing teeth for veneers, my goal is to be as conservative as possible and leave everything
in enamel. This would mean leaving the interproximal intact and not preparing through the contact.
Unfortunately, depending on the clinical situation this may not always be possible or recommended.
This article will address the clinical situations that are encountered in practice that I feel necessitate breaking
the interproximal contact.  

Existing interproximal restorations


In a majority of clinical situations, if a Class III restoration is present, I will prepare through the contact in
order to get beyond the composite and place the margin on sound tooth structure. (Figures 1a, b and c) 

Figure 1a Figure 1b Figure 1c


In these situations, I will also typically replace the existing composite during tooth preparation so I know it is
sound. The risk of placing the margin on the existing interproximal restoration is micro-leakage, which may
lead to discoloration/caries.1  
It must be noted that there are times in which I will leave my veneer preparation on composite with the idea of
being more “conservative” with the preparation. Generally, the existing interproximal restorations in these
situations extend much farther onto the palate – so much so that if you wanted to get beyond the restoration,
you would end up preparing the tooth for a partial crown. In these circumstances, it is critical to replace the
existing composite and place the veneer margin through the contact on the palatal surface so that it is
accessible if/when future leakage occurs.

Diastemas/triangular-shaped teeth/open gingival


embrasures
It is imperative to prepare teeth with diastemas or black triangles through the interproximal embrasure and
place the margin at the interproximal-palatal line angle. (Figures 2a, b and c) 

Figure 2a Figure 2b Figure 2c


This allows the technician the ability to close the diastema while maintaining a smooth emergence off of the
preparation. In addition, the tooth preparation in these situations needs to be dropped subgingival for the same
reason of creating a natural subgingival emergence profile that is easily cleansable.

Significant color change is desired


The more dramatic the color difference between the natural tooth and the desired veneer restoration, the more
apt I am to prepare through the interproximal. (Figures 3a, b and c) 

Figure 3a Figure 3b Figure 3c


The main goal here is esthetics. Moving the margin more palatally removes the risk of seeing the junction or
show-through from the tooth.
It is interesting to note that when you combine the above clinical scenarios with the many other reasons why
veneers would typically be placed, you will find that it is not uncommon for the preparation to break the
interproximal contact.
Three Impression Material
Classifications: A Comparison
Category: Dental Materials Created: Tuesday, 28 February 2017 15:59 Written by Sam Simos, DDS
  Print

 
  Email

INTRODUCTION
It is said that a first impression is everything, and this is especially true in the field of dentistry.
The art and science of taking an excellent impression plays a critical role in the restorative
process. Without a stable and accurate replica of the patient’s dentition and surrounding soft-
tissue landmarks, creating an accurate and well-fitting dental prosthesis or lab-fabricated
restoration is virtually impossible. Impression materials provide a straightforward and reliable
method of producing the negative likeness of a patient’s tooth structure and surrounding soft-
tissue landmarks needed to finalize a prosthesis or indirect restoration.
One major problem is that most dentists rely on only one impression material to
address all clinical needs. Clinicians may be better served to stock more than one type of
impression material to accommodate a variety of clinical situations.
This article will discuss the 3 most common classifications of impression materials: polyether
(PE), vinyl polysiloxane (VPS), and a hybrid material called vinyl polyether siloxane (VPES). In
addition, 3 mini case reports will be presented, focusing on the impression to give the clinician
an understanding of the rationale that may be used when choosing the best impression material
from different types of materials in various clinical situations. This article will assist the clinician
in making optimal impression material choices.

Figure 1. Impression technique: Expressing Figure 2. Impression technique: Filling the


light-body wash material around prep and dual-arch tray with a heavy-body impression
teeth. material.
Figure 3. Impression technique: Placing the Figure 4. Impression technique: Final
loaded dual-arch tray into the mouth. impression.
Impression Technique
The impression technique utilized in all cases presented in this article is the heavy-body/light-
body wash technique where the light-body or wash material is placed directly around the prep
and adjacent teeth (Figure 1). Simultaneously, the impression tray is filled with a heavy-body
material (Figure 2), and the heavy-body material is then immediately inserted into the mouth so
that both materials polymerize together (Figure 3). (It should be noted that the working times of
each category of material and set time of each category of material vary greatly. Manufacturer
suggested working times and set times were, and always should be, closely followed.) Once the
set time elapses, the impression was taken out of the mouth and evaluated under 3.5x loupe
magnification (Figure 4).
CASE 1
A new patient presented without an anterior fixed partial denture (FPD) from teeth Nos. 7 to 11
(Figure 5), stating that it had fallen out some time ago. Going without teeth did not bother him
personally; his real motivation for replacement was his wife’s unhappiness with his appearance.
This patient wanted a solution to help mend his domestic disharmony.

CASE 1

Figure 5. Anterior preparation. Figure 6. Polyether (Impregum [3M])


impression showing excellent detail of the
captured margins.
The patient was on blood thinners and, while examining his teeth, it was noted that he produced
a tremendous amount of saliva and gingival bleeding.
The existing margin of the old bridge was subgingival. The desired preparation design required
us to keep the margins of the preparations at the same depth. Care was taken in the impression
phase to minimize bleeding (Astringedent [Ultradent Products]) and saliva flow (controlled with
cotton rolls) around the preparations.
Polyether
The biggest challenge this case posed was that of moisture control. A PE impression material
(Impregum [3M]) was selected for use because of the inherent hydrophilic nature of this
material. From a practical standpoint, the PE’s ability to perform in situations where moisture
(water, saliva, blood, gingival fluids) is a challenge has set it apart from the other impression
materials.  As a hydrophilic material, PE has exceptional flow and wettability, both of which
1

allow for a high degree of detail in the impression and the resulting working model. The
international standard for detail reproduction requires impression materials to possess the ability
to reproduce a line that is 0.02 mm in width or less (a human hair is 0.04 to 0.06 mm), a
measurement that PE consistently surpasses in both wet and dry conditions. 2

PE was the first of the 3 impression materials to be introduced commercially during the 1960s.
The base generally consists of PE macromonomers along with glycol-based plasticizers and
silica fillers. The accelerator supplies the cross-linking sulfonate, which is the catalyst for the
reaction. Because there is no by-product resulting from polymerization, dimensional stability of
the material is retained as the reaction occurs.
1,2

While these traits are certainly laudable, PE impression materials exhibit high rigidity, a
drawback that essentially can result in trouble removing the material from a patient’s mouth.
Additives in the material, namely plasticizers, have been introduced to combat this rigidity issue,
allowing for easier removal from the mouth after polymerization. PE impression materials have
also historically been observed to have a bitter taste, causing more than a few complaints from
patients throughout the years. However, advancements in the chemical makeup have effectively
masked the taste with a minty flavor. Leading brands of PEs include Impregum
Penta and Impregum Soft (3M) and Polyjel NF (Dentsply Sirona).
The outcome in this case was an excellent PE impression (Figure 6) that served as the foundation
for the creation of a long-lasting FPD that this patient’s wife would appreciate for years to come.
CASE 2
A male patient fractured his molar in the upper left quadrant while eating breakfast (Figure 7).
He was a very active and healthy patient and wanted his positive lifestyle reflected in his smile.
He was worried that the tooth would need root canal therapy, or even removal. When he heard
that the tooth could be returned to normal function with a crown, he expressed relief and wanted
to get the restorative work started right away.
Because of the location and prep design (Figure 8) of the tooth, a more advanced VPS
impression material would be used in this case.

CASE 2

Figure 7. Pre-op photo of a fractured Figure 8. Preparation was done and then the
maxillary left molar. soft tissue was troughed circumferentially
(prior to taking the impression) with a diode
laser (Picasso Lite [AMD LASERS]).
Figure 9. Final vinyl polysiloxane impression
(Aquasil Ultra+ Smart Wetting [Dentsply
Sirona]).
Vinyl Polysiloxane
VPS impression materials were introduced commercially in the 1970s, about a decade after PEs.
The base consists of a siloxane co-polymer with silane terminal groups along with coloring
agents and silica fillers. The accelerator supplies a siloxane co-polymer with vinyl terminal
groups along with chloroplatinic acid, which is the catalyst for the reaction.  The chemical
3

process is designated as an addition reaction; this is distinctive from other silicone materials
because there are no by-products associated with the polymerization of these
molecules.  Similar to PEs, this clean reaction accounts for impressive dimensional stability of
1,2,4

the material.
Some studies have indicated the presence of hydrogen gas during polymerization as a result of
side reactions taking place within the base; however, the incorporation of palladium as a
hydrogen absorber/scavenger, along with improved purification techniques and better
proportioning of the materials, has negated this potential problem.2

VPS is naturally hydrophobic, which given the moist environment in which these materials are
used, would seemingly qualify as a severe limitation when compared with PEs. However, the
simple addition of an intrinsic surfactant has helped to overcome the hydrophobic nature of this
material category, providing suitable wettability characteristics. Extrinsic surfactants may also be
used on the patient’s tooth to accomplish the same end.  An example of a direct tooth surfactant
2,4

is B4 Surface Optimizer (Dentsply Sirona Restorative). It is a credit to the long list of strengths
for VPS that deter any negative feedback regarding the proclaimed hydrophobic nature of the
material.
The detail reproduction measurements match or exceed the international standard, and flow
conditions are excellent, if not ideal.  VPS materials have a desirable rigidity that provides a
2

trouble-free removal process from the patient’s mouth. In addition, and unlike PE materials, VPS
materials are naturally tasteless and odorless, though most on the market have been enhanced to
contain some form of an artificial flavoring for the comfort of patients. Some leading brands of
VPS materials include Imprint 4 (3M), EXAFAST (GC America), Take 1
Advanced (Kerr), Virtual XD (Ivoclar Vivadent), Honigum and StatusBlue (DMG
America), AFFINITY VPS (CLINICIAN’S CHOICE Dental Products), SplashMax (DenMat),
and many more.
Recent Advances in VPS Impression Materials
As a recent advancement in the VPS category, Dentsply Sirona has introduced Aquasil Ultra+
Smart Wetting impression material, which was used in case 2.
The tooth preparation was carried out with a subgingival margin. A diode laser (Picasso
Lite [AMD LASERS]) was used to trough the soft-tissue circumferentially around the
preparation. Then, a beautiful and detailed impression was taken of the final preparation (Figure
9).
The idea behind Aquasil Ultra+ Smart Wetting impression material is this: keep the framework
of a high-functioning VPS and make it even better, without transitioning to the category of
VPES hybrid. Aquasil Ultra was reformulated to produce a VPS impression material that does
not compromise in any one area of performance. Specifically, the chemistry was altered so that
the polymerization process would yield polyfunctional bonding between molecules, instead of
the standard bifunctional, or linear, bonding presented with other silicone reactions. The end
result is a matrix of “branched” macromolecules that offer stability in 2 key applications. First
and most obviously, the newly reinforced structural design allows for outstanding dimensional
stability, inherent material strength, and intraoral tear strength. Second, this improved molecular
fortification makes possible the addition of more surfactants to the material without
compromising strength. More surfactants equal better hydrophilicity.
The resulting material selection proved to be an excellent one for the patient presented in case 2.
A worry-free impression for a worry-free restoration.
CASE 3
A female patient fractured her lower incisor at the gumline and wanted to save the tooth.
Endodontic treatment was done and a post-and-core was then placed in the tooth. A
circumferential subgingival margin was placed at a depth that provided a minimum of 2.0 mm of
tooth structure (ferrule effect) to retain and support the crown (Figures 10 and 11). Because of
the depth of the margin, the presence of crevicular fluid presented one of the challenges in
capturing an accurate and detailed impression. In addition, bleeding, although minimal, was
another factor that could have potentially compromised the outcome and therefore played a role
in final impression material selection as well. A VPES was selected for the impression material
of choice in case 3.

CASE 3

Figure 10. Buccal view of the Figure 11. Occlusal view of the


preparation. preparation. Note the deep subgingival
margin placed circumferentially to
achieve an adequate ferrule effect.
Figure 12. Final vinyl polyether
siloxane impression (EXA’lence [GC
America]). Note the clear and detailed
impression of the subgingival margin.
Hybrid (VPES)
VPES impression materials are the most recent subset of dental impression
materials available on the market, having made their commercial debut in the last
few years. The developmental aim is very direct: combine the desired attributes of
both leading categories of impression materials. Simply put, why settle for a PE or
VPS when you can reap the benefits of both?
The chemical composition of a VPES is highlighted by a PE co-polymer that is
cross-linked with an organo-hydrogen polysiloxane and a VPS co-polymer. A
platinum catalyst is then employed to yield the final molecule, a vinyl
siloxanether.3 The resulting material is professed to have the premium qualities of
a VPS (stability, tear strength, elastic recovery) coupled with the natural
hydrophilicity and flowability of a PE.1,4 Theoretically, the hybrid material
represents the ultimate optimization of traits from 2 well-established and reliable
impression materials. In reality, the proposed synergistic effect of combining the
PE and VPS has not been as readily apparent as expected. While studies can
confirm that the final material is adequate to provide dentists with a third
comparable option in today’s market, such a hybrid creation has not yet been
shown to be greater than the regular sum of its 2 parts. However, this is not entirely
surprising, since the era of the VPES is still in its infancy; more clinical studies are
required to make a true comparison of the VPES to its predecessors. For now, the
hybrid’s performance with respect to detail reproduction, wettability, tear strength,
etc, is very much in line with what has been observed in PE and VPS products. The
leading brand in this category of impression materials is EXA’lence (GC America)
and was the material of choice for this case.
While this is an impression for which more could go wrong than right, the material
selection for the impression was a major part of the successful outcome of the
impression itself (Figure 12).
IN SUMMARY
Currently, the 3 most common impression materials are PE, VPS, and VPES
hybrid. Each of these material categories exhibits slight advantages over the others,
but all 3 are dependable candidates in terms of overall quality. The ever-present
competitive streak in the field of dentistry seems to have ensured that no one
product outstrips its competitor. What once may have been glaring disadvantages
of one material have been more or less mitigated by the need for manufacturers to
stay relevant. And so, while the PE, VPS, and hybrid categories differ in their
chemical makeup, the manufacturers have provided clinicians with 3 comparable
options to choose from, depending upon individual clinician preference and the
clinical situation at hand.
Clinicians face daily challenges when preparing for dentures, partials, crowns,
bridges, and implant prostheses that, because of the limitations of the impression
materials in the past, were often insurmountable. Advancement in the chemical
makeup of each impression material category now give the dentist options that can
work well in cases that are routine or in the most challenging of clinical situations.
It is up to the clinician to understand the different characteristics of the various
impression materials and to choose one that is best matched to the specific clinical
situation at hand. When this is accomplished, the most accurate reproduction of the
dentition and surrounding oral landmarks can be achieved, and the outcome will be
the creation of the best possible final prosthesis.

References
1. Burgess JO. Impression material basics. Inside Dentistry. October
2005. dentalaegis.com/id/2005/10/impression-material-basics. Accessed December
6, 2016.
2. Mandikos MN. Polyvinyl siloxane impression materials: an update on clinical
use. Aust Dent J. 1998;43:428-434.
3. Sun M. A Laboratory Evaluation of Detail Reproduction, Contact Angle, and
Tear Strength of Three Elastomeric Impression Materials [master’s thesis].
Bloomington, IN: Indiana University School of Dentistry; 2011.
4. Re D, De Angelis F, Augusti G, et al. Mechanical properties of elastomeric
impression materials: an in vitro comparison. Int J Dent. 2015;2015:428286.

Dr. Simos maintains private practices in Bolingbrook and Ottawa, Ill. He received


his DDS degree at Chicago’s Loyola University and is the founder and president of
the Allstar Smiles Learning Center and client facility (Bolingbrook), where he
teaches postgraduate courses to practicing dentists on cosmetic dentistry,
occlusion, and comprehensive restorative dentistry. He is an internationally
recognized lecturer and leader in cosmetic and restorative dentistry. He can be
reached at cmesmile50@gmail.com, via the website allstarsmiles.com, and on
Twitter @allstarlc1.

REV.CHIM.(Bucharest)♦67♦No.1♦2016 http://www.revistadechimie.ro 123


Applications of Heat-pressed Ceramics for Single Tooth Restorations
1 1 2
LILIANA POROJAN , CRISTINA SAVENCU , SORIN POROJAN *
1
V. Babes University of Medicine and Pharmacy Timisoara, Faculty of Dentistry,
Department of Prostheses Technology, Specialization
Dental Technology, 9 Revolutiei 1989 Blvd, Timisoara, Romania
2
V. Babes University of Medicine and Pharmacy Timisoara, Faculty of
Dentistry, Department of Oral Rehabilitation, Specialization
Dental Technology, 9 Revolutiei 1989 Blvd, Timisoara, Romania
Currently, several methods have been used for fabricating all-ceramic
restorations. Among them, heat-
pressing technique was a well-established method utilizing lost-wax technique.
The aim of the study was
to evaluate the feasibility of application of dental heat pressing technique in
processing dental ceramics and
the marginal fit of obtained single tooth restorations. All the processing steps
were developed to achieve
different single tooth restorations. The staining technique was used. Heat-
pressing technique used in the
present study has been considered as easy processing, less time-consuming and
with optimal marginal fit.
Key words: dental ceramic, heat pressing technique, marginal fit
The availability of different dental ceramic systems
provides solutions for different types of restorative problems
in esthetic dentistry, from conservative to extensive. Many
years ago, all-ceramic restorations were limited for the
anterior region. Today we have the ability to use ceramics
also in the posterior areas and they are sufficiently strong.
Since 1965, it can be seen an increase in strength and
fracture toughness due to an increase in the crystalline
content of the ceramic materials. This is also the period
where dentistry transitioned to advanced technology and
processing methods. Regarding to esthetics and costs,
materials for all-ceramic restorations become even more
attractive. Further, the ability to fabricate all-ceramic
restorations in a monolithic form, rather than layering,
becomes essential for dentists. Because of the large variety
of this kind of ceramics, it is important to know the
materials and their characteristics [1].
However, the demand for metal-free materials with
increased optical properties, which mimic the natural
teeth, has been increasing [2 - 5]. All-ceramic restorations
have been advocated for super esthetic accompanied with
acceptable mechanical properties [6, 7]. Properties of
ceramic materials were closely related to the
microstructure and preparation technique [8, 9]
Ceramics can be divided by their microstructure (i.e.,
amount and type of crystalline phase and glass
composition), processing technique (powder/liquid,
pressed, or machined), and clinical application. To provide
the reader with a better understanding of ceramics, the
authors give a classification based on the microstructure
of ceramics, with the inclusion of how the ceramics are
processed, which affects durability. At a microstructural
level, ceramics can be defined by their composition of
glass-to-crystalline ratio. There can be infinite variability of
materials microstructures; however, they can be divided
into four basic compositional categories with a few
subgroups [10, 11]:
• Category 1: glass-based systems (mainly silica)
• Category 2: glass-based systems (mainly silica) with
fillers, usually crystalline (typically leucite or a different
high-fusing glass)
This category has a large range of glass-cr ystalline ratios
and crystal types, so much so that the authors subdivided
this category into three groups. The difference is varying
* email: porojan_sorin@yahoo.com
amounts of crystal types have either been added to or
grown in the glassy matrix.
Subcategory 2.1: low-to-moderate leucite-containing
feldspathic glass
Subcategory 2.2: high-leucite (approximately 50%)-
containing glass, glass-ceramics
Subcategory 2.3: lithium-disilicate glass-ceramics
- Category 3: cr ystalline-based systems with glass fillers
(mainly alumina)
- Category 4: polycrystalline solids (alumina and
zirconia) [10]
Currently, several methods have been used for
fabricating all-ceramic restorations. Among them, heat-
pressing technique was a well-established method utilizing
lost-wax technique [12]. Heat pressing technology, which
involves the simultaneous application of heat and pressure
to prefabricated ingots in a previously invested mold cavity,
has been used in dentistry for over 40 years to fabricate
single crowns and partial fixed dental prostheses. Ceramics
can be pressed onto a substrate or formed as a monolithic
restoration, depending on their use and the esthetic needs
of the patient [2].
Lithium disilicate glass ceramic system was developed
with good mechanical properties and suitable translucency
for dental restorative applications [13]. As a potential
restorative material, the ability to easily fabricate dental
restorations with commercially available dental heat
pressing equipment was evaluated in different studies. The
properties of heat-pressed mica-based glass ceramic were
r e p or t e d t o b e c l o s e l y r e l a t e d t o t h e h e a t - pr e s s i n g
procedures, such as temperature and holding time [14,
15].
Heat pressing of glass ceramic materials for dental
applications is a proven method of fabricating fixed
prosthodontic restorations. These restorations are
translucent because of the absence of a metal substructure
and thus offer an excellent opportunity for achieving life-
like esthetic restorations. The microstructure consists of
70% lithium disilicate crystals embedded in a glassy
matrix. Extensive research into the mechanical properties
and clinical performance of heat-pressed glass ceramics
has been carried out over the past 2 decades [16-22]. Some
studies show promising results and give good insight into
re-pressed materials. However, studies have only tested a
http://www.revistadechimie.ro REV.CHIM.(Bucharest)♦67♦No.1♦2016
124
single reuse of the material, whereas, in practice, it could
be reused several times, depending on the amount of
material left over. Routine mechanical strength tests
revealed no adverse consequences for lithium disilicate
when it is reused and that dental laboratories may be
routinely and unnecessarily discarding excess material [21,
22].
Lithium disilicate is composed of silica, lithium dioxide,
alumina, potassium oxide, and phosphorous pentoxide,
which is melted together and then cooled. The glass is
heated at specific temperatures to produce crystalline
growth. When optimal crystalline growth has occurred to
maximize the material’s strength, the ceramic is pulverized
into powder. The powder can be pressed into ingots or
processed using other techniques. These processes result
in raw materials that are either CAD/CAM milled or heat-
pressed to reach a strong, monolithic, final restoration.
Lithium disilicate, when used properly as a monolithic, full-
contour restoration, represents a major and significant
change for dentistry [1].
The geometry of tooth preparation for ceramic prosthetic
restorations has been the subject of many debates without
clear evidence that one type of tooth preparation or method
of fabrication provides consistently superior marginal fit.
A well-designed preparation has a smooth and even
margin. Rough, irregular margins substantially reduce the
adaptation of the restoration. The cross-section
configuration of the margin has been the subject of much
analysis and debate. The minimization of marginal gaps is
an important goal in prosthodontics.
Experimetal part
The aim of the study was to evaluate the feasibility of
application of dental heat pressing technique in processing
dental ceramics and to evaluate the marginal fit of obtained
single tooth restorations.
Wax patterns of different single tooth restorations were
prepared and fixed by wax sprues with 3 mm in diameter
and 3 mm length (fig. 1).
The invested mold was transferred to a burnout furnace
and heated from room temperature to 850°C and hold for
60 min to melt down wax. Then the invested mold was
immediately transferred into commercially available
automated dental heat-pressing equipment (Multimat 2
Touch & Press, Degudent, Hanau, Germany) which had
been already preheated up to 700°C. After inserting the
lithium disilicate glass ceramic ingot (Cergo Kiss,
Degudent, Hanau, Germany), which was heated together
with the mold, and an alumina plunger, the heat pressing
procedure was used with a holding time and heat pressing
temperature of 980°C under a pressure of 4.5 bar, about 45
min (fig. 3, 4).
Then the wax patterns were invested with investment
material mixed with corresponding liquid (Cergo fit Speed,
Degudent, Hanau, Germany) (fig. 2).
Fig. 1. Spruing of the wax
patterns
Fig. 2. Invested
mold
Fig. 3. Invested
mold in the heat-
pressing
equipment
Fig. 4. Pressing
process
Cergo Kiss is a pressable ceramic system for highly
aesthetic and biocompatible restorations such as veneers,
inlays, onlays, and crowns. After heat-pressing treatment,
the specimens were carefully devested by sandblasting
with glass powders (50 µm) at a pressure of 4 bar to remove
investment material. Once the objects have become
visible, abrading across the area was continued using
reduced pressure (2 bar) (fig. 5).
The sprues were cut and the restoration to be painted
was finished using diamond burs or stones. The staining
technique was used, that means modeling and pressing
fully anatomic frameworks whose definitive shade is
created by staining with LFC (Low Fusing Ceramic) stains,
incisal stains, body stains, and glazing. The glaze firing was
performed at 800°C, 1 min long (fig. 6, 7).
Traditional tooth preparation margin designs are still
advised by most manufacturers for indirect restorations
and were used for preparations for inlays, onlays and
complete crowns. Two margin designs may be used for
complete ceramic crowns: chamfer, and shoulder. A
chamfer margin is particularly suitable for full crowns. It is
distinct and easily identified, provides space for adequate
bulk of material, although care is needed to avoid leaving a
ledge of unsupported enamel. Shoulder margins always
Fig. 5. Pressed
restorations after
devesting
REV.CHIM.(Bucharest)♦67♦No.1♦2016 http://www.revistadechimie.ro 125
offer space for the crown material. It should form a 90
degree angle with the unprepared tooth surface.
Marginal fit for all restorations were measured on the
cross-sections. Each restoration was filled with the flow
silicone, and then seated on the plaster die. The silicone
from the restoration or the die was embedded in putty
silicone and bucco-lingual and mesio-distal sections were
made. The thickness of the flow silicone was
microscopically measured in 2 points for each section.
Results and discussions
In the last few decades, there have been tremendous
advances in the mechanical properties and methods of
fabrication of ceramic materials. While porcelain-based
materials are still a major component of the market, there
have been moves to replace metal ceramics systems with
all ceramic systems. Advances in bonding techniques have
increased the range and scope for use of ceramics in
dentistry. The new generation of ceramic materials
presents interesting options, both in terms of material
selection and in terms of fabrication techniques. A closer
understanding of the dynamics of the materials with
respect to design of the restoration and the intended use is
required to enable these restorations to perform [11].
Pressed ceramic restorations are fabricated using a
method similar to injection moulding. Monochromatic
porcelain or glass-ceramic ingots are heated to allow the
material to flow under pressure into a mold formed using a
conventional lost-wax technique. The restoration may be
cast to its final contours and subsequently stained and
glazed to provide an esthetic match. Pressable ceramic
systems may be used for inlays, onlays, veneers, and
crowns [10].
The incomplete fit of restorations remains a critical
problem for dentists, leading many researchers to study
this problem. Marginal and internal accuracy of fit is valued
as one of the most important criteria for the clinical quality
and success of single tooth restorations. The means of the
marginal fit measurements were calculated. The values
are in the range of them which have been reported for the
conventional fabrication techniques [23, 24] and are
acceptable for clinical use: 156.83-180.21 µm for inlays,
79.59-107.83 µm for onlays, 59.01-116.92 µm for crowns.
Conclusions
Heat pressed ceramics single tooth restorations
represent a solution of choice in many clinical situation

Introduction
e.max® is a trade name for lithium disilicate (Li2Si2O5), which is a glass
ceramic that is both strong and esthetic. e.max® is so esthetic, in fact, that it
doesn’t require veneering ceramic. Instead, it can be built as a full-contour
monolithic restoration.
It’s for these reasons that e.max® restorations have become extremely
popular in the last several years.
Monolithic restoration cross-section
Pressing Method
The pressing method involves waxing up the restoration, investing and
burning out the wax, and then pressing the superheated lithium disilicate into
the negative space left from the burned-out wax. The original patent for lithium
disilicate stated that pressing was the desirable production method in order to
get the full benefits of the material.
Pressing superheated Lithium disilicate
Milling Method
When we talk about milling, it’s important to point out that there are actually
two different methods of milling: cutting and grinding. Cutting is the more
precise method and uses burrs that are designed to slice away at the material
in a very predictable fashion.
Grinding, on the other hand, uses diamond burrs to wear down the material
into the desired shape. The biggest issues with grinding are the uneven
milling pattern from the diamond particles. Plus, as the diamond burr wears
down, the accuracy of the milling is reduced.
A study of dental CAD/CAM systems which was published in the Journal of
Dentistry reported findings of microcracks with depths of 40-60 microns when
using the grinding method of milling.
Grinding burs can lead to micro-fractures on the restoration

Cutting burs are preferred for milling


Fracture Toughness
One of the benefits of lithium disilicate is its fracture toughness. Fracture
toughness is the ability of a material to resist surface fractures. Since a
surface fracture will continue to propagate through the material, the goal is to
avoid them in the first place. That’s why a higher fracture toughness is
important for the occlusal surface.
According to Ivoclar Vivadent® in a document published in 2010, pressed
lithium disilicate has a fracture toughness of 2.75 MPa and a flexural strength
of 400 MPa. On the other hand, milled disilicate which has a fracture
toughness of 2.25 MPa and a flexural strength of 360 MPa.
Pressed e-max crystals (simulated)

Milled e-max crystals (simulated)


Fracture toughness is based on the size and formation of the crystals. The
reason that milled e.max® has a lower fracture toughness is that the crystals
are smaller and, therefore, less resistant to fracture.
The reason for the difference in crystal size has to do with how and when the
material is crystallized. Lithium disilicate that is created for the purpose of
pressing is fully crystallized at the factory, resulting in long crystals. These
long crystals are what give lithium disilicate its high flexural strength and
fracture toughness.
They are also the reason why lithium disilicate can’t be milled in its fully
crystallized state – it’s too hard and would cause grinding burrs to wear down
too quickly. To make e.max® millable, the lithium disilicate is only partially
crystallized at the factory.
Lithium metasilicate block prepared for milling
Lithium metasilicate is a lot softer than lithium disilicate and is rated at about
130MPA of flexural strength. This allows the product to be milled with
diamond grinding burrs.
After the milling is complete, the lithium metasilicate can go through its final
crystallization cycle. The resulting lithium disilicate is very different than its
pressed counterpart. This is due to the shorter crystals formed when using the
two-step crystallization method.
These short crystals are the reason for the drop in both flexural strength and
fracture toughness when using the milling methods of cutting and grinding.
At O’Brien, it’s our goal to always give our customers the very best product we
can. This is why we only use the pressing method for fabricating our e.max®
restorations.
An Introduction to Millable Dental Materials
Determine which materials are the best fit for your laboratory.
By Chris Brown, BSEE

Before one can make the business decision to invest in CAM milling technology, it is first necessary
to ask what types of materials and substructures do I want to mill? Over a dozen companies
currently manufacture dental mills for sale in the United States. Some are wet mills while others are
dry mills, and in some cases they are both. Is it a 3-axis or 5-axis mill? Is it powered by 110 VAC or
220 VAC? Does it require a vacuum system? How fast is the spindle speed? Which CAM software
comes with the mill? There are so many options and so many questions. So how does one decide
which dental mill is the best choice? The first consideration, above all else, really comes down to the
question, “What do you want to mill?”

Today, it is possible to mill substructures, press-overs for substructures, full-contour restorations,


models, surgical guides, implant abutments, implant bars, and even removables. However, while a
mill may do a fantastic job milling wax for press-overs, it may not be able to mill titanium bars very
well. The materials being milled ultimately dictate certain features needed of the mill.
In the following article, the author will discuss a number of the common materials being milled today
in laboratories and milling centers across the United States. Those laboratories already milling in-
house may discover they can mill additional materials, while others who are considering the
purchase of a mill may discover what type of mill they should be looking to purchase. In general, all
milled dental materials require secondary processes before being delivered to the dentist.

Stock
The raw material milled in a milling machine is referred to as “stock.” Stocks come in a variety of
shapes and sizes. Some stock is in the form of small blocks, as seen in Figure 1; note the mandrel
on the bottom for the mill to hold during milling. Others are in a frame that the mill grasps on to
(Figure 2). Sometimes the stock itself has a lip or edge to be used by the mill for retention (Figure 3).
Some stock may be small enough that only a single unit—also known as a block—can be milled
from it. Other stock comes in 98-mm diameter disks that can vary in height from 8 mm to 25 mm
(Figure 4). In the case of milling stock, size certainly matters. While it is possible to mill a 5-mm
lower-anterior substructure from a disk that is 20-mm tall, it is costly, wasteful, and time-consuming.
A 5-mm zirconia substructure milled out of an 8-mm tall stock can cost half as much and mill 25% to
30% faster than the same unit milled from 20-mm stock. Whenever possible, laboratories and milling
centers try to mill shorter, single-unit cases in shorter stocks and save the taller stock for multi-unit
bridge cases with odd insertion paths or those long single-unit upper anterior cases on implant
abutments.

Post-Processing
Milled materials are generally not ready to be placed in the patient’s mouth immediately after coming
out of the mill. They all require additional processing. Sintering, shading, and polishing are excellent
examples of the majority of post-processing steps. While some materials may be available as pre-
shaded stock, often the shading is only a base shade that requires further characterization, or the
only stock shade available is white, and more in-depth shading is necessary. As the milling process
does not usually leave a particularly smooth finish, polishing is usually necessary before the dentist
can place the restoration. This is the case particularly for materials such as acrylics and resins.
Zirconia and lithium disilicate are examples of materials that are milled in one state and then need to
be fired in a furnace to change their properties. Then they may still require the subsequent layering
of porcelain, staining, and/or polishing prior to placement.

Cold Pressed (CP) Zirconia


CP zirconia is by far the most common material milled by dental laboratories and milling centers
today. It is available from a number of suppliers with varying degrees of strength (900 MPa to 1400
MPa) and translucency. It starts with zirconia powder that is pressed together or chemically bonded
with other compounds. Once in the desired shape, it is pre-sintered by the manufacturer to put it in a
form conducive for milling. In its pre-sintered state, CP zirconia is generally chalk-like and relatively
easy to mill. Sintering is required after milling and changes the zirconia from a soft material to an
extremely strong material. It also shrinks the zirconia a predictable amount, typically around 25%.
Manufacturers of CP zirconia typically provide the shrink factor for each stock so it can be
programmed into the mill’s CAM software. Stock is available in blocks, disks, and frames.

CP zirconia was initially used for copings and substructures. Recent improvements in translucency
and shading have made it increasingly popular for full-contour restorations. Advanced CAD software
tools have created the ability to design full-contour restorations with areas “cut back” for layered
porcelain and improved esthetics. New shading techniques include pre-shaded stock or the dipping
and painting of milled units with a wider assortment of coloring liquids prior to sintering.

Hot Iso-statically Pressed (HIP) Zirconia


HIP zirconia is composed of mainly the same materials as CP zirconia, but it is pressed under high
temperatures, eliminating the need for sintering. HIP zirconia is typically provided in disk form for
milling. The lack of post-mill sintering and the elimination of material shrinkage make this material
ideal for milling large-span bridge frameworks. HIP zirconia does not have as many shading and
translucency options as CP zirconia, so it is often used only as a substructure or framework. The
biggest drawback to HIP zirconia is that it is extremely difficult, time-consuming, and expensive to
mill. However, the dimensional stability for large bridges can make it worth the wait and expense.

Lithium Disilicate
Currently, IPS e.max  CAD (Ivoclar Vivadent, www.ivoclarvivadent.com) is the only lithium disilicate
®

being actively sold in the United States. It is a glass ceramic that is milled from a small block. This
material is milled in a pre-crystalized phase and has a characteristic blue color. Thirteen low-
translucency shades are available. Since lithium disilicate is more like a glass than a metal or resin,
it makes milling challenging; the material is technically “ground” rather than milled.

IPS e.max has a reputation for excellent esthetics and adequate strength (360 MPa) for anterior and
posterior restorations. Cutbacks can be designed in the CAD software, or the restorations can be
manually cut back prior to crystallization for subsequent porcelain application and improved
esthetics. Availability of this material has been restricted to machines that have been approved by
Ivoclar Vivadent to mill this material. A number of other manufacturers have similar materials under
development.

Feldspathic Porcelain
Millable blocks of feldspathic ceramic were originally designed for the CEREC  chairside milling
®

system. CEREC  Blocs (Sirona Dental, www.sirona.com) and Vitablocs  (Vident, www.vident.com)


® ®

offer numerous shades with varying translucencies that are ideal for esthetic restorations. The
relatively low strength (140 MPa) typically limits their use to anterior restorations and areas of low
occlusal impact. Restorations from the pre-shaded blocks can be simply polished after milling, but
they benefit from improved strength and esthetics if they are stained and glazed prior to seating.
These blocks are available to laboratories with Sirona inLab  mills, but are far more commonly used
®

in clinical chairside mills.

Leucite-Reinforced Porcelain
Ivoclar Vivadent offers millable blocks of a leucite-reinforced glass ceramic under the name of IPS
Empress  CAD. Available in 16 different pre-shaded blocks, restorations from this material have
®

excellent esthetics but relatively low strength (160 MPa). Milled restorations should be polished but
can also be stained and glazed for additional characterizations and strength. These blocks are
available to laboratories using the inLab or E4D milling systems (D4D Technologies, www.e4d.com),
but again are far more commonly used in clinical chairside mills.

Composite
3M produces Paradigm™ MZ100 and Lava™ Ultimate (3M Espe, www.3m.com). Paradigm MZ100
is a classic millable indirect composite available in block form. Lava Ultimate is an advanced nano-
ceramic material incorporating nano-sized particles and clusters of silica and zirconia in a resin
matrix. It is available in block form as well as a frame. These materials mill quickly and only need
polishing after milling. Both materials can be further characterized with light-curable stains and do
not require firing in an oven.

Wax
Machinable wax is used just like traditional dental wax for casting metals or pressing ceramics. Wax
suppliers often have different blends available, which are typically differentiated by color. Some
blends tend to be stronger with better handling characteristics but are not easy to modify. Others are
more easily adapted but may not be as durable. Blue, brown/red, and green are the most common
colors available, but there is no standard as to which color is more durable and which is easiest to
modify. Burnout temperature for machinable wax tends to be slightly higher than traditional dental
waxes.

PMMA, Acrylic, Resin


Polymethyl methacrylate (PMMA) is also known as “acrylic” or “resin.” It is a biocompatible material
with moderate strength and is available in a variety of colors, shapes, and sizes. Clear acrylic is
often used to test mill complex cases for fit or design, and tooth-shaded stocks are used for
temporary crowns and bridges. Several companies are now offering computer-designed removable
prosthetics and milling portions of them from blocks or disks of pink acrylic.

Chrome Cobalt
Chrome cobalt is primarily used as a base metal material for copings and frameworks. It is widely
accepted and used in Europe, but is not yet as popular in the United States. It is not a highly
challenging material to mill, but it can be time-consuming and tool life can be relatively short. Post-
processing of chrome cobalt substructures requires de-gassing prior to layering porcelain.

Titanium
Titanium is a lightweight, strong, and biocompatible metal for substructures and full-contour
restorations, as well as implant abutments and bars. It is available in several different grades, based
on purity. Grade 5, also known as Ti-6Al-4V, is a titanium alloy and is the most common grade used
in dental applications. It is a significantly stronger formulation and less prone to flex. Implant
abutments are often milled from block forms, while substructures, full-contour restorations, and bars
are milled from disks. Post-processing for substructures involves sandblasting, an oxidation wait
period, and application of a bond coat prior to applying specially designed porcelain.

Polyurethane
Polyurethane is a rigid polymer. It is a durable yet easy to mill material, making it an excellent choice
for dental models. As with most plastic materials, many shades are available and it can be formed
into nearly any shape. After milling, some cleaning of model parts may be necessary.

Conclusion
A multitude of millable materials are available in all shapes, sizes, and sometimes colors and
shades. Dental laboratories need to determine which materials best suit their individual business
models and/or rely on their milling center partners. All of these materials, except for the model
material, are subject to FDA approval. Due diligence should be taken to ensure approved materials
are used for any restorations being placed in patients’ mouths.

Future materials are likely to involve improvements in strength, esthetics, and durability. Most
materials today can be wet-milled, but not all can be dry-milled. It is difficult to say if these materials
are capable of being dry-milled or if they will have to be wet-milled. All we know for sure is that the
industry continues to evolve at a blistering pace. It should be safe to assume that the new materials
on the horizon will most likely blend the best characteristics of the materials currently in use.

Effect of the shades of background substructures on the overall


color of zirconia-based all-ceramic crowns
Kallaya Suputtamongkol, DDS, PhD,  Chantana Tulapornchai, DDS, MSc, Jatuphol Mamani,
DDS, Wannaporn Kamchatphai, DDS, MSc, and Noparat Thongpun, BSc
Author information Article notes Copyright and License information Disclaimer
This article has been cited by other articles in PMC.

Abstract
Go to:

INTRODUCTION
Fixed partial denture generally consists of two parts, the enduring substructure or coping and the
esthetic veneer overlaid on the coping to match the neighboring tooth color. Different types of
dental alloys have been used as substructures for metal-ceramic restorations. Because of their
high fracture resistance and high reliability, metal-ceramic restorations have shown exceptionally
high success rates even for a long-term treatment.1-3 However, it is more difficult to create a
metal-based restoration with a natural tooth color and translucency when compared to an all-
ceramic prosthesis because a metal substructure does not allow light transmission through a
restoration.4,5 In contrast, various core ceramics have been shown to have varied degree of
translucency, ranging from a very translucent material to an opaque core ceramic.4-6 The varied
translucency of these core ceramics has made it uncomplicated to fabricate a restoration that can
match its neighboring teeth in terms of color and translucency.
The translucency of ceramic materials depends on several factors such as relative refractive
index, wavelengths of the light sources, numbers and sizes of porosity and inclusion
etc.7 Contrast ratio is an optical parameter that is used to represent the degree of translucency of
a material.7,8 Several studies have used this parameter to compare the light transmission
capability or the masking ability of tooth-colored dental restorative materials.4-6,9,10 For
determining the contrast ratio, the ratio of light reflectance of a material over black and white
backgrounds (Yb/Yw) are measured using a spectrophotometer.7,8 While the contrast ratio for a
completely opaque material is 1, a lower contrast ratio represents a more translucent substance.
For tooth enamel and dentin, the contrast ratio was approximately 0.55 at 1 mm thickness.11
As previously mentioned, color match between a dental restoration and the adjacent natural teeth
depends partly on the degree of translucency of dental restorative material. However, the ability
to conceal the discolored abutment tooth or a metal post and core is also a subject of interest for
all-ceramic restorations. The masking or covering ability of materials can be defined as a
measure of the capability of a coating to hide a colored background and the contrast ratio of 0.98
of a covering layer is proposed for the perfect masking ability in industrial production.8 If the
color of an underlying structure can be observed, it would result in the color difference between
the covering material and the target color. The color difference in CIELAB units is given by the
following equation;12
ΔE*ab = [(ΔL*)2 + (Δa*)2 + (Δb*)2]1/2
(1)
When ΔL*, Δa* and Δb* are the differences in lightness, chroma in red-green axis (a*), and
chroma in blue-yellow axis (b*) of the measured colors of two objects.
In dentistry, the color difference or ΔE*ab is used to evaluate the color match between a dental
restoration and the adjacent natural teeth. The perceptible thresholds (ΔE*ab ≈ 2.6-3.7) were set as
guidelines for color matching determination according to the results from few in
vivo studies.13,14 The perceptible thresholds obtained from in vitro studies were lower (ΔE*ab ≈
0.4-1) because of their better viewing and measuring conditions.15,16 However, the color
differences were reported even for the matched natural teeth.17 For a perfect match between
natural upper central incisors, the reported ΔE*ab values ranged from 0.1 to 1.6. For a perfect
match between an upper natural central incisor and a contralateral all-ceramic crown,
ΔE*ab values varied from 0.2 to 2.9 with an average of 1.6.
Zirconium dioxide or zirconia has presently received considerable attention from dental
practitioners because of its high fracture resistance and excellent biocompatibility. Most
zirconia-based core materials obtained from different manufacturers are yttria-stabilized
tetragonal zirconia polycrystals or Y-TZP. Even though they are all polycrystalline materials
with comparable compositions, they could have slightly different microstructures.18 As a result,
Y-TZPs could have different degree of opacity because of an increase or decrease in light
scattering caused from the microstructural variations.19 For example, an increase in the light
scattering inside the bulk material results from an increase in porosities or inclusions or the
discontinuity of refractive indices at the grain boundary.19 In an esthetic viewpoint, a ceramic
material with limited light transmission is not desirable because it does not imitate the optical
characteristics of a natural tooth. On the contrary, a high opacity ceramic material is required
when a restoration is made on an abutment such as a discolored tooth or a metal post and core.
For zirconia-based core materials, they appear to be opaque materials and they could be used to
mask the dark colors of an underlying substructure.3,20 However, there is limited information
about the translucency of zirconia core materials. The objective of this study was to determine
the effect of color of an underlying substructure on the overall color of zirconia all-ceramic
crowns.
Go to:

MATERIALS AND METHODS


The protocol for this study was approved by the Mahidol University Institutional Review Board
(MU-IRB 2008/031.0506). Twenty adult subjects, of good to excellent dental health, were
recruited from the pool of subjects on the waiting list of the Faculty of Dentistry at Mahidol
University in Bangkok, Thailand. All selected subjects had healthy periodontal tissues, were free
of active periodontal disease and caries, and showed no evidence of bruxing. For these 20 adult
subjects, 7 were men and 13 were women. Their ages ranged from 17 to 55 years. Eligible
subjects had at least one posterior endodontically treated tooth opposed by natural dentition in
the maxillary or mandibular arch. This endodontically treated tooth was used as an abutment for
an all-ceramic crown. After obtaining an informed consent to participate in the study, a
maximum of two crowns were placed per patient.
There were 7 premolars and 14 molars prepared for all-ceramic crowns. Color shade of a
contralateral or an adjacent tooth was determined for use as an all-ceramic crown shade. All
abutment teeth were prepared either for a metal cast post and core or a prefabricated fiber post
with a composite core build-up as foundations for the all-ceramic crowns. The following
minimal guidelines for the final tooth preparation were followed; axial tooth reduction of 1 mm,
shoulder or deep chamfer design, occlusal reduction of 1.5-2 mm, and rounded inner
angles/edges/transitions. The ceramic material used in this study was a zirconia-based core
ceramic and a compatible glass-based veneering ceramic (ZENO®, Wieland Dental and Technik
GmbH & Co, Germany, and IPS e. max Ceram, Ivoclar Vivadent AG, Liechtenstein). Twenty
one crowns were made using a layering technique according to the manufacturer's instructions.
These crowns were divided into three groups according to the remaining abutment tooth
structure and the appropriate reconstruction of the core foundation as shown in Table 1.

Table 1
Experimental groups assigned according to the remaining abutment tooth structure and the
appropriate reconstruction of the core foundation

After try-in and adjustment to produce proper anatomical contours and proper occlusion, each
crown was glazed before insertion. The all-ceramic crowns were cemented using resin cement
which had simply one shade (Multilink, Ivoclar Vivadent AG, Liechtenstein). Only one shade
resin cement was used in order to limit the influence of cement shade on the color of all-ceramic
crowns. Examinations of the overall prosthesis, the marginal area, the adjacent gingival tissues,
and the occlusion were evaluated before cementation. The thicknesses of the core and veneering
materials were measured before cementation at the middle 1/3 of the buccal, lingual and occlusal
surfaces of each ceramic crown. Color measurements of all crowns were made using a shade
measuring device (ShadeEye NCC®, Shofu Inc., Kyoto, Japan) before and after cementation
using the CIE-Lab parameters. L* represents the lightness of a material, +a* represents color in
the red axis and -a* indicates color in the green axis. The blue and yellow axes were designated
by -b* and +b*, respectively. This ShadeEye NCC® intraoral shade measuring device has been
used in some previous studies with an acceptable performance.21,22 The color of all-ceramic
crowns was measured using a pulsed xenon Lamp as an optical light source and a vertical light
receiving system. The instrument was calibrated against a standard calibration according to a
manufacturer's recommendation before each color measurement. The contact plastic tip, having a
diameter of 3 mm, was positioned at the middle 1/3 of the buccal surface of each crown, the tip
made an intimate contact with the crown surface during the measurement. After each
measurement, L*, a* and b* value was obtained and used for calculation of color differences
between before try-in, before and after cementation of all-ceramic crowns. A repeated measure
ANOVA was used for a statistical analysis of a color change between before try-in, before and
after cementation of all-ceramic crowns at α=.05.
In order to obtain the optical properties of a zirconia-based material used in this study, 24
zirconia core ceramic (ZENO®, Wieland Dental and Technik GmbH & Co, Germany) were
prepared in a laboratory. These rectangular core specimens (15 mm × 15 mm) with four different
thicknesses (0.4, 0.6, 0.8 and 1.0 mm) were prepared considering a shrinkage during the
sintering process. After sintering, ZirLiner and dentine shade A3 (IPS e. max Ceram,
IvoclarVivadent AG, Liechtenstein) was applied onto the specimens and fired in a furnace
(Programat P100, IvoclarVivadent AG, Liechtenstein) according to the manufacturer's firing
instructions. A lithia-disilicate-based core ceramic (15 mm × 15 mm × 0.8 mm, Empress 2,
Ivoclar Vivadent AG, Liechtenstein) and a base metal alloy (Wiron 99, Bego, Germany) were
also prepared as controls. For a metal-ceramic system, six rectangular specimens, with a
dimension of 15 mm × 15 mm × 0.3 mm were casted using a lost wax technique. For veneering
procedures, lithia-disilicate-based ceramic and metal samples were veneered with dentin
porcelains (IPS Eris, Ivoclar Vivadent AG, Liechtenstein and Vita VMK95, Vita Zahnfabrik,
Germany).
After veneering, all specimens were ground to a final thickness of 1.5 ± 0.1 mm using a milling
machine (Schick Dentalgerate S master 3, Vacalon, USA) and a diamond grinding disc with a
grit size of 80 µm (S327010, Bredent GmbH & Co. KG, Senden, Germany). All specimens were
glazed by applying a thin layer of the glaze paste onto the grinding surface and fired according to
the recommended schedule. After firing, the thickness of each specimen was measured four
times and the mean thickness was calculated prior to color measurement.
The contrast ratios of all specimens were measured before and after veneering, respectively,
using a spectrocolorimeter (ColorFlex, Model 45/0, Hunter Associates Laboratory, Inc., Reston,
VA, USA). All specimens were measured using the 45°/0° geometry with CIE illuminant D65
and 2 degree observer function. Calibration of the machine was made using a black glass and a
white tile as recommended by the manufacturer. Each specimen was placed at the specimen port
with the measuring window of 13 mm in diameter. The spectral reflectance data was obtained in
the range of 400 - 700 nm at 10 nm intervals. Three measurements were made for each specimen
and the mean contrast ratio was calculated. The mean contrast ratios before and after veneering
of each group were calculated and statistical analysis of the data was performed using a
mixed/split-plot design ANOVA (SPANOVA) test at α=.05. The Tukey's multiple comparison
test was used to determine the rank of each group.
Go to:

RESULTS
For all-ceramic crowns, the mean thicknesses of core ceramic at the buccal surface were 0.6, and
0.7 ± 0.1 mm for premolar and molar crowns, respectively. The mean thicknesses at the occlusal
surface of the core ceramic were 0.7 ± 0.1 for premolar, and 0.8 ± 0.2 mm for molar crowns. The
total thicknesses of all-ceramic crowns after veneering were 1.8 ± 0.3 mm for premolars, and 2.0
± 0.3 mm for molars at the buccal surface.
ΔL*, Δa* and Δb* values of all-ceramic crowns cemented on a metal cast post and core are shown
in Fig. 1, Fig. 2 and Fig. 3. L*, a*, and b* values did not show a significant change when the
values obtained before try-in, before and after cementation were compared. However, L* value of
all-ceramic crowns with a metal cast post and core appeared to be decreased (-ΔL1*)
and a* tended to be increased (+Δa1*) when compared between before try-in and before
cementation. In contrast, b* was not affected much by the color of a background substructure. For
the metal cores in these groups, the height of the metal part was approximately 2-3 mm covered
the remaining tooth structure. L* and b* values of molar crowns with prefabricated post and
composite core build-up also did not show a significant change when compared the values
obtained before try-in, before and after cementation. But a* was significantly increased when
compared between before try-in and before cementation (+Δa1*).
Fig. 1
The differences in L* between before try-in and before cementation (ΔL*1), and between before try-in and
after cementation (ΔL*2) of molar crowns with metal cast post and cores, and molar crowns with
prefabricated post and core build-up.

Fig. 2
The differences in a* between before try-in and before cementation (Δa*1), and between before try-in and
after cementation (Δa*2) of molar crowns with metal cast post and cores, and molar crowns with
prefabricated post and core build-up.
Fig. 3
The differences in b* between before try-in and before cementation (Δb*1), and between before try-in and
after cementation (Δb*2) of molar crowns with metal cast post and cores, and molar crowns with
prefabricated post and core build-up.
The color differences or ΔE*ab between the colors of an all-ceramic crown obtained before try-in
and before cementation (ΔE*1), and between before try-in and after cementation (ΔE*2) were
determined using Equation 1. The mean color differences of all-ceramic crowns are shown
in Table 2.

Table 2
The color differences of all-ceramic crowns between before try-in and before cementation (ΔE*1),
and between before try-in and after cementation (ΔE*2)

For in vitro ceramic specimens, the mean contrast ratios before and after veneering of all-
ceramic and metal-ceramic materials are summarized in Table 3. The contrast ratio of zirconia
core specimen was significantly increased from 0.71 to 0.86 as their thickness was increased
from 0.4 to 0.8 mm, and no significant difference was found at the thickness of 0.8 and 1.0 mm.
After veneering to a final thickness of 1.5 mm, their contrast ratios was increased to 0.92 - 0.95.
Compared with a translucent lithia disilicate-based all-ceramic system, zirconia core with a
thickness of 0.4 mm had a level of contrast ratio similar to this system after veneering. The
metal-ceramic system was used as a control group for a completely opaque system, and its
contrast ratio was 1.00.

Table 3
The mean contrast ratios before and after veneering of all-ceramic and metal-ceramic specimens
Same superscript letters mean that no significant differences were found between groups

Go to:

DISCUSSION
The aesthetic values of all-ceramic restorations are essentially based on their translucency. Even
the zirconia-based restorations have significantly higher fracture resistance than other all-
ceramic systems, but their limited light transmission as reported from few previous studies is
their critical disadvantage.6,20 The contrast ratio of 1.00 was reported for one commercially
available zirconia-based core ceramic, whereas the results from another study have shown that a
zirconia-based restoration might not be as opaque as expected. An excellent match between a
veneered zirconia crown and the adjacent natural anterior tooth has been reported.17 However,
the information about the opacity of the zirconia core materials is still limited even though there
are many zirconia core systems that are currently available in the market.
Light transmission of zirconia-based restorations is limited by its composition and
microstructure. Different levels of light transmission were allowed for dissimilar polycrystalline
zirconia-based specimens because of their microstructural dissimilarities.20 These
microstructural dissimilarities such as grain sizes, the amount of porosity or inclusions, and the
extent and orientation of the grain boundary, are the key factors which responded for the changes
in light reflected and transmitted through zirconia materials.19,20 The variation in grain sizes
and porosity of Y-TZP used in dentistry has been reported in few previous studies.18,23
For posterior zirconia-based all-ceramic crowns used in this study, the changes
of L*, a* and b* values were detected after try-in and cemented on the abutment teeth with either
metal post and core or prefabricated post and composite core build-up, even though these
changes were not statistically significant. The color differences (ΔE*1 and ΔE*2) values were
determined from the changes of L*, a* and b* values. Regarding to ΔE*1, these values indicated
that the color of a background substructure could affect the overall color of posterior zirconia-
based all-ceramic crowns. The cement layer would have minor influence on the overall color
because ΔE*2 was comparable to ΔE*1. However, the mean ΔE*1 and ΔE*2 values (1.2-3.1) did not
exceed the clinically acceptable limit (ΔE*ab< 3.7). This result implied that the color modification
observed on all-ceramic crowns in this study was instrumentally detectable, but it would still be
clinically acceptable. In a previous study conducted in both in vivo and in vitro environments, the
contrast ratio could be related to the masking ability of ceramic veneers placed on the discolored
abutment.10 Even the results from that study showed that the masking ability of 1 mm thick
veneer was insufficient to conceal the discolored teeth but the threshold contrast ratio was
determined to indicate the value above which the restoration could mask the discolored
background. In order to produce ΔE*ab value of less than 3.7, the contrast ratio of a material
should be at least 0.93-0.94 as indicated by the results from that previous study.10 In this study,
the contrast ratio of zirconia all-ceramic crowns should be at least 0.945 at the core thickness of
0.6-0.8 mm according to the results from the in vitro investigation (Table 3). Therefore, the
masking ability of the zirconia crowns was acceptable at these clinically relevant thicknesses.
For a thin zirconia coping (0.4 mm) that had comparable opacity as that of a lithia-disilicate-
based ceramic, it would be advisable to use a tooth-colored core materials to prevent the
traceable dark shadow of a restoration.
As mentioned earlier that the changes of L*, a* and b* values were detected after try-in and
cemented on the abutment teeth with either metal post and core or prefabricated post and
composite core build-up, but the changing patterns were not similar. The optical properties of
metal and nonmetallic material are different because of the differences in their atomic
structures.24 For metallic materials, the incident light is absorbed superficially and then reflected
from the surface. Therefore, metals are opaque and highly reflective and the metallic color could
be effectively reflected through overlying translucent materials. For nonmetallic substance, the
occurrence of refraction, and transmission of light at the interface and inside the bulk material is
unavoidable. As a result of refraction phenomenon, the amount of light scattering plays an
important role on the optical properties of nonmetallic materials.
For non-zirconia-based all-ceramic crowns, the effects of the core and cement shades were
investigated in several in vitro studies.25-33 The results from few studies indicated that the core
shade and color of the luting agents had minor influence on the overall color of all-ceramic
restorations, especially when the ceramic thickness was more than 1.5 mm.25-27 On the
contrary, the effect of the core shades and cement layer on the overall color of all-ceramic
materials was significant in other studies as represented by the high ΔE*ab values that exceeded
the acceptable limits.28-33 However, a similar suggestion has been drawn from these studies that
the thickness of non-zirconia-based all-ceramic materials is a vital factor for this effect because
they are translucent materials. With the thickness less than 1.5 mm, the background shade could
be partly detectable through the all-ceramic materials. When the thickness of a ceramic is more
than 1.5 mm, the final color of an all-ceramic crown would not be significantly affected by the
color of a background substructure or cement. The effects of the core and cement shades were
also investigated in anterior zirconia-based all-ceramic crowns.33 The perceptible color
difference caused from the substrate and cement shades was observed in that study.
Because the limited numbers of zirconia crown were observed in this study, the results obtained
from this study were the preliminary information for only one type of a zirconia restorative
material. Another limitation of this study would be a geometric limitation of color measurement.
To compensate the discrepancy between the measuring window and specimen size, the covering
opaque backgrounds were used during the contrast ratio measurement and it could minimize the
edge-loss effect. Future researches in this topic are required to obtain more information that can
be used in choosing and designing of materials for all-ceramic fixed partial dentures.
Go to:

CONCLUSION
No significant differences were observed between the L*, a* and b* values obtained before try-in,
before and after cementation of posterior zirconia crowns cemented either on a metal cast post
and core or a prefabricated post and composite core. However, the color of a background
substructure could affect the overall color of premolar and molar zirconia restorations with
clinically recommended core thickness based on the changes of ΔE*ab in this study.
Go to:

Footnotes

The authors do not have any financial interest in the companies whose materials are included in this
article.

This study was supported by Thailand Research Fund Grant No. MRG5180145.

Вам также может понравиться