You are on page 1of 10

Clinically relevant approach to failure testing of all-ceramic restorations

J. Robert Kelly, DDS, MS, DMedSca


National Institute of Standards and Technology, Gaithersburg, and Naval Dental Research Institute
Detachment, Naval Dental School, Bethesda, Md.
Statement of problem. One common test of single-unit restorations involves applying loads to clinically
realistic specimens through spherical indenters, or equivalently, loading curved incisal edges against flat
compression platens. As knowledge has become available regarding clinical failure mechanisms and the
behavior of in vitro tests, it is possible to constructively question the clinical validity of such failure testing
and to move toward developing more relevant test methods.
Purpose. This article reviewed characteristics of the traditional load-to-failure test, contrasted these with
characteristics of clinical failure for all-ceramic restorations, and sought to explain the discrepancies. Litera-
ture regarding intraoral conditions was reviewed to develop an understanding of how laboratory testing
could be revised. Variables considered to be important in simulating clinical conditions were described,
along with their recent laboratory evaluation.
Conclusions. Traditional fracture tests of single unit all-ceramic prostheses are inappropriate, because they
do not create failure mechanisms seen in retrieved clinical specimens. Validated tests are needed to elucidate
the role(s) that cement systems, bonding, occlusion, and even metal copings play in the success of fixed
prostheses and to make meaningful comparisons possible among novel ceramic and metal substructures.
Research over the past 6 years has shown that crack systems mimicking clinical failure can be produced in
all-ceramic restorations under appropriate conditions. (J Prosthet Dent 1999;81:652-61.)

CLINICAL IMPLICATIONS
Ideally, laboratory failure tests should provide meaningful information to clinicians
comparing all-ceramic restorative systems and aid in elucidating clinically important
variables. For test data to be relevant, laboratory tests should cause the same type(s) of
damage observed to occur in cases of clinical failure. Laboratory tests that create dam-
age uncharacteristic of clinical situations provide misleading guidance to clinicians.

L aboratory load-to-failure tests attempt to simulate


clinical failure to investigate variables thought to influ-
features reported from clinical experience or the exam-
ination of failed prostheses. Examination of clinically
ence the success of fixed prostheses and to evaluate new failed glass-ceramic crowns revealed that failures initiat-
materials or designs. These tests involve loading the ed from flaws and stresses existing at the cementation
occlusal surface of fixed prostheses with spherical surface as opposed to damage on the occlusal sur-
indenters, or equivalently using a flat compression face.16-18 In contrast, as will be discussed later, tradi-
platen against a curved incisal edge.1-15 Although tional laboratory tests induce failure from indentation
investigators have evaluated parameters such as tooth damage at the occlusal surface. Another problematic
preparation, various core and veneering materials, lut- characteristic of traditional load-to-failure testing is
ing agent type and thickness, and substructure elastic that extremely high loads are required (often 1500 to
modulus, little attention has been paid to the stress 5000 N)1,2,4,5,9,13,15 compared with those measured
state at failure or to the mechanism by which failures during mastication and swallowing (approximately 5 to
occurred during testing. More importantly, no work 364 N),19-23 or the maximum force recorded during
has been done to validate the clinical relevance of data clenching efforts (approximately 216 to 890 N).24-28
from such testing. Further, traditional laboratory tests are suspect in that
Many important characteristics associated with labo- failures involve the production of a great number of
ratory testing are simply not consistent with common fragments (versus the 2 fragments typical of clinical fail-
ure). Energy is required to create new surfaces during
Presented at the 80th annual meeting of the Academy of Prostho-
fracture, with the stress at failure being directly related
dontics, Colorado Springs, Colo., May 1998.
aGuest Scientist, Dental and Medical Research Group, National to the surface area created. Crushing damage accompa-
Institute of Standards and Technology; and Associate Professor, nied by formation of powder-like debris is not uncom-
Research Department, Naval Dental School. mon beneath the loading site during such tests, indi-

652 THE JOURNAL OF PROSTHETIC DENTISTRY VOLUME 81 NUMBER 6


KELLY THE JOURNAL OF PROSTHETIC DENTISTRY

Fig. 1. Cemented all-ceramic restorations comprise tri-material structures in which ceramic is


fully supported by and often bonded to dentin. High tensile stresses develop directly below
loaded area in ceramic at its interface with cement. These tensile stresses are far more sensi-
tive to stiffness of supporting material (core material or dentin) than to ceramic thickness.

cating that a high amount of energy was stored and supports and load points (namely, the moment arm),
then released during testing. the width of the bar, and the square of the thickness of
On the basis of the above mentioned discrepancies, the bar in a simple mathematical relationship derived
the conditions created during traditional failure tests do from solid mechanics.32 Unlike the simple specimen for
not appear to faithfully reproduce those seen clinically. bending tests, clinical ceramics are uniformly support-
Therefore it is of interest to explore reasons why this ed on a relatively elastic foundation (and often bonded
seemingly sensible traditional approach has led to spuri- to it) (Fig. 1). This raises the complexity of the stress
ous results and how such testing can be improved. Stress solution such that only finite element methods or iter-
conditions during contact loading and the subsequent ative mathematical approximations can be used to solve
failure response of brittle materials are governed by for the stresses as a function of load.33
some well-known variables, including load, contact area, For a ceramic layer uniformly supported by and
and the elastic moduli of the 2 materials in contact.29 In bonded to a less stiff material, high tensile stresses
addition, intraoral failure most likely involves the active develop in the ceramic at its interface with the cement,
participation of water, a phenomenon termed “chemi- directly below the loaded area as shown in the repre-
cally assisted crack growth” or “static fatigue.”30,31 sentation of a finite element result in Figure 1.34,35
This article examines these variables along with their These interfacial stresses arise from strain differences in
likely role in clinical failure from the perspective of a the ceramic, cement, and dentin because of the ceram-
person engaged in designing laboratory research. This ic being much stiffer (higher elastic modulus) than
information will then be used to explore the develop- either the cement or the dentin. These tensile stresses
ment of a meaningful laboratory fracture test for all- are extremely sensitive to the ratios of elastic moduli
ceramic single-unit fixed partial dentures. between the ceramic and the cement and dentin, and to
a much lesser extent the thickness of the ceramic and
DEFINING THE PROBLEM
the cement.35 Figure 2 illustrates the calculated maxi-
Ceramics supported by dentin mum tensile stresses as a function of thickness for iden-
Elementary beam theory cannot be used to examine tical bars of veneering ceramic either bonded to dentin
a cemented crown or to predict its clinical behavior. or loaded in 3-point bending (3 × 12 mm bar; 400 N
Homogeneous all-ceramic restorations consist of a load; 10 mm support for 3-point bending). Compared
layer of ceramic (approximately 1.0 to 2.0 mm thick) with the bar in 3-point bending, the bonded ceramic is
atop a layer of cement (approximately 30 to 120 µm virtually insensitive to thickness variations from 0.5 to
thick) supported by a thickness of dentin (Fig. 1). This 2.5 mm and is actually predicted to sustain less stress
structure is not well represented by simple bar-shaped at 0.5 mm than at 1.5 mm or 2.5 mm (35.6 MPa,
specimens such as those used in 3-point or 4-point 44.6 MPa, and 39.5 MPa, respectively). Better perfor-
bending tests. For such bar specimens, tensile stresses mance at 0.5 mm is attributed to the compressive
leading to fracture depend on distances between the stresses directly below the indenter countering tensile

JUNE 1999 653


THE JOURNAL OF PROSTHETIC DENTISTRY KELLY

Fig. 2. Calculated maximum tensile stresses as function of thickness for identical bars of
veneer porcelain either bonded to dentin or loaded in 3-point bending (10 mm support).
Stresses in bend bar follow (thickness)2 relationship, whereas bonded ceramic is virtually
insensitive to thickness variations from 1.0 to 2.5 mm.

stresses at the ceramic-cement interface.35 It remains to that can develop in a uniformly supported piece of
be seen whether this effect can be taken advantage of ceramic, arising from 3 distinctly different stress
clinically, although the phenomenon may partially help states29,35: (1) sharp indentation stresses, (2) blunt
to explain the high success experienced with veneer indentation stresses, and (3) cementation surface
restorations. (interfacial) stresses. If the contact area is small, as in
In addition to not following the thickness-squared the case of sharp indentation, median-radial crack sys-
relationship for stresses developed in simple beams, tems occur (Fig. 3, A). These cracks open onto the sur-
bonded ceramics develop higher stresses for a given face on unloading, and the surface crack traces are
load as the elastic modulus of the supporting materials clearly visible. Lateral cracks can also develop during
is decreased; 33 therefore stiffer cores may be better. the unloading phase of sharp indentation (Fig. 3, A).
One prediction from this effect is that at a given load, Such lateral cracks can propagate to the surface resulting
ceramic crowns cemented to cores of either a high alu- in material loss by chipping. Under blunt indentation,
mina ceramic or a nickel-chromium alloy will develop Hertzian cone cracks will pop-in from initial ring cracks
less cementation surface stress than crowns cemented located just outside of the contact area on the loaded
to resin composite cores. Taking this reasoning one surface (Fig. 3, B). For bonded dental ceramics, Hertz-
step further, the predicted ranking of core materials ian-origin cracks may not propagate to the cement inter-
based on such structural considerations would be as face, but instead have been observed to turn away from
follows: high alumina ceramic (most appropriate; for the interface as indicated by the dotted line extension to
example, In-Ceram, Vita Zahnfabrik, Bad Sackingen, the cone crack illustrated in Figure 3, B.35 When surface
Germany) ≅ nickel-chromium alloy > gold-based or contact stresses around a blunt indenter are kept low
palladium-based alloy > high leucite ceramic (OPC, enough to avoid Hertzian ring and cone formation, a
Jeneric/Pentron, Wallingford, Conn.; or Empress, single crack can preferentially be developed from the
Ivoclar AG, Schaan, Liechtenstein) > dentin > resin cementation surface of the ceramic and propagated
composite (least appropriate). toward the loaded surface (Fig. 3, C). Initially, while
still under the original load, this crack will be arrested
Contact stresses and crack systems
as it encounters a large compressive stress field below
There are at least 3 distinctly different crack systems the indenter. However, once this large crack exists in

654 VOLUME 81 NUMBER 6


KELLY THE JOURNAL OF PROSTHETIC DENTISTRY

A B

Fig. 3. Three crack systems that can result from indentation or loading of bonded ceramics:29, 35
(A) sharp indentation radial, median, and lateral cracks (shown in both surface and sectional
views); (B) blunt indentation Hertzian cone cracks (surface and sectional views shown); and
(C) cracks originating from cementation surface (sectional view only). Hertzian cone cracks
(B) observed to deviate away from bonded interface for dental porcelain bonded to less stiff
material, as indicated by dashed line.35

the ceramic part, complete failure is virtually assured


under loading from a slightly different angle or under
cyclic loading conditions.
Cracks in clinically failed and laboratory-
tested prostheses
Clinically failed ceramic crowns reveal features that
are quite consistent with those of the cementation sur-
face crack type in Figure 3, C.16-18 Cracks in clinical
crowns were not reported to begin from damage at the
occlusal surface, as would be the case for the crack sys-
tems presented in Figure 3, A and B. Clinical failures
were found to initiate from flaws and stresses located at
the cementation surface of ceramic crowns. In some
fracture surface views from clinical cases, wear facets are Fig. 4. Gold-coated fracture surface of glass-ceramic canine
loaded to failure against flat compression platen (at row of
seen directly opposite from the failure origin, providing
small arrows). Numerous fracture surface features,36 includ-
strong evidence that these crowns were broadly loaded ing twist hackle (A), crack branching (A), and Wallner lines
in the manner of Figure 3, C and that failures did not (B) indicate that crack propagated from damage at cusp tip,
originate from damage at the wear facets. not from ceramic-cement interface (C).
In contrast, traditional laboratory load-to-failure
tests appear to create cracks consistent with those
shown in Figure 3, A and B. Figure 4 illustrates a gold-
coated, glass-ceramic canine broken as part of a tradi- interpretation that crack propagation occurred from
tional load-to-failure experiment. Numerous fracture contact damage at the cusp tip, not from the ceramic-
surface features, for example, twist hackle, crack cement interface. The crown in Figure 4 was from a set
branching, Wallner lines,36 are consistent with the of 50 glass-ceramic canines found to belong to a single

JUNE 1999 655


THE JOURNAL OF PROSTHETIC DENTISTRY KELLY

failure distribution and for which crushing damage and (1) simple contact between 2 facets of similar material;
both Hertzian cone cracks and median cracks (Fig. 3, A (2) contact occurring simultaneously at 4 circular wear
and B) were commonly observed with no instances of facets of between 0.5 mm and 3.0 mm in diameter; and
cementation interface crack initiation.37 Similar obser- (3) the load range being 100 to 700 N.
vations can be made from published photographs of in
Simulating clinical contact stresses
vitro test specimens.7,9,10,13 Thus, data collected dur-
ing this type of laboratory testing are clinically invalid It also is of interest to know how to design a labora-
as a result of: (1) an incorrect stress state, (2) failure tory experiment with spherical indenters to create the
occurring from contact damage flaws instead of “nat- range of contact pressures calculated previously (3.5 to
ural” cementation surface flaws, (3) crack systems caus- 890 MPa). Contact pressures (P) between spheres and
ing failure not seen in bulk clinical fracture, and (4) the flat surfaces of different materials, as used in tradition-
failure loads being too high for clinical significance. al load-to-failure testing, can be calculated with the
These crowns failed from artificial damage caused by following relationships29:
inappropriate loading under the wrong environmental
conditions.
1 4kr2
3E1 2⁄3 L1⁄3
P= •
UNDERSTANDING CLINICAL
π
CONDITIONS TO BE MODELED
Clinical contact stresses where k is given by:
Loaded restorations and teeth develop features that
have come to be called “wear facets.” Clinicians recog- 9 E1
nize that wear facets are usually not point contacts, but k= ·[(1 – v12)+(1 – v22)· ]
16 E2
have dimensions of approximately 0.5 to 3.0 mm in
diameter. Unfortunately, the sizes and size distributions
of wear facets does not appear to have been document- with elastic modulus E1 of the flat surface material and
ed in the literature, except for anthropologic studies of E2 of the spherical surface material; v1 and v2 are the
aboriginal populations38, 39 and hominid remains.40,41 respective Poisson’s ratios; L the applied load; and r the
However, the size of this contact area along with the radius of the spherical indenter. Figures 5 and 6 depict
loads generated during chewing and bruxism have a the contact pressures and contact radii developed
significant influence on which damage mechanisms are below steel spheres (E2 = 180 GPa; v2 ≈ 0.25) loading
likely to predominate intraorally. Therefore it is impor- dental porcelain (E1 = 70 GPa; v1 ≈ 0.25) for various
tant to at least determine a likely range for these values ball sizes.
as a guide to laboratory test development. Steel indenter balls would need to have diameters
Average total contact area on posterior teeth during between 40 mm to 1 m so as to develop clinically real-
maximum intercuspation has been reported to be istic contacts (Fig. 5); namely, 0.5 to 3 mm diameter
52 mm2 for females and 64 mm2 for males.42 On the (clinical wear facet size) at contact pressures of 5 to
basis of these contact areas, even at an extreme load of 890 MPa for loads between 100 to 700 N (encom-
3500 N, simple contact pressures would only reach passing the range of realistic average maximum bite
between 55 and 67 MPa. However, chewing or brux- forces as developed previously). Traditional load-to-
ism forces are more likely being delivered not to the failure tests have typically used spherical steel indenters
whole arch but to opposing quadrants or tooth groups. of well below 10 mm in diameter, conditions that can
Further, it is more realistic for purposes of experimen- be placed into perspective with reference to Figure 6.
tal design to consider functional forces (that is, chew- Contact pressures beneath a 5-mm diameter steel ball
ing, swallowing, and bruxism) than to base tests on the on dental porcelain (for 100 to 700 N loads) would be
maximum force achievable by extreme persons. It is 1384 to 2648 MPa on contacts of only 0.3 to 0.6 mm
also common in experimental design to chose average in diameter. Traditional tests often require loads of
values or a range of values that fall within the “enve- 1500 to 3000 N to generate failure.1,2,4,5,9,13,15 For
lope” of reported data. One research group extensively comparison purposes, Figure 6 illustrates this tradition-
involved with simulating the conditions of mastication al failure test region (2 to 4 mm diameter balls; 1500
has chosen the range of 9 to 180 N.43 Average maxi- to 3000 N load).
mum biting forces on premolars and molars are report- Tooth-to-tooth contacts do not appear to be well
ed to be between 150 and 665 N.44 Posterior forces, represented experimentally by small steel balls. Such
especially when high, are not likely to be delivered to spherical indenters cause failure from mechanisms not
only 1 tooth but will be distributed. With the previ- seen clinically. Similarly, high contact stresses would be
ously mentioned caveats in mind, calculated contact generated between a curved incisal edge and a flat com-
pressures range between 3.5 and 890 MPa assuming: pression platen. High contact pressures favor the medi-

656 VOLUME 81 NUMBER 6


KELLY THE JOURNAL OF PROSTHETIC DENTISTRY

Fig. 5. Calculated contact facet size and contact pressure developed by steel intenter balls of
various sizes on dental porcelain. For loads of between 100 to 700 N, steel indenter balls
would need to have diameters between 40 mm to 1 m to develop clinically sized contacts of
0.5 to 3 mm diameter at pressures of 5 to 890 MPa. Load ranges for mastication and bruxism
are defined in text.

Fig. 6. Expansion of Figure 5 to include higher contact pressures developed during tradition-
al tests. Such tests often use 2 to 4 mm diameter balls and require loads of 1500 to 3000 N
to generate failure, conditions contained within gray box to right. Clinically relevant region
from Figure 5 appears as tall box region to left (filled dots represent realistic ball sizes/load
conditions).

JUNE 1999 657


THE JOURNAL OF PROSTHETIC DENTISTRY KELLY

an-radial cracks and Hertzian cone cracks that occur at shown to be sensitive to water in static strength
the contact surface as illustrated in Figure 3, A and B, tests,51,52 under cyclic loading,53 and in measurements
and in an actual test specimen in Figure 4, which was of the fatigue (or slow crack growth) exponent.54,55 At
produced using a curved canine against a flat platen. least 1 research group has evaluated hydrophobic
These cracks do not involve either the flaws or stresses silanes as moisture barriers to improve clinical longevi-
located at the cementation interface that are responsi- ty.56 Water is available to all external surfaces from sali-
ble for clinical failure. In addition, the critical load to va; however, it should be noted that the cementation
initiate a Hertzian crack (Pc) increases linearly as the surface is also exposed to water available from the
radius of the indenter (r) increases in a proportionality dentin and from transport through dental cements.
known as the Auerbach Law (Pc ∝ r).45 For spherical Although chemically assisted crack growth is likely to
indenters larger than 35 mm in radius the proportion- be involved in clinical failure, it is rarely if ever includ-
ality shifts from linear to quadratic (Pc ∝ r2).46 The ed in traditional laboratory tests. No known work
Auerbach effect has been shown for dental porcelains37 attempts to link slow crack growth behavior of various
and poses special problems where the prosthesis pro- ceramics to clinical failure rates.
vides a curved surface (namely, central incisor and pos-
TOWARD SIMULATING CLINICAL
terior occlusal anatomy) because radii of curvature are
FAILURE
not easily controlled in fired dental porcelain.
Producing the appropriate crack system
Influence of cyclic loading under low loads
The use of large balls (>40 mm in diameter) to load
Chewing obviously consists of high numbers of low all-ceramic prostheses is unrealistic for laboratory
cyclic loads. Certain engineering ceramics, including one experimentation. However, large radii can be machined
similar to the dental glass-ceramic Dicor (Dentsply onto the ends of steel pistons to create contact stresses
International, York, Pa.), accumulate damage during and contact areas within the target range calculated
low-load cyclic blunt indentation that is not seen during previously, based on clinical observations and measure-
single-load failure testing.47,48 This accumulated subsur- ments. Flat pistons should be avoided because they can
face damage can weaken the ceramic specimen and is become edge-loading as the ceramic deforms slightly
thought by some to explain clinical failure.49 However, beneath the piston. Another benefit to large radii is that
2 factors suggest that this indentation-related concept contact pressures do not increase as fast with increasing
lacks any general clinical relevance in the case of bulk load as they do with small balls (Figs. 5 and 6), so that
fracture. First, no accumulated microstructural damage, it is easier to avoid the initiation of Hertzian and medi-
as reported in other works,47-49 was identified during an-radial cracks. Thin sheets of polyethylene can also be
SEM examination of the failure origins of clinically failed interspersed between the large radius piston and the
glass-ceramic crowns.16-18 Such damage would have specimen to further assure a broad even contact. Such
been quite prominent if it existed. Second, cemented a technique35 has been tested successfully, routinely
feldspathic (leucite-filled) crowns were not weakened by creating cementation surface cracks beneath bars of a
cyclic loading for 1 million cycles at loads well above feldspathic porcelain, and it has been further developed
those associated with chewing (up to 800 N).50 The for single units, to be discussed as follows.
mean failure loads for these crowns was found to be
Creating meaningful failure data
1563 N, regardless of whether they were cyclically
loaded or what load was used (200, 400, 600, or Ceramic restorations should fail in the laboratory
800 N). It may be significant that the cementation sur- from the same flaw type and size, causing failure intra-
face of these silane-treated and resin-bonded crowns was orally. To enhance this probability, the cementation
likely quite dry and well-protected from atmospheric surface of laboratory specimens should be created and
moisture. Under dry conditions, low cyclic loads alone treated as closely as possible to clinical situations. Flaws
do not appear to lead to any cementation surface dam- can either be inherent to the microstructure of the
age accumulation in feldspathic ceramics and cannot ceramic, for example, cracking around leucite grains
alone be used to induce failure under appropriate clini- during cooling, or can result from processing steps
cal loads. The mean load of 1563 N is still too high for (such as sandblasting, diamond grinding). It is unlikely
clinical significance, even though the correct crack sys- that the same flaw populations would exist at the sur-
tem was involved (cementation surface cracks without face of ceramic test bars machined from large blocks as
any Hertzian ring or cone cracks). in test bars fabricated of the same material that was
pressed into and then recovered from a gypsum mold.
Role of water in brittle fracture
Correct failure stresses also need to be created to
Since 1958, it has been recognized that water can induce realistic fracture. Cementation surface stresses
act chemically at crack tips to decrease the strength of are a function of the elastic moduli of the support and
glasses and ceramics.30,31 Dental ceramics have been cement, the thickness of the ceramic and cement, the

658 VOLUME 81 NUMBER 6


KELLY THE JOURNAL OF PROSTHETIC DENTISTRY

shape of the ceramic, and the applied load. Again, as for 1 million cyclic loads that included contributions from
flaws, the best approach is to test specimens that are as mastication and swallowing.44,59 Such a mean cyclic
close to clinical restorations as possible. Thickness and load would be expected to be lower than that measured
shape are best controlled by creating clinically relevant from a static test (single increasing load until failure)
test specimens. Popular cements should be chosen (and under wet conditions as predicted from mathematical
these have a narrow range of elastic properties), which models describing the relationship between static and
leaves only the dentin to be correctly modeled. Because cyclic stress states.60 Likewise, mean failure load might
the stresses of interest are being created during blunt also be expected to shift to even lower loads at 2 or 3
loading, it makes sense to evaluate candidate dentin years. Time to failure under cyclic versus static condi-
substitutes in a similar fashion. Stress-strain behavior tions is quite sensitive to the slow crack growth expo-
(both elastic and plastic) can be evaluated by blunt nent (n, describing the sensitivity of the ceramic to
indentation at various loads.57 On the basis of such chemically assisted crack growth), as per the following
testing, an epoxy matrix containing woven glass fibers approximation60:
was chosen as approximately representing the elastic
and plastic (permanent deformation) response of tc σs n
dentin during blunt loading.58 Such materials are easi- ts
≅ g –11 2
σa
ly machined into replicas of prepared teeth (or disks)
for use during laboratory testing.
As previously stated, cracks originating from the where tc = time to failure under cyclic conditions;
cementation surface can be reproducibly formed dur- ts = time to failure under static conditions; σs = static
ing static testing when blunt loading is used under con- failure stress; σa = cyclic stress amplitude; and factor,
ditions that avoid Hertzian cone cracking. However, g-1, is obtained using a series solution for sinusoidal
the usual method of detecting breakage or failure, a stresses.60
sharp drop in load, cannot be used. The “pop-in” of a On the basis of the above relationship, the predicted
cementation surface crack is not accompanied by any wet static mean failure load would be 684 N, or
drop in the load, because the test piece is still uniform- approximately 44% of the 1563 N value found during
ly supported on its base and continues to fully support dry static testing (assuming: [1] linear elastic fracture
the loading piston. For bars of feldspathic porcelain, mechanics, so load and stress can be directly related;
the load for crack pop-in could be determined by trans- [2] a static failure time of approximately 5 seconds; [3]
illumination.35 Cracks were identified both visually and a mean 1-year cyclic failure load of 275 N; and [4] n =
with a digital photometer. However, when testing full 27 for the slow crack growth exponent.)55
crowns, it is quite difficult to detect the occurrence of Figure 7 illustrates calculated mean failure loads ver-
crack pop-in by such a technique during loading. For sus days of function, assuming 1 million cycles per year,
full-crown specimens, acoustic monitoring was found data that were collected from analogs of single-unit
to be a reliable method for crack detection.58 prostheses that were fabricated and cemented accord-
ing to common dental laboratory and clinical prac-
Failure under wet cyclic loading
tices.50 It is noted that the prediction in Figure 6 that
Loads for cementation-surface cracking under dry, mean failure loads decrease most dramatically in the
static testing were still too high to be meaningful clini- first days after insertion and are not predicted to
cally (1563 N), even after cyclic loading for 1 million decrease in likewise fashion beyond 1 year. The predic-
cycles at 800 N, which represents approximately 1 year tions in Figure 7 and the relationship between failure
of constant bruxism. However, when identical speci- loads and failure rates of a population of crowns are
mens modified with tiny channels (pseudodentin worth exploring.
tubules) were submerged in water for 2 months before This lack of substantive linkage between traditional
testing and then cyclically loaded under water, the mean laboratory results and clinical behavior leaves a number
failure load dropped to 275 N, well within the range of basic design questions unanswered. For example, is
reported for clinical loads.50 In this experiment, the the success of metal-ceramic crowns simply because of
presence or absence of cracking was evaluated by transil- the rigidity of the cast metal substructure? Alternative-
lumination at the end of the 1 million cycles of loading. ly, does this successful prosthesis depend on unique
Subsurface (cementation surface originating) cracks stress distributions, moisture barrier effects, or crack
were easily detected when present, and there were no bridging that can be replicated by materials other than
instances of Hertzian ring or cone cracking or cracking cast alloys? How thin can the metal be and where must
from damage at the loaded surface (examined by it be used within the substructure? How does one eval-
Nomarski interference contrast microscopy at ×10).50 uate different coping designs and materials, including
This mean failure load of 275 N was that measured powder metallurgy and various foil products? Can an
after the equivalent of approximately 1 year of function; appropriate cement be developed to substitute for the

JUNE 1999 659


THE JOURNAL OF PROSTHETIC DENTISTRY KELLY

Fig. 7. Measured and calculated failure loads for wet and dry conditions under both static and
cyclic loading. Measured failure loads (dots) were collected from analogs of single unit pros-
theses fabricated and cemented according to common practices.50 Graph assumes that 1 mil-
lion cyclic loads (mastication and swallowing) occur per year.

metal coping if the cement is well-bonded to the Much information in dental literature may be mis-
ceramic and dentin? Why did glass-ceramic crowns leading regarding comparative “strengths” of single
bonded with methacrylate-based cements have such unit prostheses and the function of some clinically
improved survival rates versus crowns cemented with important variables, including tooth preparation,
zinc phosphate cements?61 A new cement adaptation ceramic thickness, coping and veneering materials,
theory addressing this last question was recently pro- cementation and bonding procedures, cement type and
posed and awaits testing.62 thickness, and coping design. Cracks that mimick clin-
These are all questions of genuine interest that ical failure can be produced at realistic intraoral loads
deserve far more fundamental thought than they have under the modified testing protocol described in this
received. Answers to these and related questions await article. It remains to be seen whether this testing
the use of validated structural tests and mathematical approach creates damage having relevance to modeling
models developed in conjunction with further elucida- the clinical behavior of metal-ceramic restorations.
tion of clinical failure mechanisms.
CONCLUSIONS REFERENCES
1. Josephson BA, Schulman, A, Dunn ZA, Hurwitz W. A compressive
Significant differences were found between the fail- strength study of an all-ceramic crown. J Prosthet Dent 1985;53:301-3.
ure behavior created during traditional load-to-failure 2. Josephson BA, Schulman, A, Dunn ZA, Hurwitz W. A compressive
tests and that observed to have occurred during clinical strength study of complete ceramic crowns. Part II. J Prosthet Dent
1991;65:388-91.
failure of all-ceramic restorations. Traditional loading 3. Dickinson AJ, Moore BK, Harris RK, Dykema RW. A comparative study of
conditions created contact stresses that favor the for- the strength of aluminous porcelain and allceramic crowns. J Prosthet
mation of median-lateral crack systems, Hertzian cone Dent 1989;61:297-304.
4. Smith TB, Kelly JR, Tesk JA. In vitro fracture behavior of ceramic and
cracks, and localized crushing damage: None of which metal-ceramic restorations. J Prosthodont 1994;3:138-44.
are reported to cause the bulk failure of clinical all- 5. Carrier DD, Kelly JR. In-Ceram failure behavior and core-veneer interface
ceramic restorations. Thus, it appears that traditional quality as influenced by residual infiltration glass. J Prosthodont
1995;4:237-42.
laboratory tests do not (1) create appropriate stress
6. Vrijhoef MM, Spanauf AJ, Renggli HH. Axial strengths of foil, all-ceramic
states, (2) cause failure from clinically relevant flaws, or and PFM molar crowns. Dent Mater 1988;4:15-9.
(3) create crack systems modeling clinical failure. 7. Munoz CA, Goodacre CJ, Moore BK, Dykema RW. A comparative study

660 VOLUME 81 NUMBER 6


KELLY THE JOURNAL OF PROSTHETIC DENTISTRY

of the strength of aluminous porcelain jacket crowns constructed with the 39. Molnar S, McKee JK, Molnar I. Measurements of tooth wear among Aus-
conventional and twin foil technique. J Prosthet Dent 1982;48:271-81. tralian aborigines: I. Serial loss of the enamel crown. Am J Phys Anthro-
8. Miller A, Long J, Miller B, Cole J. Comparison of the fracture strengths of pol 1983;61:51-65.
ceramometal crowns versus several all-ceramic crowns. J Prosthet Dent 40. Puech PF, Albertini H. Dental microwear and mechanisms in early
1992;68:38-41. hominids from Leatoli and Hadar. Am J Phys Anthropol 1984;65:87-91.
9. Ferro KJ, Myers ML, Graser GN. Fracture strength of full-contoured ceram- 41. Whittaker DK, Ryan S, Weeks K, Murphy WM. Patterns of approximal
ic crowns and porcelain-veneered crowns of ceramic copings. J Prosthet wear in cheek teeth of a Romano-British population. Am J Phys Anthropol
Dent 1994;71:462-7. 1987;73:389-96.
10. Yoshinari M, Derand T. Fracture strength of all-ceramic crowns. Int J 42. Julien KC, Buschang PH, Throckmorton GS, Dechow PC. Normal masti-
Prosthodont 1994;7:329-38. catory performance in young adults and children. Arch Oral Biol
11. Burke FJ, Watts DC. Fracture resistance of teeth restored with dentin- 1996;41:69-75.
bonded crowns. Quintessence Int 1994;25:335-40. 43. DeLong R, Douglas WH. Development of an artificial oral environment
12. McCormick JT, Rowland W, Shillingburg HT Jr, Duncanson MG Jr. Effect for the testing of dental restoratives: bi-axial force and movement control.
of luting media on the compressive strengths of two types of all-ceramic J Dent Res 1983;62:32-6.
crown. Quintessence Int 1993;24:405-8. 44. Craig RG. Restorative dental materials. 10th ed. St Louis: Mosby; 1997. p.
13. Brukl CE, Ocampo RR. Compressive strengths of a new foil and porcelain- 56-57, 76.
fused-to-metal crowns. J Prosthet Dent 1987:57:404-10. 45. Frank FC, Lawn BR. On the theory of Hertzian fracture. Proc Roy Soc
14. Passi P, Girardello GB, Vesentini A. Resistance to fracture of ceramic jack- 1967:A299:291-306.
et crowns. Quintessence Int 1992;23:845-7. 46. Tillet JPA. Fracture of glass by spherical indenters. Proc Roy Soc
15. Scherrer SS, de Rijk WG. The fracture resistance of all-ceramic crowns on 1956;B69:47-54.
supporting structures with different elastic moduli. Int J Prosthodont 47. Guiberteau F, Padture NP, Cai H, Lawn BR. Indentation fatigue: a simple
1993;6:462-7. cyclic Hertzian test for measuring damage accumulation in polycrys-
16. Kelly JR, Campbell SD, Bowen HK. Fracture-surface analysis of dental talline ceramics. Phil Mag 1993;68(A):1003-16.
ceramics. J Prosthet Dent 1989;62:536-41. 48. Lawn BR, Padture NP, Cai H, Guiberteau F. Making ceramics “ductile”.
17. Kelly JR, Giordano R, Pober R, Cima MJ. fracture surface analysis of dental Science 1994;263:1114-6.
ceramics: clinically failed restorations. Int J Prosthodont 1990;3:430-40. 49. Peterson IM, Pajares A, Lawn BR, Thompson VP, Rekow ED. Mechanical
18. Thompson JY, Anusavice KJ, Naman A, Morris HF. Fracture surface char- characterization of dental ceramics by Hertzian contacts. J Dent Res
acterization of clinically failed all-ceramic crowns. J Dent Res 1998;77:589-602.
1994;73:1824-32. 50. Kelly JR, Hunter BD, Brenyo MR, Peterson IM. Simulating clinical failure
19. Gibbs CH, Mahan PE, Lundeen HC, Brehnan K, Walsh EK, Holbrook WB. during in vitro testing of all-ceramic crowns. J Dent Res 1998;77:778
Occlusal forces during chewing and swallowing as measured by sound (abstract 1175).
transmission. J Prosthet Dent 1981;46:443-9. 51. Jones DW, Sutow EJ. Stress corrosion failure of dental porcelain. Br Ceram
20. Lundgren D, Laurell L. Occlusal force patterns during chewing and biting Tran J 1987;86:40-3.
in dentitions restored with fixed bridges of cross-arch extension. I. Bilat- 52. Fairhurst CW, Lockwood PE, Ringle RD, Twiggs SW. Dynamic fatigue of
eral end abutments. J Oral Rehabil 1986;13:57-71. feldspathic porcelain. Dent Mater 1993;9:269-73.
21. De Boever JA, McCall WD Jr, Holden S, Ash MM. Functional occlusal 53. Drummond JL, Van Scoyoc JP, Racean DC. Aging of ion exchanged porce-
forces: an investigation by telemetry. J Prosthet Dent 1978;40:326-33. lain. In: Fischman G, Clare A, Hench L, editors. Bioceramics: materials
22. Anderson DJ, Pincton DC. Masticatory stresses in normal and modified and applications. Ceramic transactions. Vol. 48. Westerville (OH): Amer-
occlusion. J Dent Res 1958;37:312-7. ican Ceramic Society; 1995. p. 191-200.
23. Hagberg C. Electromyography and bite force studies of muscular function 54. Myers ML, Ergle JW, Fairhurst CW, Ringle RD. Fatigue characteristics of a
and dysfunction in masticatory muscles. Swed Dent J Suppl 1986;37:1-64. high-strength porcelain. Int J Prosthodont 1994;7:253-7.
24. Mansour RM, Reynik RJ. In vivo occlusal forces and moments: I. Forces 55. Myers ML, Ergle JW, Fairhurst CW, Ringle RD. Fatigue failure parameters
measured in terminal hinge position and associated moments. J Dent Res of IPS-Empress porcelain. Int J Prosthodont 1994;7:549-53.
1975;54:114-20. 56. Rosenstiel SF, Denry IL, Zhu W, Gupta PK, Van der Sluys RA. Fluo-
25. Hellsing E, Hagberg C. Changes in maximum bite force related to exten- roalkylethyl silane coatings as a moisture barrier for dental ceramics. J
sion of the head. Eur J Orthod 1990;12:148-53. Biomed Mater Res 1993;27:415-7.
26. Helkimo E, Carlsson GE, Helkimo M. Bite force and state of dentition. 57. Swain MV, Hagan JT. Indentation plasticity and the ensuing fracture of
Acta Odontol Scand 1977;35:297-303. glass. J Phys D Appl Phys 1976;9:2201-14.
27. Hagberg C, Agerberg G, Hagberg M. Regression analysis of electromyo- 58. Ellert DO, Kelly JR. All-ceramic crown failure as a function of occlusal
graphic activity of masticatory muscles versus bite force. Scand J Dent Res contact location. (IADR abstracts) J Dent Res 1997;76:392 (abstract
1985;93:396-402. 3030).
28. Kikuchi M, Korioth TW, Hannam AG. The association among occlusal 59. Pehlivan M, Yüceyar N, Ertekin C, Celebi G, Ertas M, Kalayci T, et al. An
contacts, clenching effort, and bite force distribution in man. J Dent Res electronic device measuring the frequency of spontaneous swallowing:
1997;76:1316-25. digital phagometer. Dysphagia 1996;11:259-64.
29. Lawn BR. Fracture of brittle solids. 2nd ed. Cambridge, U.K.: Cambridge 60. Evans AG, Fuller ER. Crack propagation in ceramic materials under cyclic
University Press; 1993. p. 249-306. loading conditions. Met Trans 1994;5:27-33.
30. Charles RJ. Static fatigue of glass. I. J Appl Phys 1958;29:1549-53. 61. Malament KA, Socransky SS. Survival of Dicor glass-ceramic dental
31. Charles RJ. Static fatigue of glass. II. J Appl Phys 1958;29:1554-60. restorations over 14 years: part I. Survival of Dicor complete coverage
32. Cook NH. Mechanics and materials for design. New York: McGraw-Hill; restorations and effect of internal surface acid etching, tooth position,
1984. p. 238-49. gender and age. J Prosthet Dent 1999;81:23-32.
33. Young, WC. Roark’s formulas for stress and strain. New York: McGraw- 62. Thompson JY, Rapp MM, Parker AJ. Microscopic and energy dispersive x-
Hill; 1989. p. 449. ray analysis of surface adaptation of dental cements to dental ceramic sur-
34. Anusavice KJ, Hojjatie B. Tensile stresses in glass-ceramic crowns: effect faces. J Prosthet Dent 1998;79:378-83.
of flaws and cement voids. Int J Prosthodont 1992;5:351-8.
35. Morris DR, Kelly JR. Failure loads of bonded ceramics influenced by Reprint requests to:
hydrostatic compressive stresses. (AADR Abstracts) J Dent Res 1995;74: DR J. ROBERT KELLY
220 (abstract 1666). DENTAL AND MEDICAL MATERIALS GROUP
36. Frechette VD. Failure analysis of brittle materials, advances in ceramics. NATIONAL INSTITUTE OF STANDARDS AND TECHNOLOGY
Vol. 28. Westerville (OH): American Ceramic Society; 1990. 100 BUREAU DR, MAIL STOP 8545
37. Harvey CK, Kelly JR. Contact damage as a failure mode during in vitro GAITHERSBURG, MD 20899-8545
testing. J Prosthodont 1996;5:95-100. FAX: (301)963-9143
38. Kaidonis JA, Richards LC, Townsend GC. Nature and frequency of dental E-MAIL: robert.kelly@nist.gov
wear facets in an Australian aboriginal population. J Oral Rehabil 10/1/97805
1993;20:333-40.

JUNE 1999 661

You might also like