You are on page 1of 27

P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA

May 21, 1997 14:19 Annual Reviews AR34-14

Annu. Rev. Mater. Sci. 1997. 27:443–68

CERAMICS IN RESTORATIVE
AND PROSTHETIC DENTISTRY1
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

J. Robert Kelly
Dental and Medical Materials Group, National Institute of Standards and Technology,
Gaithersburg, Maryland 20899; email: robert.kelly@nist.gov

KEY WORDS: dentistry, ceramics, leucite, prostheses, feldspar, clinical failure, fracture, wear,
aesthetics, history

ABSTRACT
This review is intended to provide the ceramic engineer with information about
the history and current use of ceramics in dentistry, contemporary research top-
ics, and potential research agenda. Background material includes intra-oral de-
sign considerations, descriptions of ceramic dental components, and the origin,
composition, and microstructure of current dental ceramics. Attention is paid to
efforts involving net-shape processing, machining as a forming method, and the
analysis of clinical failure. A rationale is presented for the further development
of all-ceramic restorative systems. Current research topics receiving attention
include microstructure/processing/property relationships, clinical failure mecha-
nisms and in vitro testing, wear damage and wear testing, surface treatments, and
microstructural modifications. The status of the field is critically reviewed with an
eye toward future work. Significant improvements seem possible in the clinical
use of ceramics based on engineering solutions derived from the study of clini-
cally failed restorations, on the incorporation of higher levels of “biomimicry” in
new systems, and on the synergistic developments in dental cements and adhesive
dentin bonding.

OVERVIEW
Ceramics were first used successfully in dentistry in 1774, at a time when tooth-
colored replacements were made of ivory, bone, wood, animal teeth, or expen-
sive extracted teeth obtained from human “donors.” None of these biologically
1 The US Government has the right to retain a nonexclusive, royalty-free license in and to any

copyright covering this paper.

443
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

444 KELLY

derived products was stable toward oral flora or corrosive components of food
and saliva. Ceramics immediately solved the problems of stained, decayed,
and terminally malodorous dentures. While initially prized for their hygienic
qualities, early ceramics did not faithfully replicate optical characteristics of
natural teeth and were awkward to fit as prosthetic replacements. Most ad-
vances in dental ceramics over the next 190 years related to improvements in
aesthetics, particularly with respect to translucency and to fabrication methods
in support of dental practice. In the 1960s, dental ceramics were formulated for
routine fusion onto metal substructures, greatly broadening the use of ceramics.
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Dispersion strengthening was also employed, creating a material suitable for


all-ceramic substructures. During the past decade, attention has been paid to
improving structural properties of all-ceramic systems, net-shape processing,
machining as a forming method, and analyzing clinical failure.
One constant historical and current theme in dental ceramics has been the
close collaboration between dentists, dental materials scientists, and ceramic
engineers. Significant future improvement seems possible in the clinical use
of ceramics, based partly on engineering approaches derived from the study
of clinically failed restorations and partly on developments in dental cements
and adhesive dentin bonding. This review begins with background information
on the use and nature of dental ceramics to provide a perspective to readers
unfamiliar with dentistry.

BACKGROUND
Designing with Ceramics for the Intraoral Environment
Ceramic structures are used to replace missing teeth, tooth structure lost to
disease or trauma, and unaesthetic, but otherwise healthy, tooth enamel. Recre-
ating aesthetics and function are the two practical goals of such restorative
treatment. Aesthetics is an obvious attribute of ceramics and, along with dura-
bility, is the primary reason dentists often choose ceramics over other materials.
Function includes chewing, speech, and the maintenance of a healthy harmony
among teeth, supporting bone and soft tissues, and the muscles of mastication.
As discussed below, the goals of aesthetics and function are rarely achieved in
concert with one material—one goal is often compromised for the other. Cur-
rently, dental ceramics having high strength and toughness are not generally
aesthetic.
Certain aspects of the intra-oral environment are listed in Table 1. Dur-
ing mastication, an artificial structure must withstand average cyclic loads of
approximately 60 to 250 N in a moist environment (1, 2). Higher forces can
readily be achieved for brief periods (≈500–800 N) (1, 2) and are more likely to
be encountered during parafunctional behaviors such as clenching and grinding
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 445

Table 1 Intra-oral conditionsa

Applied load ranges


masticationb 6–130 N
maximumc 200–800 N
Cycles/day 1000–1400
Temperature range 65◦ C
pH range 0.5–8.0
aReferences 1, 2.
bForces measured during chewing of different
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

types of foods.
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

cMaximum average biting forces sustained for

a brief period (s).

(bruxism) than during chewing. Actual tooth-to-tooth contact is also more likely
during parafunctional behavior than during mastication, making these contacts
well worth considering for engineering design purposes. Individual areas of
contact between opposing teeth (wear facets) are in the range of 1–4 mm2.
The number of chewing cycles per day is approximately 800–1400. Water is
obviously available to all external surfaces exposed to saliva. It should also be
realized that dentin (inner tooth structure containing extensions of living cells)
is a likely source of water to all internal ceramic surfaces, as well as microleak-
age and diffusion of water through dental cements. Wide temperature and pH
shifts are also intra-oral realities but are probably of only secondary concern
with respect to clinical survival of ceramic restorations.
Ceramic Dental Components
The simplest use of ceramics involves the adhesive bonding of thin shells (0.5–
0.8 mm thick veneers) of material to the visible surfaces of front teeth or the
bonding of small, generally rectangular pieces (inlays) within the chewing
surfaces of back teeth as an alternative to silver-tin-mercury fillings (amalgam).
For veneers and inlays, high tensile strength and fracture toughness of ceramics
appear to be less important criteria for clinical success than aesthetic potential,
fit to the prepared tooth, and an ability to be selectively acid-etched to form
micro-mechanically retentive features for the adherence of dental cements.
These simple structures serve mainly as aesthetic surfaces or to obturate cavities
and are clinically quite successful, although they are often made of the weakest
dental porcelains (3, 4). Denture teeth represent another simple use of ceramics
that demands little beyond aesthetics.
Crowns are more complex prostheses that completely replace all external
tooth structure on single teeth. Crowns are essentially thin-walled (1.0–2.5
mm thick) cylinders closed at the chewing surface. Wall dimensions of crowns
are limited by biologic considerations involving soft tissue health and tooth
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

446 KELLY
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Figure 1 Cross-sectional view of a fixed partial denture fabricated with an internal core material
(metal or ceramic) that is veneered with an aesthetic dental ceramic.

vitality. Crowns can be composed of dental porcelains fused to either metal


or high-strength ceramic substructures or can be composed entirely of an aes-
thetic dental ceramic. Missing teeth are often replaced by integrating two or
more crowns with the replacement tooth in one continuous prosthesis called a
fixed partial denture (FPD). A cross-sectional illustration of an FPD is shown in
Figure 1. More complex prostheses such as crowns and FPDs restore function
as well as aesthetics. Structural properties including tensile strength, fracture
toughness, and resistance to chemically assisted crack growth are generally
thought to be important for determining the clinical durability of all-ceramic
prostheses.

Origin, Composition, and Microstructure of Dental Ceramics


AESTHETIC DENTAL CERAMICS These ceramics are classified as aesthetic based
upon their ability to mimic the translucency and color of vital teeth. Some may
be used to construct the total wall thickness of restorations, while many are
used to form a veneer layer covering either metal or opaque ceramic substruc-
tures. They are processed into dental prostheses by various methods including
viscous sintering of powders, hot-pressing of glassy ingots into lost-wax invest-
ment molds, glass casting and ceramming, and machining by both CAD/CAM
and copy-milling systems (5). Most aesthetic dental ceramics are glass-matrix
composites derived from mined felspar minerals.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 447

Traditional feldspathic dental porcelains evolved in concert with the devel-


opment of European triaxial whiteware formulations (clay-quartz-feldspar).
Around 1774, a Parisian apothecary, Alexis Duchǎteau, with the assistance
of a Parisian dentist, Nicholas Dubois de Chémant, made the first successful
porcelain dentures to replace the stained and malodorous ivory prostheses of
Duchǎteau (6, 7). Dubois de Chémant continually improved porcelain formu-
lations, receiving both French and British patents, and fabricated porcelain
dentures as part of his practice (6, 7). While in England, he obtained supplies
from collaborations with Josiah Wedgwood during the formative years of the
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

famous porcelain manufacturing concern that still bears the Wedgwood name
(6, 7). Improvements in aesthetics seem to have been accomplished with the
minimization or elimination of quartz and clay from most dental compositions.
Essentially, feldspathic glasses became the predominant matrix material, as op-
posed to simply being utilized as a flux, with dental formulations containing
approximately 75–85 mass% feldspar compared with 15–25 mass% in white-
ware formulations (8). Historically, this dental triaxial composition is most
closely associated with Parian porcelain of England, which was made with a
high feldspar content to imitate marble (9, 10).
Many aesthetic dental porcelains today remain essentially alkali-modified
aluminosilicate glasses processed from feldspathic minerals and have relatively
little crystalline filler (8, 11, 12). Although amorphous filler particles of higher
melting temperature glasses are often abundant, crystalline material is limited
to mineral remnants such as feldspathic minerals and quartz, added colorant
and opacifying oxides/spinels, and minor amounts of phases formed at elevated
temperatures, such as mullite and leucite (8, 11, 13, 14).
Exceptions to this include high-expansion porcelains containing purposeful
additions of crystalline leucite (KAlSi2O6) and a micaceous phase glass-ceramic
(discussed below). Leucite forms by incongruent melting in feldspathic glasses
having a K2O content higher than approximately 12 mass% (15, 16). Leucite-
containing porcelains (17–25 mass%) (17) were first developed in the early
1960s for metal-ceramic systems (18, 19), taking advantage of earlier observa-
tions regarding the role of crystalline leucite in the nonlinear thermal expan-
sion of high-potassium feldspar glasses (20, 20a). Leucite raises the composite
coefficient of thermal expansion of the porcelain (to match that of casting al-
loys) owing to both its high thermal expansion coefficient and its martensitic
tetragonal-cubic transition (400–500◦ C) (21).
Powders for firing onto metal substrates are generally physical blends of
high-fusing glass particles containing leucite with softening temperatures of
≈1100◦ C and lower-fusing amorphous particles with softening temperatures
around 875◦ C (18, 22). More recently, leucite has been added at significantly
higher concentrations (40–55 mass%) as a dispersion-strengthening phase that
retains acceptable translucency in the ceramic (for aesthetics) because of the
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

448 KELLY

relatively close match in index of refraction between leucite and feldspathic


glasses (17, 23, 24).
Another type of aesthetic dental ceramic with a substantial volume fraction
of crystalline phase is a glass-ceramic based upon tetrasilicic fluoromica (25).
Glass-ceramics had been envisioned for dental use as early as 1968 (26), and
a commercial material along with its inherently net-shape processing method
was finally introduced into dentistry around 1985 (27). The first version of
this material was cast into phosphate-bonded refractory molds in a dedicated
centrifugal casting machine at temperatures between 1350 and 1365◦ C. Recov-
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

ered glass castings were then cerammed within a special embedment material
at 1070◦ C for 6 h. This material was reported to consist of 55 vol% tetrasili-
cic mica arranged in an interlocking “house of cards” manner within 45 vol%
residual glass (28). Another version of this ceramic, having a higher volume
fraction (70 vol%) of 2 µm diameter and 0.5 µm thick mica platelets, became
available as precerammed blocks for CAD/CAM machining (29).

HIGHER STRENGTH CORE CERAMICS Most aesthetic ceramics are of limited


use where functional demands are high (e.g. crowns on molar teeth and FPDs)
unless they are applied to a metal or high-strength ceramic substructure or core
(as depicted in Figure 1). Although all higher-strength ceramics developed
for dentistry are too opaque to be classified as aesthetic, a far more pleasing
optical result can be achieved with ceramic substructures instead of cast metal.
A certain amount of translucency is desirable in the substructure material.
The concept of the core-veneer ceramic design came with the extension to
dental use of a dispersion-strengthened, alumina-filled glass composition (≈55
vol% Al2O3 in feldspathic glass) (30). Powders of this composition were sin-
tered on thin sheets of platinum foil adapted to replicas of prepared teeth. This
composition and all later higher-strength ceramics discussed below comprise
the core materials upon which tooth-like contours and natural aesthetics were
developed in the standard feldspathic veneering porcelains.
The process of transfer molding was introduced into the dental laboratory
along with a novel net-shape core material (31). The refractory components
of this material included Al2O3 (60 mass%), MgO (9 mass%), and a barium
aluminosilicate glass (13 mass%). In its green state this material also contained
enough silicone (12 mass%) and kaolin clay (4 mass%) to impart sufficient
plasticity for transfer molding at 160◦ C onto an epoxy replica of the prepared
tooth (32). Its net-shape capability, following programmed firing up to 1300◦ C,
was ascribed to the formation of magnesium aluminate spinel (having lower
density than the parent oxides), oxidation of the silicone resin, and expansion
of a closed pore phase (32). This core ceramic was removed from the market
within 3 to 4 years of introduction because, in part, of a high clinical failure
rate for single-tooth crowns.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 449

Slip-casting was recently introduced into dental laboratory use for the fab-
rication of another unique net-shape core material comprised of two three-
dimensionally interpenetrating phases: alumina (≈70–85 mass%) and a lan-
thanum aluminosilicate glass (33). Lanthana additions decrease the viscosity
of silicate melts, which assists infiltration, and raise the index of refraction of the
glass closer to that of alumina, improving the translucency of the core ceramic
(34). The alumina is reportedly provided in a bimodal distribution consisting
of tabular platelets 2 µm thick and up to 12 µm in diameter and fine particles
under 0.3 µm (35). Following slip-casting, the alumina is initially sintered to
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

form a porous body, with neck formation between touching particles probably
involving differential sintering of the colloidal-sized particles as well as surface
diffusion sintering. This porous body is subsequently infiltrated with the glass,
forming a core ceramic containing approximately 5 vol% porosity (35), yet hav-
ing properties (tensile strengths 400–600 MPa; KIC 3.7 MPa · m1/2 − 5.34 MPa ·
m1/2) approaching or exceeding sintered polycrystalline alumina (33, 35, 36).
In a variation of this material, magnesium aluminate spinel (MgAl2O4) is sub-
stituted for alumina, which increases the translucency of the finished ceramic
but does not provide as much strength. Presintered blocks of porous alumina
and spinel are also available for the fabrication of the core structures by copy
milling or CAD/CAM machining prior to glass infiltration (37, 38).
Now, sintered polycrystalline alumina substructures can also be made for sin-
gle crowns with a system that digitizes a replica of the prepared tooth and directs
the machining of an oversized die onto which alumina powder is dry-pressed
(39). This oversized die compensates for expected densification shrinkage, and
sintered parts are reported to have strengths consistent with high-quality dense
alumina (≈600 MPa) (39). As with most advanced ceramics processes, this
system combines some standard dental laboratory methods with CAD/CAM
technology (40).

RATIONALE AND PARADIGMS FOR ALL-CERAMIC


SYSTEM DEVELOPMENT
Concerns over the biocompatability of dental alloys, the aesthetic limitations
of metal-ceramic systems, and the added cost and complexity of metal frame-
work fabrication are the main factors motivating continued research into all-
ceramic restorations. However, clinicians remain concerned about the structural
longevity and predictability of all-ceramic prostheses and the observation that
most dental ceramics become abrasive toward opposing dentition. Therefore,
fracture and wear behavior also have been key focuses in recent developmental
research.
In substructure design, ceramics have simply been substituted for cast metals
in ways that mimic their use. Most clinicians and manufacturers have assumed
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

450 KELLY

that metal-ceramic prostheses were structurally successful simply because of


the support given by the casting to aesthetic dental porcelains. Framework
design principals taught in dental texts invariably include the need for proper
support of porcelain (41, 42). Fracture mechanics concepts involving identified
clinical failure origins, validated stress distributions, possible accumulation of
cyclic fatigue damage, and the probable role of water-assisted fracture were
generally not considered until quite recently.
One far more comprehensive approach is represented by a university-based
development program utilizing the decision tree shown in Figure 2 (43). This
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

approach is an enormous improvement over the majority of investigations that


tend simply to use strength as a their primary criterion. Strength is quite often
ill-defined and can even include rather dubious data from “crunch-the-crown”
tests described below. However, even for the well-founded program represented
in Figure 2, expert opinion necessarily provides the only guidance available
regarding cut-off values for flexural strength and toughness. Values chosen in
Figure 2 have been set slightly higher than the properties of aesthetic ceramics
currently found to be successful as veneers and single crowns for anterior

Figure 2 Decision diagram from a comprehensive university-based dental ceramics development


program. (From Reference 43, with permission from the author.)
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 451

teeth and as inlays in posterior teeth. It is not known whether these strength
and toughness values will serve as screening criteria for ceramics that will be
successful as single posterior crowns.
As yet, there appears to be no expert opinion regarding values for the
strength/environmental tests specified in Figure 2. As discussed below, slow
crack growth exponents have only recently been reported for dental ceram-
ics, and no work has yet attempted to match these values with clinical failure
rate data or to calculate their influence on failure probabilities under intra-oral
conditions. Additionally, it is not clear that appropriate defect populations are
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

being sampled during standard strength testing of dental ceramics. Especially


in the case of multilayer prostheses, there may also be influential structural
factors, including unique stress distributions and failure mechanisms, that are
not addressed by standardized testing of monolithic materials. Structural tests
involving realistic clinical prostheses or validated analogues may be required
for multilayered structures, even including simple bonded inlays that form a
trimaterial system with tooth structure (ceramic-cement-dentin). In general,
however, there are no widely used structural tests that have been validated as
recreating known modes of clinical failure. The program represented in Figure 2
does not address concerns regarding the abrasivity of dental ceramics toward
opposing teeth. In fairness, however, the contact damage mechanisms at work
intra-orally to produce roughened surfaces have not been elucidated or repro-
duced verifiably as part of laboratory testing. Figure 2, then, represents the
most comprehensive development paradigm described to date, but it must also
be recognized as being only a preliminary framework for directing future work.

CURRENT TOPICS IN DENTAL CERAMICS RESEARCH


Overview of Current Research
Dental ceramics research falls into three broad categories: (a) work devoted
to the improved understanding of microstructure/processing/property relation-
ships and the clinical behavior and laboratory testing of current ceramics and
metal-ceramic systems; (b) studies committed to the incremental improvement
of existing ceramics through the application of post-fabrication surface treat-
ments, microstructural modification, and improved processing schemes; and
(c) development of novel materials or the employment of advanced processing
technologies allowing for the flexible manufacturing of dental prostheses from
advanced ceramics. Topics of particular current importance from the first two
categories are covered in some depth below. Most important advanced pro-
cesses and materials under the last category have received attention elsewhere
in this review. It may be far more important to develop a perspective toward
advanced materials and processes with an eye toward future research agenda.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

452 KELLY

Therefore, the final section of this review focuses on future directions from a
critical assessment of current dental ceramics research.

Microstructure/Processing/Property Relationships
LEUCITE-CONTAINING PORCELAINS Leucite is not an equilibrium phase in the
dental ceramics formulated for application to metal substructures, and not
all commercial porcelains appear to be similarly located with respect to the
K2O-Al2O3-SiO2 ternary (44, 45). Quantitative X-ray diffraction analysis of
multiple-fired dental porcelains has revealed that leucite concentrations are
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

definitely altered during repeated firings, with the leucite content of certain
commercial porcelains increasing and others decreasing (46, 47). Leucite dis-
solution with sanidine precipitation, retention of cubic (high) leucite, formation
of leucite, and the precipitation of other feldspathic minerals have all been dis-
cussed in the literature (16, 44–47).
Well-matched thermal expansion behavior (ceramic-metal) is important for
the practical success of many dental laboratory procedures. Although standard
dilatometry data alone is not considered sufficient to predict compatibility (48),
a safe tolerance of as little as 0.6 × 10−6/◦ C in coefficients of thermal expansion
has been proposed, based upon thermal stress calculations (49). Slow cooling
accomplished in a furnace muffle without power has caused 11 to 56 vol%
increases in leucite content of many porcelains (50). Under an isothermal hold
at 750◦ C for 4 to 16 min, conditions that simulated post-soldering (the joining
of two portions of a prothesis already veneered with porcelain), the leucite
content of six commercial porcelains increased 6 to 21% (51). These percentage
increases in leucite content are sufficient to cause substantial alterations in the
coefficients of thermal expansion (52).
Slow cooling and multiple firing of metal-ceramic prostheses can promote
immediate and delayed porcelain cracking (53). This cracking has been at-
tributed to differential thermal stresses that develop owing to differences in
heat transfer rates and overall thermal history. It is conceivable that the alter-
ations in the coefficient of thermal expansion previously discussed might also
influence porcelain cracking during normal dental laboratory procedures. Ad-
ditionally, it is reported that porcelains fired once can be stronger than those
fired multiple times, providing another indicator of an essential compositional
or microstructural change occurring with repeated firing (54).
Microcracking is another microstructural feature that can influence physi-
cal properties in dental porcelains. Microcracking around leucite grains is a
common observation because the cubic-to-tetragonal transformation of leucite
upon cooling involves a substantial volume change (≈1.2 vol%) occurring be-
low the glass transition temperature. In one review of the biaxial strength of five
commercial leucite-reinforced ceramics, it was concluded that ceramics were
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 453

stronger if their leucite crystals were smaller and more uniformly distributed
(55). More recently, evidence for a critical leucite particle radius for sponta-
neous microcracking was reported (56). Heat-pressing of leucite-containing
porcelain into molds via a sprue resulted in a more uniform distribution of
leucite particles and a coincident, and possibly related, increase in strength
(23).
Clinical Failure Mechanisms and In Vitro Testing
of Ceramic Prostheses
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

CLINICAL FAILURE OF SINGLE CROWNS Clinical failure modes have only rarely
been identified, and they have even more rarely been used to guide structural
modeling or in vitro test development in dental research. Fracture surface anal-
ysis of a limited number of all-ceramic crowns (mainly of a micaceous glass-
ceramic) revealed that most clinical failures had initiated from the internal
(or cementation) surface as opposed to the chewing surface (57, 58). This key
finding of cementation surface failure origin was later confirmed in an indepen-
dent study involving a similar number of clinically retrieved micaceous glass-
ceramic specimens (59). Observations of likely processing defects as well as
defects inherent to the microstructure of this dental glass-ceramic are explored
in these limited fractographic studies (57–59).
Such fractographic findings imply that the internal surface of single-tooth
crowns can become the location of the highest tensile stresses and/or accumu-
lated damage during clinical loading. Finite element analyses are becoming
available that endeavor to link their modeling with the mode of failure ac-
tually observed in all ceramic restorations (60–62). Indirect evidence of the
importance of this cementation (or internal surface) on single crowns is also
available from failure rate studies of micaceous glass-ceramic crowns cemented
with acid-base cements (zinc oxide-phosphoric acid) versus crowns that were
etched to create micromechanically retentive features and then cemented with
methacrylate-based resin cements. Crowns that are etched and bonded to resin
cement have much lower failure rates. For example, three-year failure rates for
molar, premolar, and anterior glass-ceramic crowns cemented with acid-base
cements were 35.3, 11.8, and 3.5% respectively (63). In comparison, an overall
failure rate of 1.3% at two years and 2.9% at four years was reported from a
study of 143 anterior and 254 posterior glass-ceramic crowns bonded with resin
cement (64, 65). This result is consistent with the discussed fracture-surface
observations that failures originate from the cementation surface. Etching and
polymer coating of tensile surfaces have been shown to substantially improve
the strength of glasses and dental porcelain (66–68). It is also worth mentioning
that every all-ceramic restorative system currently available treats the cemen-
tation surface rather harshly by requiring the removal of mold or embedment
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

454 KELLY

materials via air-oxide abrasion or grinding or by forming the surface via a


machining process.
IN VITRO TESTING OF SINGLE CROWNS Much of the guidance dentists obtain
regarding new ceramic systems comes from comparative studies that utilize
load-to-failure testing of single crowns. Test protocols almost invariably involve
loading the biting surface with a ball indenter (or equivalently with a flat platen
against a curved incisal edge) until failure. Such testing is intuitively satisfying
because it seems to replicate the loading experienced intra-orally. However,
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

given that teeth contact at wear facets that are often 1–3 mm in diameter, ball
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

indenters produce clinically unrealistic contact pressures. Nothing, however,


seems to have been published validating that this endemic in vitro test re-
creates failure modes seen clinically. In one contrary report, the failure of
50 glass-ceramic cuspid crowns loaded to failure (biting cusp against a flat
compression platen) was shown to have initiated from indentation damage at the
loading site (69). A photomicrograph of the fracture surface from one of these
glass-ceramic cuspids appears in Figure 3. A Weibull plot of data from all 50
crowns reinforced the fractographic finding that all were from the same failure
distribution, most likely involving propagation of a median crack originating
from sharp indentation conditions associated with localized crushing directly
beneath the indenter (69). Damage such as these crowns suffered has not been
reported to be involved in clinical failure (57–59).

Figure 3 Photomicrograph of the gold-coated fracture-surface of a glass-ceramic crown loaded


to failure on the cusp tip. Failure clearly originated from damage at the loading site.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 455

Table 2 Failure origins in all-ceramic fixed partial denturesa

In vitro In vivo

Interface origin 14 7
(core-veneer)
Free surface origin 6 2
(veneer)
aReference 70.
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

CLINICAL FAILURE AND TESTING OF ALL-CERAMIC FPDS Fractographic findings


have been reported for 9 clinically failed and 20 laboratory-tested FPDs that
had utilized the alumina-glass interpenetrating phase composite core ceramic
discussed above (70). All these prostheses were found to have failed from their
connectors (Figure 1), although many of the laboratory-tested FPDs did ex-
hibit partially formed Hertzian cone cracks and localized crushing, as discussed
above for single glass-ceramic crowns. In contrast to apparent crunch-the-crown
test outcomes, the 9 retrieved clinical prostheses qualitatively shared two major
failure mode characteristics with the 20 laboratory specimens: (a) All prosthe-
ses failed from their connectors, and (b) more subtly, 70–78% of the cracks
originated from the core-veneer interface, as seen in Table 2 (70). As discussed
previously, core ceramics usually have high elastic moduli and strengths com-
pared with veneering ceramics. Because core ceramics are used internally in
FPD pontics and connectors (Figure 1), those portions of the prostheses become
laminate structures for which elastic property mismatches, thickness ratios, and
part geometry determine stress distributions and therefore failure behavior(s).
In addition, interfaces between different materials can be the site of unique
defects, boundary phases, and thermal incompatibility stresses.
Finite element and Weibull probability of failure modeling was used to an-
alyze the laboratory test and to replicate the laboratory test data and failure
behavior (70). Three major features of the model were necessary in order to
faithfully reproduce stresses and failure probabilities consistent with the ob-
served failure origins (Table 2) and the laboratory failure data. First, a small
degree of tooth rotation (tipping) was allowed about a dentin node within the
root portion of the model. This feature is very realistic from a clinical view-
point and suggests that significant abutment mobility (such as that encountered
with periodontal disease) may contraindicate the use of all-ceramic prosthe-
ses. Second, the core-veneer interface needed to be a rich source of structural
defects (modeling required a Weibull modulus of 3.6). Scanning electron pho-
tomicrographs provided solid evidence for this model feature. Third, the core
ceramic required a higher elastic modulus and strength than the veneering ce-
ramic. An elastic moduli ratio (core/veneer) of 3.06 was measured as part of the
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

456 KELLY

modeling, and strengths of 540 MPa (core) and 70 MPa (veneer) were found,
which are within the bounds of published strength data. An important finding
of this analysis was that the veneering ceramic and the quality of the veneer-
core interface control failure of the all-ceramic FPD connector (70). Related
modeling and physical testing demonstrated that layered bend-bars served as
poor test analogues for evaluating FPD connector improvements and that sig-
nificant load-bearing improvements could be derived from use of an improved
veneering material in lieu of a stronger core ceramic (71–73).
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

Wear Damage and Wear Testing


by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Although wear of the ceramic restoration itself is not a critical clinical problem,
wear of enamel opposing ceramics can be significant and is considered to be one
major contraindication to the broader use of ceramics (74–79). Dental profes-
sionals have been inclined to believe that ceramic hardness is a good predictor
of potential abrasivity, as is generally the case for many metals and ceramics
tested against abrasive papers or with loose abrasive particles (80). However,
it is becoming increasingly clear that such conditions may not faithfully model
intra-oral circumstances where the character, size, and shape of abrasive fea-
tures (other factors known to influence wear) also depend upon microstructural
elements of the ceramic, its fracture toughness, and the mechanism of contact
damage (80–82). Wear of enamel pins tested against rotating ceramic discs
was not found to correlate with Knoop hardness (KHN) for five dental ceram-
ics having KHN ranging from 379 to 443 (83). Similar wear studies involving
two much harder ceramics (KHNs = 709 and 1040) revealed extremely low
enamel wear compared with traditional dental-veneering ceramics (KHN '
560), significantly extending the hardness range over which factors other than
hardness appear to predominate in determining enamel wear (84, 85).
The amount and character of damage (number, size, and shape of abrasive
features) developing on dental ceramics during contact appear to determine
enamel wear (84, 85). In general, contact damage characteristics are known
to be a function of the fracture toughness of ceramics, microstructural scale
(grains, filler phases, pores), and local property variations in its microstructure
(80). Two examples from dental ceramics serve to illustrate the influence of
microstructural features and their size on the wear of enamel. First, a mica-
ceous phase glass-ceramic was discovered to be substantially more abrasive
against enamel if a unique microstructure on its external surfaces remained in-
tact, as is normally done in clinical practice (86). This outer skin layer forms on
external surfaces that are in contact with a magnesium-rich embedment mate-
rial during ceramming and contains needle-like crystals of enstatite (MgSiO3)
oriented perpendicularly to the surface, and it is also the site of considerable
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 457

microporosity (87, 88). Second, for feldspathic compositions, a fine microstruc-


ture porcelain for CAD/CAM-fabricated restorations caused appreciably less
wear than a similar porcelain based upon traditionally sized porcelain powder
(≈25–50 µm) when machined inlays cemented in extracted teeth were tested
under mastication conditions against unrestored teeth (89).
An understanding of clinical surface damage is basic to designing relevant
laboratory wear studies and interpreting wear data and is only beginning to be
fundamentally approached. Wear testing is normally quite complex, since mul-
tiple processes may occur simultaneously in service environments. The oral
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

environment provides a large number of wear variables thought to be influen-


tial (77) but not necessarily well modeled during laboratory testing. Review of
dental literature reveals the existence of several types of wear machines; contact
pressures that vary a thousandfold; loading that has been identified as contin-
uous or intermittent; lubricants of differing chemistry; and a variety of flow
volumes (74–76, 79, 83, 89, 90). Any of these variables can alter the mecha-
nism of damage produced (91–94). That very different damage mechanisms are
being studied for similar ceramics is evident from the SEM photomicrographs
of DeLong et al (76) versus Krejci et al (89). Only quite rarely have investiga-
tors attempted to identify actual contact damage mechanisms operating either
under laboratory conditions (95–97) or intra-orally (82). In the examination of
a limited number of retrieved clinical specimens (2 months to 22 years of ser-
vice), oriented microcracks were reported that may have resulted from sliding
Hertzian contact damage and a secondary type of damage was found associated
with clusters of leucite grains (82).

Chemical and Thermal Surface Treatments


CHEMICAL STRENGTHENING Potassium-for-sodium ion exchange strengthen-
ing has been successfully applied to feldspathic dental ceramics (98–102).
However, the glass phase of a micaceous glass-ceramic was not amenable to
ion exchange with potassium salts, as judged by indentation fracture toughness
(103). One enhanced approach involved exchanging rubidium for potassium
because the ionic volume of rubidium exceeds that of potassium, and many
dental porcelains contain more potassium than sodium (104). In this same in-
vestigation, a two-step exchange process was utilized: Lithium was exchanged
for sodium above the strain point, and then potassium was exchanged for lithium
below the strain point. Flexural strengths were higher for ceramics given the
two-step treatment or the rubidium exchange than for the standard potassium-
for-sodium treatment.
Several practical concerns exist that limit the use of chemical strengthening
in dentistry. Dental prostheses commonly require adjustment by the dentist
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

458 KELLY

prior to being cemented. Adjustments can involve physical removal of ceramic


by diamond grinding and air-oxide abrasion, as well as aesthetic adjustments
involving colorant glazes that are sintered well above glass transition temper-
atures of dental porcelains. Chemical strengthening involves only a thin layer
of ceramic, and removal of only 16 to 18 µm of ceramic by air-oxide abra-
sion has been shown to eliminate the strengthening induced by one commercial
potassium-for-sodium product (105), although treatment via an experimental
dual ion treatment related to the one discussed above did survive air abrasion
(106). While it is an advantage that the ion-exchange process does not alter
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

aesthetics, dentists have no visual clues or other method of validating that this
treatment was performed. Residual compressive stresses would be annealed
out during any subsequent glaze firing. There is as yet no data regarding either
the effect of etching treated units (commonly performed to create microme-
chanical retention for bonding resins) or the treatment of ceramics that have
already been etched. Although prostheses could be treated in dental offices just
before cementation, ion exchange protocols can require well over an hour to
accomplish, making such sequencing unrealistic for both the patient and the
doctor.

THERMAL TEMPERING Advantages of thermal versus chemical tempering in-


clude the fact that thermal compressive stress profiles can extend deeper than
chemically derived stresses and that tempering can be accomplished more sim-
ply. Thermal tempering with compressed air or immersion in silicone oil has
been successfully applied to simple shapes such as discs of metal-ceramic porce-
lain (107, 108). This strengthening of dental porcelain appears to result from
an inhibition of crack initiation rather than crack propagation (109). Forced-
air tempering is more effective than ion-exchange strengthening alone or a
combination of tempering and ion-exchange strengthening (110). The thermal
tempering effect for one porcelain survived grinding to a depth of 150 µm
(111), consistent with the greater treatment depth commonly associated with
thermal versus chemical stress profiles.
One potential disadvantage of thermal versus chemical tempering might be
the difficulty in controlling cooling rates (and hence the effect) that may be exac-
erbated for objects having complex shapes, such as dental prostheses. However,
a three-dimensional viscoelastic stress analysis of a thermally tempered metal-
ceramic crown indicated that relatively uniform surface compressive stresses
were calculated irrespective of its complex shape and cooling behavior, al-
though high-tensile stresses may form within the opaque porcelain adjacent to
the metal (112). As discussed for chemical tempering, the same caveats apply
with respect to the timing of the tempering step and adjustments to improve fit
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 459

or aesthetics, although the thermal strengthening effect may be less sensitive to


air-oxide abrasion, grinding, or etching.

Microstructural Modifications
LEUCITE-CONTAINING CERAMICS Leucite remains the most widely employed
crystalline filler in dental ceramics, initially used to increase the composite
thermal expansion coefficient and more recently as a strengthening phase (as
discussed above in Background). Leucite is known to increase the fracture
toughness of dental porcelains over that of pure glasses (113), but leucite grains
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

(or particles containing leucite) are also associated with considerable residual
thermal stress and microcracking within and around leucite grains due to the
cubic-tetragonal transition that occurs during cooling (114). Recently, work
has begun to assess the influence that the residual stresses and microcrack-
ing associated with tetragonal leucite have on physical properties of dental
ceramics. Substitution of K+ by Cs+ ions in leucite has been found to sup-
press the cubic-tetragonal transition, thereby stabilizing cubic leucite down to
room temperature (115). Cubic leucite stabilization was achieved by the addi-
tion of varying amounts of pollucite (CsAlSi2O6) to a leucite-containing den-
tal ceramic (115). It was concluded from this study that the cubic-tetragonal
transition during cooling was necessary to achieve high flexural strengths in
leucite-containing porcelains.
Leucite-containing feldspathic glasses have been produced by hydrolysis of
tetraethoxysilane in the presence of varying amounts of Al3+ and K+, followed
by calcination and sintering at temperatures of 1100, 1200, and 1300◦ C (116).
These treatments yielded amorphous materials or composites of glass-kanasite-
leucite or glass-leucite. Strength, density, and thermal expansion were found
to increase with increasing leucite content. Similar work varied the ratio of K+
to Na+, producing decreasingly lower concentrations of leucite with increasing
sodium content (117). It was suggested that such a chemical route might provide
feldspathic glass composites containing submicrometer leucite filler particles
(116).
Interesting crystallization behavior, noted as the result of ion exchange treat-
ment of feldspathic powders (118), may provide another route to dental ceramics
containing submicrometer leucite (119). Ion exchange treatment in RbNO3 of
potassium aluminosilicate glass powders promoted the crystallization of cubic
leucite. The glass system used, normally resistant to leucite formation, yielded
between 19 and 43 vol% crystallinity, with average particle sizes ranging be-
tween 0.6 and 1.18 µm.
In a microstructural modification driven by ion exchange, external surfaces
exposed to NaI or NaCl salt baths were enriched in sodium, shifting the chemical
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

460 KELLY

composition toward soda feldspar and away from the leucite phase field (120).
Such treatment led to a decreased leucite concentration and a stabilization of
cubic leucite in the treated layer, both of which reduced the thermal expansion
coefficient relative to untreated internal ceramic. This surface treatment resulted
in a 75% increase in flexural strength.

NON-LEUCITE CRYSTALLINE FILLED GLASS SYSTEMS Alumina-filled glass sys-


tems and dental ceramics based upon interpenetrating phases of alumina and
infiltration glass have been utilized in commercial products, as discussed in
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

the Background section. The challenge with using crystalline fillers for elastic
modulus or strength improvements is that concentrations sufficient for a notable
effect are usually coupled with high opacity and fusion temperatures. While a
limited amount of filler experimentation continues, alumina and leucite remain
the only crystalline inclusions to achieve widespread use.
One approach to decreasing the amount of filler required utilizes tetragonal
zirconia in an effort to capitalize on the transformation toughening phenomenon
exhibited in polycrystalline composites (121). Modest increases in toughness
and strength were achieved at 30 mass% of yttria-stabilized zirconia, and evi-
dence was presented to indicate that tetragonal to monoclinic transformation did
occur during grinding and therefore might be operative in the filled-glass com-
posite. Thermal shock resistance was dramatically improved with the addition
of tetragonal zirconia.
More recently, alumina and zirconia were added to glass formulations based
upon the commercial micaceous glass-ceramic composition discussed in the
Background section (122). Aluminum was found to incorporate into the glass
phase and improved physical properties of the cerammed material. Specimens
containing zirconia additions had properties generally worse than the control
glass-ceramic, possibly related to microstructural alterations in the crystalline
phase.
Magnesium oxide has been used as the basis for one core ceramic reported
to have strengths equivalent to alumina-filled core materials but which is dis-
tinguished by having thermal expansion characteristics parallel to those of the
leucite-containing porcelains used for metal veneering (123). This MgO-based
core is apparently slightly porous, and strengths can be improved with the
infiltration of a low viscosity glaze.
Glass-ceramic compositions based upon nepheline (NaAlSiO4) precipitation
have been investigated for dental use (124). These glass-ceramics had high
enough thermal expansion coefficients and sufficient residual glass phase to
allow powders of the cerammed material to be fired onto a dental alloy. Ce-
ramming dramatically decreased their solubility in 1 N HCl compared with
parent glass compositions. Lithia-based glass-ceramics (125, 126), canasite
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 461

glass-ceramics (127), and β-quartz glass-ceramics (128) have also received at-
tention for dental use. Structures fabricated from a β-quartz glass-ceramic were
developed for insertion into tooth preparations being filled with composite resin
materials in order to ameliorate the negative results caused by polymerization
shrinkage. This novel use of a dental ceramic is receiving increasing interest,
especially in Europe.

RESEARCH AGENDA FOR THE FUTURE


Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

Traditional thinking in dental ceramics research has centered on a desire for


by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

higher strengths (or more recently toughness), which has resulted both in a
search for materials novel to dentistry and in the incremental improvement of
existing ceramics. Recently, this approach has been significantly supplemented
by the increasing availability and sophistication of machining processes (numer-
ically controlled and copy-milling) that promise to make a wider variety of ma-
terials available for dental use by removing processing constraints (129, 129a).
It is likely, however, that broadened use of ceramics in dentistry will stem from
more than the development (or cooption) of an advanced material.
Some key structural lessons probably remain to be learned about designing
with brittle materials for oral function. As can be appreciated from the data in
Table 3 (130, 131), properties of dental ceramics are already equal or superior to
enamel or dentin (taken separately). Why this enamel-dentin couple is so highly
successful has not been fully elucidated or, more importantly, well-replicated

Table 3 Strength of dental ceramics and tooth structure

Tensile strength
(MPa)

Ceramics
Alumina/glass interpenetrating
phase composite (IPC) 400–600 (35, 36, 130, 131)a
Magnesium aluminate
spinell/glass IPC 290–380 (36, 131)
Leucite-filled feldspathic glass
40–55 mass% 130–180 (36, 130, 131)
17–25 mass% 60–70 (36, 130)
Micaceous phase glass-ceramic
Dental laboratory processed 115–125 (36, 130)
CAD/CAM processed 140–220 (29, 131)
Tooth Structure
Dentin 232–305 (2)
Enamel 261–288 (2)
aNumbers in parentheses are references.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

462 KELLY

in ceramic systems. Strength alone seems too narrow a standard. Neither the
anisotropic fracture toughness of both enamel and dentin (132, 133) nor the ex-
tremely tough enamel-dentin interface (134) have been well modeled in dental
ceramic systems but probably play important roles in the functional success
of teeth. Whether such biomimetics are possible within traditional ceramics
engineering (or in combination with polymer science) remains to be seen. The
elastic modulus mismatch between enamel and dentin is not targeted as an engi-
neering design. These observations, taken together with the improved clinical
performance of bonded ceramics (64, 65) and recent failure and modeling results
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

from multilayer structures (61, 70), lead to the conclusion that much remains to
be learned regarding structural design with multilayer systems, the role of mi-
crostructures in optimizing resistance to failure, and contact damage. All these
factors may alter the way ceramics are used and teeth are prepared. Recent
work regarding the role of “ductile” microstructures in optimizing resistance to
failure from contact damage (135) may be directly applicable, as may work on
laminates tailored to optimize wear and fracture resistance (136, 137).
Stress corrosion (slow crack growth) exponents remain another underappre-
ciated materials design criterion. Mastication obviously involves low cyclic
loads in a wet environment. Based on relatively well-accepted formulations for
the stress corrosion exponent (138) and relationships between static and cyclic
loading, it can be appreciated that strength differences from standard laboratory
testing can be easily overwhelmed by small differences in resistance to chem-
ically assisted crack growth. Although an awareness of chemically assisted
crack growth in dental ceramics has existed for some time (139, 140), very lim-
ited attention has been paid to formal fracture mechanics measurements. Slow
crack growth exponents (n) reported for three dental ceramics (all leucite-filled
feldspathic glasses) ranged from n = 26–31 (141, 142). No work yet attempts
to link slow crack growth behavior to clinical failure rates. Hydrophobic silane
treatments have been shown to significantly increase the strength of feldspathic
dental porcelains (143).
Very little effort has been spent examining clinical failure (or wear) mecha-
nisms from a fundamental perspective. Such information would be invaluable
in developing tests to replicate clinical behavior and to help validate (or restrict
boundary conditions) for mathematical modeling. This lack of substantive link-
age between laboratory results and clinical behavior leaves a number of quite
basic structural and material design questions unanswered today. For example,
is the success of metal-ceramic crowns simply due to the flexural rigidity of the
cast metal substructure? Alternatively, does this successful structure depend on
unique stress distributions, moisture barrier effects, or crack bridging that can
be replicated in other materials? How thin can the metal coping be and where
must it be used within the substructure? How does one evaluate different coping
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 463

designs and materials, including powder metallurgy and various foil products?
Can an appropriate cement substitute for the metal coping if well bonded to the
ceramic and dentin? Why do glass-ceramic crowns bonded with methacrylate-
based cements have such improved failure rates versus crowns cemented with
zinc oxide/phosphoric acid cements? These are all questions of real interest
that deserve far more fundamental thought than they have generally received.
Answers to these and related questions probably await the development and val-
idation of structural tests and mathematical models dependent upon the further
elucidation of clinical failure mechanisms.
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Visit the Annual Reviews home page at


http://www.annurev.org.

Literature Cited

1. Craig RG. 1985. Restorative Dental Ma- als: Properties and Selection. Chicago:
terials, St. Louis, MO: Mosby. pp. 60–63 Quintessence. pp. 397–418
2. Anusavice KJ. 1996. Phillips’ Science of 13. Hodson JT. 1959. Phase composition of
Dental Materials. 10th ed. Philadelphia, crown and inlay porcelains. J. Dent. Res.
PA: Saunders. pp. 65–66 38:483–90
3. Mörmann WH, Krejci I. 1991. Clinical 14. Mysen BO. 1988. Structure and Proper-
and SEM evaluation of CEREC inlays ties of Silicate Melts. Amsterdam: Else-
after 5 years in situ. In Int. Symp. Com- vier. p. 99
puter Restorations: State of the Art of 15. Hermansson L, Carlsson R. 1978. On
the CEREC Method, ed. WH Mörmann, the crystallization of the glassy phases
pp. 25–32. Chicago: Quintessence in whitewares. Trans. Br. Ceram. Soc.
4. Rucker LM, Richter W, MacEntee M, 77(2):32–35
Richardson A. 1990. Porcelain and resin 16. Rouf MA, Hermansson L, Carlsson
veneers clinically evaluated: 2-year re- R. 1978. Crystallization of glasses in
sults. J. Am. Dent. Assoc. 121:594–96 the primary phase field of leucite in
5. Giordano RA. 1996. Dental ceramic the potassium oxide-aluminum oxide-
restorative systems. Compendium 17: silicon dioxide system. Trans. Br. Ce-
779–802 ram. Soc. 77(2):36–39
6. Ring ME. 1985. Dentistry, an Illus- 17. Piché PW, O’Brien WJ, Groh CL,
trated History, pp. 160–81, 193–211. Boenke KM. 1994. Leucite content of
New York: Abrams selected dental porcelains. J. Biomed.
7. Jones DW. 1985. Development of dental Mater. Res. 28:603–9
ceramics. Dent. Clin. North Am. 29:621– 18. Weinstein M, Katz S, Weinstein AB.
44 1962. U.S. Patent No. 3,052,982
8. Southan DE. 1975. Dental porcelain. In 19. Weinstein M, Weinstein AB. 1962. U.S.
Scientific Aspects of Dental Materials, Patent No. 3,052,983
ed. JA von Fraunhofer, pp. 277–305. 20. Orlowski HJ. 1944. Dental porcelain.
London: Butterworths Ohio State Univ. Eng. Exper. Stat.
9. Norton FH. 1978. Fine Ceramics. Tech- Bull.118:1–34
nology and Applications, pp. 17, 181. 20a. Orlowski HJ, Koenig CJ. 1941. Thermal
Malabar, FL: Krieger expansion of silicate fluxes in the crys-
10. Kingery DW, Vandiver PB. 1986. Ce- talline and glassy states. J. Am. Ceram.
ramic Masterpieces. Art, Structure, and Soc. 24:80–84
Technology, pp. 322. New York: Free 21. Mackert JR, Butts MB, Fairhurst CW.
Press 1986. The effect of the leucite transfor-
11. Lehman ML, Isard JO. 1969. X-ray mation on dental porcelain expansion.
diffraction analysis of dental porcelain. Dent. Mater. 2:32–36
J. Dent. Res. 48:543–45 22. Barreiro MM, Riesgo O, Vincente EE.
12. O’Brien WJ. 1989. Dental Materi- 1989. Phase identification in dental
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

464 KELLY

porcelains for ceramo-metallic restora- strength alumina crowns and fixed


tions. Dent. Mater. 5:51–57 partial dentures generated by copy-
23. Dong JK, Luthy H, Wohlwend A, milling technology. Quint. Dent. Tech-
Schärer P. 1992. Heat-pressed ceram- nol. 18:31–38
ics: technology and strength. Int. J. 38. Bindl A, Mörmann WH. 1996. Clinical
Prosthodont. 5:9–16 and technical aspects of the CEREC In-
24. Shareef MY, Van Noort R, Messer PF, Ceram crown. In CAD/CIM in Aesthetic
Piddock V. 1994. The effect of mi- Dentistry, ed. WH Mörmann, pp. 441–
crostructural features on the biaxial flex- 62. Chicago: Quintessence
ural strength of leucite reinforced glass- 39. Andersson M, Oden A. 1993. A new all-
ceramics. J. Mater. Sci. Mater. Med. ceramic crown. A dense-sintered, high
5:113–18 purity alumina coping with porcelain.
25. Grossman DG. 1972. Machinable glass- Acta Odont. Scand. 51:59–64
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

ceramics based upon tetrasilicic mica. J. 40. Heganbarth EA. 1996. Procera alu-
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Am. Ceram. Soc. 55:446–49 minum oxide ceramics: a new way to


26. MacCulloch WT. 1968. Advances in achieve stability, precision, and aesthet-
dental ceramics. Br. Dent. J. 124:361– ics in all-ceramic restorations. Quint.
65 Dent. Technol. 20:21–34
27. Grossman DG. 1985. Cast glass- 41. Rosentiel SF, Land MF, Fujimoto
ceramics. Dent. Clin. North Am. 29:719– J. 1995. The metal-ceramic crown
23 preparation. In Contemporary Fixed
28. Grossman DG. 1985. Processing a den- Prosthodontics. St. Louis, MO: Mosby-
tal ceramic by casting methods. In Proc. Year Book. pp. 180–93
Conf. Recent Developments in Dental 42. McLean JW. 1980. The Science and
Ceramics, ed. WJ O’Brien, RG Craig, Art of Dental Ceramics. Volume II:
pp. 19–40. Westerville, OH: Am. Ceram. Bridge Design and Laboratory Pro-
Soc. cedures in Dental Ceramics. Chicago:
29. Grossman DG. 1991. Structure and Quintessence. pp. 192, 343–46
physical properties of Dicor/MGC glass- 43. Anusavice KJ. 1995. Development and
ceramic. In Int. Symp. Computer testing of ceramics for dental restora-
Restorations: State of the Art of the tions. In Bioceramics: Materials and
CEREC Method, ed. WH Mörmann, pp. Applications, ed. G Fischman, A Clare,
103–115. Chicago: Quintessence L Hench, 48:101–24. Westerville, OH:
30. McLean JW, Hughs TH. 1965. The re- Am. Ceram. Soc.
inforcement of dental porcelain with ce- 44. Mackert JR, Butts MB, Morena R,
ramic oxides. Br. Dent. J. 119:251–67 Fairhurst CW. 1986. Phase changes in a
31. Sozi RB, Riley EJ. 1985. Shrink-free ce- leucite-containing dental porcelain frit.
ramic. Dent. Clin. North Am. 29:707–17 J. Am. Ceram. Soc. 69:C69–72
32. Starling LB. 1985. Transfer molded “all 45. Barreiro MM, Vincente EE. 1993. Ki-
ceramic crowns”: the Cerestore system. netics of isothermal phase transforma-
In Proc. Conf. Recent Developments in tions in a dental porcelain. J. Mater. Sci.
Dental Ceramics: Ceramic Engineering Mater. Med. 4:431–36
and Science, ed. WJ O’Brien, RG Craig, 46. Fairhurst CW, Anusavice KJ, Hashinger
6(1–2):41–56. Columbus, OH: Am. Ce- DT, Ringle RD, Twiggs SW. 1980
ram. Soc. Thermal expansion of dental alloys
33. Clause H. 1990. Vita In-Ceram, a new and porcelains. J. Biomed. Mater. Res.
system for producing aluminum oxide 4:435–46
crown and bridge substructures (Ger- 47. Mackert JR, Evans AL. 1991. Quan-
man). Quint. Zahntech. 16:35–46 titative X-ray diffraction determination
34. Bansal NP, Doremus RH. 1986. Hand- of leucite thermal instability in dental
book of Glass Properties. Orlando, FL: porcelain. J. Am. Ceram. Soc. 74:450–
Academic. pp. 305, 604 53
35. Hornberger H, Marquis PM. 1995. Me- 48. International Standards Organization,
chanical properties and microstructure ISO/TC 106 (Dentistry). 1991. Specifi-
of In-Ceram, a ceramic-glass compos- cation ISO 9693, Dental porcelain fused
ite for dental crowns. Glastech. Ber. Sci. to metal restorations. Geneva, Switzer-
Technol. 68:188–94 land
36. Seghi RR, Sorensen JA. 1995. Relative 49. Nielsen JP, Tuccillo JJ. 1972. Calcu-
flexural strength of six new ceramic ma- lation of interfacial stresses in dental
terials. Int. J. Prosthodont. 8:239–46 porcelain bonded to gold alloy substrate.
37. McLaren EA, Sorensen JA. 1995. High- J. Dent. Res. 51:1043–47
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 465

50. Mackert JR, Evans AL. 1991. Effect of 64. Malament KA, Grossman DG. 1990.
cooling rate on leucite volume fraction in Clinical application of bonded Dicor
dental porcelains. J. Dent. Res. 70:137– crowns: a two year report. J. Dent. Res.
39 69(IADR Abstr.) 299:A1523
51. Mackert JR, Twiggs SW, Evans AL. 65. Malament KA, Grossman DG. 1992.
1995. Isothermal anneal effect on leucite Bonded vs non-bonded Dicor crowns. J.
content in dental porcelains. J. Dent. Dent Res. 71(AADR Abstr.) 321:A1720
Res. 74:1259–65 66. Phillips CJ. 1965. The strength and
52. Mackert JR, Butts MB, Fairhurst CW. weakness of brittle materials. Am. Sci.
1986. The effect of leucite transforma- 53:20–51
tion on dental porcelain expansion. Dent. 67. Ritter JE, Lin MR. 1991. Effect of poly-
Mater. 2:32–36 mer coatings on the strength of indented
53. Anusavice KJ, Gray AE. 1989. Influence soda-lime glass. Glass Technol. 32:51–4
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

of framework design, contraction mis- 68. Rosenstiel SF, Gupta PK, Van der Sluys
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

match, and thermal history on porcelain RA, Zimmerman MH. 1993. Strength
checking in fixed partial dentures. Dent. of a dental glass-ceramic after surface
Mater. 5:58–63 coating. Dent. Mater. 9:274–79
54. Fairhurst CW, Lockwood PE, Ringle 69. Harvey CK, Kelly JR. 1996. Contact
RD, Thompson WO. 1992. The effect of damage as a failure mode during in vitro
glaze on porcelain strength. Dent. Mater. testing. J. Prosthodont. 5:95–100
8:203–7 70. Kelly JR, Tesk JA, Sorensen JA. 1995.
55. Shareef MY, Van Noort R, Messer PF, Failure of all-ceramic fixed partial den-
Piddock V. 1994. The effect of mi- tures in vitro and in vivo: analysis and
crostructural features on the biaxial flex- modeling. J. Dent. Res. 74:1253–58
ure strength of leucite reinforced glass- 71. Kelly JR, Tesk JA. 1995. Clinically
ceramics. J. Mater. Sci. Mater. Med. relevant all-ceramic FPD mechanical
5:113–18 test analogs. J. Dent. Res. 74(Ab-
56. Mackert JR, Russell CM, Evans- str.):159:A1184
Williams AL. 1996. Evidence of a crit- 72. Grant GT, Kelly JR. 1995. Fabrication
ical leucite particle radius for microc- and failure behavior of an all-ceramic
racking in dental porcelains. J. Dent. FPD test analog. J. Dent. Res. 74(Ab-
Res. 75(IADR Abstr.):126:A866 str.):160:A1185
57. Kelly JR, Campbell SD, Bowen HK. 73. Fahncke CR, Kelly JR, Antonucci JM.
1989. Fracture-surface analysis of dental 1997. Influence of veneering material
ceramics. J. Prosthet. Dent. 62:536–41 on failure of In-Ceram FPD analogs. J.
58. Kelly JR, Giordano R, Pober R, Cima Dent. Res. 76(IADR Abstr.) 311:A2378
MJ. 1990. Fracture surface analysis 74. Mahalik JA, Knap FJ, Weiter EJ. 1971.
of dental ceramics: clinically-failed Occlusal wear in prosthodontics. J. Am.
restorations. Int. J. Prosthodont. 3:430– Dent. Assoc. 82:154–59
40 75. Monasky GE. 1971. Studies on the wear
59. Thompson JY, Anusavice KJ, Morris H. of porcelain, enamel and gold. J. Pros-
1994. Fracture surface characterization thet. Dent. 25:299–306
of clinically failed all-ceramic crowns. 76. DeLong R, Douglas WH, Sakaguchi RL,
J. Dent. Res. 73:1824–32 Pintado MR. 1986. The wear of dental
60. Anusavice KJ, Hojjatie B. 1992. Ten- porcelain in an artificial mouth. Dent.
sile stresses in glass-ceramic crowns: ef- Mater. 2:214–19
fect of flaws and cement voids. Int. J. 77. Ekfeldt A, Ilo G. 1988. Occlusal con-
Prosthodont. 5:351–58 tact wear of prosthodontic materials:
61. Morris DR, Kelly JR. 1995. Failure loads an in vivo study. Acta Odontol. Scand.
of bonded ceramics influenced by hydro- 46:159–69
static compressive stresses. J. Dent. Res. 78. Wiley MG. 1989. Effects of porcelain
74(Abstr.):220:A666 on occluding surfaces of restored teeth.
62. Ellert DO, Kelly JR. 1997. All-ceramic J. Prosthet. Dent. 61:133–37
crown failure as a function of occlusal 79. Hacker CH, Wagner WC, Razzoog ME.
contact location. J. Dent. Res. 76(IADR 1996. An in vitro investigation of the
Abstr.) 392:A3030 wear of enamel on porcelain and gold
63. Moffa JP. 1988. Clinical evaluation of in saliva. J. Prothet. Dent. 75:14–17
dental restorative materials—final re- 80. Larsen-Basse J. 1994. Abrasive wear of
port. Interagency Agreement No. 1YO1- ceramics. In Friction and Wear of Ce-
DE40001–05, San Francisco: Letter- ramics, ed. S Jahanmir, pp. 99–115. New
man Army Inst. Res. York: Dekker
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

466 KELLY

81. Kelly JR, Nishimura I, Campbell SD. Monitoring, Materials, Synthetic Lubri-
1995. Ceramics in dentistry: historical cants and Applications, ed. ER Booser,
roots and current perspectives. J. Pros- pp. 103–20. Boca Raton, FL: CRC Press
thet. Dent. 75:18–32 95. Craig RG, Powers JM. 1976. Wear of
82. Gentile MJ, Kelly JR. 1996. Intra-oral dental tissues and materials. Int. J. Dent.
wear mechanisms on dental ceramics. J. 26:121–33
Dent. Res. 75(Abstr.):50:A264 96. Powers JM, Bayne SC. 1992. Friction
83. Seghi RR, Rosenstiel SF, Bauer P. 1991. and wear of dental materials. In Friction
Abrasion of human enamel by different Lubrication and Wear Technology, ASM
dental ceramics in vitro. J. Dent. Res. Handbook, ed. PJ Blau, 18:665–81. Ma-
70:221–25 terials Park, OH: ASM Int.
84. Reeves N, Gore K, Meiers JC, Kelly JR. 97. Jahanmir S, Dong X. 1995. Wear mech-
1993. Enamel wear against In-Ceram anism of a dental glass-ceramic. Wear
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

and Vitadur-N with various finishes. J. 181–83:821–25


by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

Dent. Res. 72(IADR Abstr.) 187:665 98. Southan DE. 1970. Strengthening mod-
85. George LA, Eichmiller FC. 1996. Wear ern dental porcelain by ion exchange.
of enamel against glass-ceramic, porce- Aust. Dent. J. 15:507–10
lain and amalgam. In Challenges in 99. Dunn B, Levy MN, Reisbeck MH. 1977.
Ceramic Product Development, Manu- Improving the fracture resistance of den-
facture, and Commercialization, ed. A tal ceramics. J. Dent. Res. 56:1209–13
Gosh, R Barks R Rusin, 66:181–88. 100. Seghi RR, Crispin BC, Mito W. 1990.
Westerville, OH: Am. Ceram. Soc. The effect of ion exchange on the flex-
86. Jacobi R, Shillingburg HT, Duncanson ural strength of dental porcelains. Int. J.
MG. 1991. A comparison of the abra- Prosthodont. 3:130–34
siveness of six ceramic surfaces and 101. Piddock V, Qualtrough AJE, Brough
gold. J. Prosthet. Dent. 66:303–9 I. 1991. An investigation of an ion
87. Campbell SD, Kelly JR. 1989. The in- strengthening paste for dental porce-
fluence of surface preparation on the lains. Int. J. Prosthodont. 4:132–37
strength and surface microstructure of a 102. Giordano RA, Campbell S, Pober R.
cast dental ceramic. Int. J. Prosthodont. 1994. Flexural strength of feldspathic
2:459–66 porcelain treated with ion exchange,
88. Denry IL, Rosenstiel SF. 1993. Flex- overglaze, and polishing. J. Prosthet.
ural strength and fracture toughness Dent. 71:468–72
of Dicorglass-ceramic after embedment 103. Seghi RR, Denry I, Brajevic F. 1992. Ef-
modification. J. Dent. Res. 72:572–76 fects of ion exchange on hardness and
89. Krejci I, Lutz F, Reimer M. 1994. Wear fracture toughness of dental ceramics.
of CAD/CAM ceramic inlays: restora- Int. J. Prosthodont. 5:309–14
tions, opposing cusps, and luting ce- 104. Denry IL, Rosenstiel SF, Holloway JA,
ments. Quint. Int. 25:199–207 Niemiec MS. 1993. Enhanced chemi-
90. DeLong R, Pintado MR, Douglass WH. cal strengthening of feldspathic dental
1992. The wear of enamel opposing porcelain. J. Dent. Res. 72:1429–33
shaded ceramic restorative materials: an 105. Osborn GD, Denry IL, Rosenstiel SF,
in vitro study. J. Prosthet. Dent. 68:42– Holloway JA. 1994. Effect of abrasion
48 on the flexural strength of ion-exchanged
91. Evans AG, Marshall DB. 1981. Wear porcelain. J. Dent. Res. 73(IADR Abstr.)
mechanisms in ceramics. In Fundamen- 320:A1745
tals of Friction and Wear of Materials. 106. Holloway JA, Denry IL, Rosenstiel
ed. DA Rigney, pp. 439–452. Metals SF. 1994. Strengthening of felds-
Park, OH: Am. Soc. Met. pathic porcelain by one-step dual ion-
92. Lawn BR, Wiederhorn SE, Roberts DE. exchange. J. Dent. Res. 73(IADR Abstr.)
1984. Effect of sliding friction forces on 320:A1746
the strength of brittle materials. J. Mater. 107. Anusavice KJ, Hojjatie B. 1991. Ef-
Sci. 19:2561–69 fect of thermal tempering on strength
93. Fischer TE, Anderson MP, Jahnmir S. and crack propagation behavior of felds-
1989. Influence of fracture toughness on pathic porcelains. J. Dent. Res. 70:1009–
the wear resistance of yttria-doped zirco- 13
nium oxide. J. Am. Ceram. Soc. 72:252– 108. Hojjatie B, Anusavice KJ. 1993. Ef-
57 fects of initial temperature and temper-
94. Jahanmir S, Fischer TE. 1994. Friction ing medium on thermal tempering of
and wear of ceramics. In CRC Handbook dental porcelains. J. Dent. Res. 72:566–
of Lubrication and Tribology, Vol. III: 71
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

CERAMICS IN PROSTHETIC DENTISTRY 467

109. Anusavice KJ, Gray A, Shen C. 1991. 122. Tzeng JM, Duh JG, Chung KH, Chan
Influence of initial flaw size on crack CC. 1993. Al2O3- and ZrO2-modified
growth in air-tempered porcelain. J. dental glass ceramics. J. Mater. Sci.
Dent. Res. 70:131–36 28:6127–35
110. Anusavice KJ, Shen C, Vermost B, 123. O’Brien WJ. 1985. Magnesia ceramic
Chow B. 1992. Strengthening of porce- jacket crowns. Dent. Clin. North Am.
lain by ion exchange subsequent to ther- 29:719–23
mal tempering. Dent. Mater. 8:149–52 124. Wang M-C, Wu N-C, Hon M-H. 1994.
111. Hojjatie B, Anusavice KJ, Hsiao LI. Preparation of nepheline glass-ceramics
1992. Effect of abrasion on strength of and their application as dental porcelain.
thermally tempered dental ceramics. J. Mater. Chem. Phys. 37:370–75
Dent. Res. 71(AADR Abstr.) 208:A817 125. Anusavice KJ, Zhang N-Z, Moorhead
112. DeHoff PH, Vontivillu SB, Anusavice
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

JE. 1993. Influence of colorants on crys-


KJ, Wang Z. 1994. Viscoelastic analy- tallization and properties of lithia-based
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

sis of PFM crowns using finite element glass ceramics. J. Dent. Res. 72(IADR
methods. J. Dent. Res. 73(IADR Abstr.) Abstr.) 262:A1271
369:A2141 126. Parsell D. 1993. The optimization of
113. Morena R, Lockwood PE, Fairhurst CW. glass-ceramic heat treatment schedules.
1986. Fracture toughness of commercial PhD thesis. Univ. Florida, Gainesville.
dental porcelains. Dent. Mater. 2:58–62 200 pp.
114. Mackert JR. 1988. Effect of thermally 127. Anusavice KJ, Zhang N-Z. 1996. Effect
induced changes on porcelain-metal of nucleation temperature on toughness
compatibility. In Proc. 4th Int. Symp. on and flexural strength of canasite glass-
Ceramics, ed. JD Preston, pp. 53–64. ceramic. J. Dent. Res. 75(IADR Abstr.):
Chicago: Quintessence 67
115. Denry IL, Rosenstiel SF. 1995. Leucite 128. George LA, Eichmiller FC, Bowen RL.
stabilization in feldspathic dental porce- 1992. An intrinsically colored micro-
lain. In Bioceramics: Materials and Ap- crystalline glass-ceramic for use in den-
plications II. Ceramic Transactions, ed. tal restorations. Bull. Am. Ceram. Soc.
RP Rusin, GS Fischman, 63:115–21. 71:1073–76
Westerville, OH: Am. Ceram. Soc. 129. Mörmann WH, Bindl A. 1996. The new
116. Yoon CK, Rasmussen ST, O’Brien WJ, creativity in ceramic restorations: dental
Tien T-Y. 1994. Properties of leucite- CAD-CIM. Quint. Int. 27:821–28
glass composites prepared by a co- 129a. McLaren EA, Sorensen JA. 1994. Fabri-
precipitation process. J. Mater. Res. cation of conservative ceramic restora-
9:2285–89 tions using copy milling technology.
117. Sheu T-S, O’Brien WJ, Rasmussen ST, Quint. Dent. Technol. 17:19–25
Tien T-Y. 1994. Mechanical proper- 130. Lehner CR, Shärer P. 1989. All-ceramic
ties and thermal expansion behavior in crowns. Curr. Opin. Dent. Curr. Sci.
leucite containing materials. J. Mater. Prosthodont. Endodont. pp. 45–52
Sci. 29:125–28 131. Luthy H. 1996. Strength and toughness
118. Denry IL, Rosenstiel SF. 1996. Effect of dental ceramics. In CAD/CIM in Aes-
of ion-exchange on the thermal expan- thetic Dentistry, ed. WH Mörmann, pp.
sion behavior of leucite-reinforced den- 229–39. Chicago: Quintessence
tal porcelain. Presented at Am. Ceram. 132. Hassan R, Caputo AA, Bunshah RF.
Soc. Annu. Meet. Indianapolis, IN. Ab- 1981. Fracture toughness of human
str. p. 24 enamel. J. Dent. Res. 60:820–27
119. Denry IL, Holloway JA, Rosenstiel SF. 133. Rasmussen ST, Patchin RE. 1984. Frac-
1997. Effect of ion-exchange on the ture properties of human enamel and
crystallization behavior of a low ex- dentin in an aqueous environment. J.
pansion feldspar glass. J. Dent. Res. Dent. Res. 63:1362–68
76(IADR Abstr.) 138:A993 134. Xu HHK, Smith DT, Jahanmir S,
120. Fischer J, Pospiech P, Gernet W. 1992. Romberg E, Kelly JR, Thompson VP,
Chemical strengthening of leucite- Rekow ED. 1997. Indentation damage
containing dental ceramics. J. Eur. Ce- and mechanical properties of human
ram. Soc. 10:221–27 enamel and dentin. J. Dent. Res. In re-
121. Morena R, Lockwood PE, Evans AL, view
Fairhurst CW. 1986. Toughening of den- 135. Lawn BR, Padture NP, Cai H, Guib-
tal porcelain by tetragonal ZrO2 ad- erteau F. 1994. Making ceramics “duc-
ditions. J. Am. Ceram. Soc. 69:C75– tile.” Science 263:1114–16
77 136. Wuttiphan S, Lawn BR, Padture NP.
P1: ARK/SDA/SNY P2: SDA/VKS QC: SDA
May 21, 1997 14:19 Annual Reviews AR34-14

468 KELLY

1996. Crack suppression in strongly 140. Sherrill CA, O’Brien WJ. 1974. Trans-
bonded homogeneous/heterogeneous verse strength of aluminous and felds-
laminates: a study on glass/glass- pathic porcelain. J. Dent. Res. 53:683–
ceramic bilayers. J. Am. Ceram. Soc. 90
79:634–40 141. Meyers ML, Ergle JW, Fairhurst CW,
137. Liu H, Lawn BR, Hsu SM. 1996. Ringle RD. 1994. Fatigue characteris-
Hertzian contact response of tailored sil- tics of a high strength porcelain. Int. J.
icon nitride multilayers. J. Am. Ceram. Prosthodont. 7:253–57
Soc. 79:10009–14 142. Meyers ML, Ergle JW, Fairhurst CW,
138. Ritter JE. 1978. Engineering design and Ringle RD. 1994. Fatigue failure param-
fatigue failure of brittle materials. In eters of IPS-Empress porcelain. Int. J.
Fracture Mechanics of Ceramics, ed. Prosthodont. 7:549–53
RC Bradt, DPH Hasselman, FF Lange, 143. Rosenstiel SF, Denry IL, Zhu W, Gupta
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org

4:667–86. New York: Plenum PK, Van der Sluys RA. 1993 Fluo-
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

139. Jones DW, Sutow EJ. 1987. Stress corro- roalkylethyl silane coatings as a mois-
sion failure of dental porcelain. Br. Ce- ture barrier for dental ceramics. J.
ram. Trans. J. 86:40–43 Biomed. Mater. Res. 27:415–17
Annual Review of Materials Science
Volume 27, 1997

CONTENTS
FUTURE DIRECTIONS IN CARBON SCIENCE, M. S. Dresselhaus 1
THE NEW GENERATION HIGH-TEMPERATURE
SUPERCONDUCTORS, Z. Fisk, J. L. Sarrao 35
CERAMIC SCINTILLATORS, C. Greskovich, S. Duclos 69
CLAYS AND CLAY INTERCALATION COMPOUNDS: Properties and
Physical Phenomena, S. A. Solin 89
BISTABLE CHOLESTERIC REFLECTIVE DISPLAYS:Materials and
Drive Schemes, Deng-Ke Yang, Xiao-Yang Huang, Yang-Ming Zhu 117
BINDER REMOVAL FROM CERAMICS, Jennifer A. Lewis 147

CHARACTERIZATION OF POLYMER SURFACES WITH ATOMIC


FORCE MICROSCOPY, Sergei N. Magonov, Darrell H. Reneker 175
Annu. Rev. Mater. Sci. 1997.27:443-468. Downloaded from www.annualreviews.org
by b-on: Universidade do Minho (UMinho) on 03/21/11. For personal use only.

ELECTRICAL CHARACTERIZATION OF THIN-FILM


ELECTROLUMINESCENT DEVICES, J. F. Wager, P. D. Keir 223
LAYERED CERAMICS: Processing and Mechanical Behavior, Helen M.
Chan 249
MATERIALS FOR FULL-COLOR ELECTROLUMINESCENT
DISPLAYS, Yoshimasa A. Ono 283
LIQUID CRYSTAL MATERIALS AND LIQUID CRYSTAL
DISPLAYS, Martin Schadt 305
CHEMICAL FORCE MICROSCOPY, Aleksandr Noy, Dmitri V.
Vezenov, Charles M. Lieber 381
RECENT LIQUID CRYSTAL MATERIAL DEVELOPMENT FOR
ACTIVE MATRIX DISPLAYS, K. Tarumi, M. Bremer, T. Geelhaar 423
CERAMICS IN RESTORATIVE AND PROSTHETIC DENTISTRY, J.
Robert Kelly 443
LOCALIZED OPTICAL PHENOMENA AND THE
CHARACTERIZATION OF MATERIALS INTERFACES, Paul W.
Bohn 469
CERAMIC COMPOSITE INTERFACES: Properties and Design, K. T.
Faber 499
AN ATOMISTIC VIEW OF Si(001) HOMOEPITAXY, Zhenyu Zhang,
Fang Wu, Max G. Lagally 525
SUPERTWISTED NEMATIC (STN) LIQUID CRYSTAL DISPLAYS,
Terry Scheffer, Jürgen Nehring 555
PHOTOREFRACTIVE POLYMERS, W. E. Moerner, A. Grunnet-
Jepsen, C. L. Thompson 585
POLYCRYSTALLINE THIN FILM SOLAR CELLS: Present Status and
Future Potential, Robert W. Birkmire, Erten Eser 625

You might also like