26.11.2012 Views

Fracture behavior of lithia disilicate- and leucite-based ceramics

Fracture behavior of lithia disilicate- and leucite-based ceramics

Fracture behavior of lithia disilicate- and leucite-based ceramics

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Dental Materials (2004) xx, xxx–xxx<br />

<strong>Fracture</strong> <strong>behavior</strong> <strong>of</strong> <strong>lithia</strong> <strong>disilicate</strong>- <strong>and</strong><br />

<strong>leucite</strong>-<strong>based</strong> <strong>ceramics</strong><br />

Alvaro Della Bona a, *, John J. Mecholsky Jr. b , Kenneth J. Anusavice c<br />

a School <strong>of</strong> Dentistry, The University <strong>of</strong> Passo Fundo, P.O. Box 611/613, Passo Fundo, RS 99001 970, Brazil<br />

b Department <strong>of</strong> Material Sciences <strong>and</strong> Engineering, University <strong>of</strong> Florida, Gainesville, FL, USA<br />

c Department <strong>of</strong> Dental Biomaterials, University <strong>of</strong> Florida, Gainesville, FL, USA<br />

Received 8 July 2003; received in revised form 3 February 2004; accepted 19 February 2004<br />

KEYWORDS<br />

Structural reliability;<br />

<strong>Fracture</strong>; Fractography;<br />

<strong>Fracture</strong> toughness;<br />

Weibull modulus<br />

Introduction<br />

The appeal <strong>of</strong> <strong>ceramics</strong> as structural dental<br />

materials is <strong>based</strong> on their esthetics, low density,<br />

high hardness, chemical inertness, <strong>and</strong> wear resistance.<br />

A major goal <strong>of</strong> ceramic research <strong>and</strong> development<br />

is to produce stronger, tougher <strong>ceramics</strong><br />

that are structurally reliable in dental applications.<br />

Limited information is available on the fracture<br />

<strong>behavior</strong> <strong>and</strong> toughness <strong>of</strong> hot-pressed dental<br />

*Corresponding author. Tel.: þ55-54-3115142; fax: þ55-54-<br />

3168403.<br />

E-mail address: dbona@upf.br<br />

ARTICLE IN PRESS<br />

www.intl.elsevierhealth.com/journals/dema<br />

Summary Objective: This study was designed to characterize the fracture <strong>behavior</strong> <strong>of</strong><br />

<strong>ceramics</strong> <strong>and</strong> test the hypothesis that variation in strength is associated with a<br />

variation in fracture toughness.<br />

Methods: The following four groups <strong>of</strong> 20 bar specimens (25 £ 4 £ 1.2 mm) were<br />

fabricated (ISO st<strong>and</strong>ard 6872): E1, a hot-pressed <strong>leucite</strong>-<strong>based</strong> core ceramic<br />

(IPS Empress); E2, a hot-pressed <strong>lithia</strong>-<strong>based</strong> core ceramic (IPS Empress 2); ES, a<br />

hot-pressed <strong>lithia</strong>-<strong>based</strong> core ceramic (Experimental); <strong>and</strong> GV, a glass veneer (IPS<br />

Empress2 body). Specimens were subjected to four-point flexure loading in 37 8C<br />

distilled water. Fractographic analysis was performed to determine the fracture origin<br />

ðcÞ for calculation <strong>of</strong> fracture toughness ðK ICÞ: Weibull analysis <strong>of</strong> flexure strength ðsÞ<br />

data was also performed.<br />

Results: Differences in mean s <strong>and</strong> K IC were statistically significant for E1 <strong>and</strong> GV<br />

ðp , 0:05Þ: These differences are associated with processing effects <strong>and</strong> composition.<br />

Significance: The higher mean s <strong>and</strong> K IC values <strong>of</strong> E2 <strong>and</strong> ES core <strong>ceramics</strong> suggest<br />

potentially improved structural performance compared with E1 although the Weibull<br />

moduli <strong>of</strong> E1 <strong>and</strong> E2 are the same.<br />

Q 2004 Academy <strong>of</strong> Dental Materials. Published by Elsevier Ltd. All rights reserved.<br />

<strong>ceramics</strong>. Brittle fracture is a complex process. 1,2<br />

Quantitative fractographic analysis facilitates the<br />

identification <strong>of</strong> fracture mechanisms through the<br />

use <strong>of</strong> fracture mechanics. 3<br />

<strong>Fracture</strong> toughness is <strong>of</strong>ten used to characterize<br />

the fracture resistance <strong>of</strong> brittle materials 4–6 <strong>and</strong> it<br />

is usually controlled by the fracture in Mode I<br />

(opening mode, tensile load). Irwin 16 defined<br />

failure at the point when the Mode I stress intensity<br />

(KI) reaches or exceeds a critical value ðK I $ K ICÞ:<br />

The critical stress intensity factor ðK ICÞ is in many<br />

cases a material constant <strong>and</strong> is one measure <strong>of</strong> the<br />

toughness <strong>of</strong> the material, i.e. the resistance<br />

to crack propagation. Therefore, the fracture<br />

0109-5641/$ - see front matter Q 2004 Academy <strong>of</strong> Dental Materials. Published by Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.dental.2004.02.004


2<br />

Figure 1 Diagram <strong>of</strong> the typical fracture surface features occurring in brittle materials. The regions are not drawn to<br />

scale. Adapted from Mecholsky (1993). 27<br />

toughness or critical stress intensity factor ðK ICÞ can<br />

<strong>of</strong>ten be determined using the Griffith–Irwin<br />

equation:<br />

K IC ¼ Ys fc 1=2<br />

ð1Þ<br />

where Y is a geometrical factor that accounts for<br />

the location <strong>and</strong> geometry <strong>of</strong> the critical flaw <strong>and</strong><br />

type <strong>of</strong> loading, 7 s f is the stress at fracture, <strong>and</strong> c is<br />

the radius <strong>of</strong> an equivalent semicircular flaw for a<br />

semi-elliptical crack <strong>of</strong> semiminor axis ‘a’ <strong>and</strong><br />

semimajor axis ‘b’ (Fig. 1). 8,9<br />

Little information is available on the structural<br />

reliability <strong>of</strong> hot-pressed <strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> <strong>and</strong><br />

<strong>leucite</strong>-<strong>based</strong> <strong>ceramics</strong>, <strong>and</strong> the sensitivity <strong>of</strong> processing<br />

variables is unknown. This study was designed to<br />

characterize the fracture <strong>behavior</strong> <strong>and</strong> fracture<br />

toughness <strong>of</strong> a glass veneer (GV), a <strong>leucite</strong>-<strong>based</strong><br />

ceramic, <strong>and</strong> two <strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> <strong>ceramics</strong><br />

using fractographic principles. The objective <strong>of</strong> this<br />

study was to test the hypothesis that variation in<br />

strength is associated with the variation in fracture<br />

toughness for the same surface preparation.<br />

Materials <strong>and</strong> methods<br />

The ceramic materials (E1, E2, ES, <strong>and</strong> GV) <strong>and</strong><br />

firing procedures used for bar specimens<br />

Table 1 Firing temperatures ðTÞ <strong>and</strong> times ðtÞ for the four <strong>ceramics</strong>.<br />

Dental <strong>ceramics</strong> Starting T<br />

(8C)<br />

(25 £ 4 £ 1.2 mm) are summarized in Table 1. The<br />

GV specimens were fabricated using a metal mold<br />

<strong>and</strong> fired according to the manufacturer’s instructions<br />

(Table 1). The hot-pressed <strong>leucite</strong>-<strong>based</strong> core<br />

ceramic (E1) <strong>and</strong> the two hot-pressed <strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong><br />

core ceramic specimens (E2 <strong>and</strong> ES)<br />

were prepared using the lost-wax technique. 10<br />

After removal <strong>of</strong> the hot-pressed ceramic from<br />

the investment, the interaction layer was removed<br />

by grit blasting with 80 mm glass beads at a pressure<br />

<strong>of</strong> 2 bar (30 psi). The bar specimens were cleaned in<br />

1% hydr<strong>of</strong>luoric acid (HF) for 30 min, grit blasted<br />

with 100 mm Al2O3 at a pressure <strong>of</strong> 2 bar, <strong>and</strong><br />

polished through 1200 grit SiC metallographic paper<br />

to a thickness <strong>of</strong> 1.2 mm. All specimens were<br />

finished with 1 mm polishing alumina (Mark V<br />

Laboratory, East Granby, CT, USA) to the final<br />

dimensions (25 £ 4 £ 1.2 mm). They were ultrasonically<br />

cleaned in distilled water <strong>and</strong> steam cleaned<br />

using distilled water. Specimens were examined for<br />

flaws using light microscopy at 30 £ (microscope<br />

model SCW30L, Fisher Scientific, Thail<strong>and</strong>). Specimens<br />

with any obvious large flaws would be<br />

eliminated, if detected. However, no major flaws<br />

were detected. The specimens were stored for 48 h<br />

in distilled water at 37 8C <strong>and</strong> then subjected to<br />

four-point flexure loading (applied along rollers<br />

at 1/3 <strong>and</strong> 2/3 <strong>of</strong> the length <strong>of</strong> the bars) at<br />

Heating rate<br />

(8C/min)<br />

Firing T<br />

(8C)<br />

Holding t<br />

(min)<br />

Vacuum T on–<strong>of</strong>f<br />

(8C)<br />

E1—IPS Empress core a (<strong>leucite</strong>-<strong>based</strong> ceramic) 700 60 1180 20 500–1180<br />

E2—IPS Empress2 core a (<strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> ceramic) 700 60 920 20 500–920<br />

ES—Experimental core a (<strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> ceramic) 700 60 910 15 500–910<br />

GV—IPS Empress2 body a (amorphous glass) 403 60 800 2 450–799<br />

a Ivoclar AG, Schaan, Liechtenstein.<br />

ARTICLE IN PRESS<br />

A. Della Bona et al.


a cross-head speed <strong>of</strong> 0.5 mm/min in a universal<br />

testing machine (Model 1125, Instron Corp., Canton,<br />

MA, USA) while immersed in distilled water.<br />

Note that any local residual stresses induced during<br />

preparation, i.e., around any flaws, were relieved<br />

by the storage under water because <strong>of</strong> slow crack<br />

growth in a conducive environment. An Isotemp<br />

Immersion Circulator (Model 730, Fisher Scientific,<br />

Pittsburgh, PA, USA), which was connected to the<br />

testing chamber, was used to maintain the water at<br />

37 8C. 11 Failure loads were recorded <strong>and</strong> the<br />

flexural strength values were calculated using the<br />

following equation:<br />

s ¼ PL=wt 2<br />

ð2Þ<br />

where P is the applied load at failure, L is the<br />

length <strong>of</strong> the outer (total) span, w is the width <strong>of</strong><br />

the specimen, <strong>and</strong> t is the thickness <strong>of</strong> the<br />

specimen. 11,12<br />

<strong>Fracture</strong> surfaces were examined using optical<br />

microscopy (OM) <strong>and</strong> scanning electron microscopy<br />

(SEM) to determine the mode <strong>of</strong> failure <strong>based</strong> on the<br />

fracture origin <strong>and</strong> fractographic principles.<br />

3,13 – 15<br />

In preparation for SEM (JSM 6400, Jeol Ltd, Tokyo,<br />

Japan) examination, the fracture surfaces were<br />

sputter-coated with gold–palladium for 3 min in a<br />

Hummer II Sputter Coater (21020, Technics Inc.,<br />

Alex<strong>and</strong>ria, VA, USA) at a current <strong>of</strong> 10 mA <strong>and</strong> a<br />

vacuum <strong>of</strong> 130 mTorr.<br />

The critical flaws were located <strong>and</strong> their size<br />

ðcÞ was determined using fractographic principles.<br />

The flaw size calculation ½c ¼ðabÞ 1=2 Š is the same as<br />

that used for an equivalent semicircular surface<br />

crack <strong>and</strong> corner crack. In the case <strong>of</strong> a corner<br />

crack, ‘c’ corresponds to the distance from the<br />

crack corner to the limit <strong>of</strong> the critical flaw-mirror<br />

region (Fig. 1).<br />

The mean crack size <strong>and</strong> strength values were<br />

used to calculate the fracture toughness ðK ICÞ via<br />

Eq. (1). The geometrical factor ðYÞ was calculated<br />

<strong>based</strong> on R<strong>and</strong>all’s 7 interpretation <strong>of</strong> Irwin’s 16<br />

work:<br />

Y ¼½0:515Q 1=2 Š 21<br />

ð3Þ<br />

Q ranged in value from 1.00 for long, shallow<br />

cracks ðY ¼ 1:94Þ to 2.46 for semicircular cracks<br />

ðY ¼ 1:24Þ: For a corner crack, the factor Y can be<br />

approximated by<br />

Y ¼ 1:12 2 2=p 1=2<br />

ð4Þ<br />

where 1.12 2 represents the surface flaw corrections<br />

<strong>and</strong> 2=p 1=2 represents the correction for a pennyshaped<br />

embedded crack, i.e. Y ¼ 1:4. 17<br />

Extra samples <strong>of</strong> each ceramic were fabricated<br />

<strong>and</strong> ground to a powder, mixed with an adhesive<br />

ARTICLE IN PRESS<br />

<strong>Fracture</strong> Behavior <strong>of</strong> Ceramics 3<br />

(collodian–amyl acetate in a ratio <strong>of</strong> 1:7), <strong>and</strong><br />

placed on a glass slide sample area. To identify the<br />

ceramic crystal phases, X-ray diffraction (XRD) was<br />

performed using a 3720 Phillips room temperature<br />

diffractometer (serial #1470, Phillips, Netherl<strong>and</strong>s)<br />

<strong>and</strong> PW1877 automated powder diffraction s<strong>of</strong>tware.<br />

A copper tube anode in a continuous scan<br />

from 3.01 to 89.978 (,3–908) with step size <strong>of</strong> 0.028<br />

every 0.4 s was used at a voltage <strong>of</strong> 40 kV, <strong>and</strong><br />

current <strong>of</strong> 20 nA. A slow scan from 10 to 408 was also<br />

used. The IPS Empress2 GV was analyzed for any<br />

crystal phase using a DRIFT (Diffuse Reflectance<br />

Infrared Fourier Transform, Spectra Tech Inc.,<br />

Shelton, CT, USA) analyzer with a Gemini Stage<br />

(Nicolet Magna 760, Nicolet Inc., Madison, WI, USA)<br />

<strong>and</strong> OMNIC 4.1b s<strong>of</strong>tware (Nicolet Inc., Madison,<br />

WI, USA). The following four powder samples were<br />

analyzed: (1) 0.5 g <strong>of</strong> KBr powder (Spectra Tech<br />

Inc., Shelton, CT, USA) for the background control;<br />

(2) 0.5 g <strong>of</strong> KBr powder mixed with 0.025 g <strong>of</strong><br />

unfired incisal GV powder; (3) 0.5 g <strong>of</strong> KBr powder<br />

mixed with 0.025 g <strong>of</strong> unfired dentin GV powder;<br />

<strong>and</strong> (4) 0.5 g <strong>of</strong> KBr powder mixed with 0.025 g <strong>of</strong><br />

fired dentin GV powder. Analysis was performed in<br />

64 scans for four wave number resolutions. The data<br />

were corrected for H2O <strong>and</strong> CO2 interferences <strong>and</strong><br />

results were recorded.<br />

Ceramic disks specimens <strong>of</strong> each material used in<br />

this part <strong>of</strong> the study (Table 1) were fabricated<br />

<strong>and</strong> polished through 1 mm alumina abrasive. The<br />

specimen thickness was measured using a caliper<br />

(Digimatic caliper, Mitutoyo Co., Kawasaki, Japan)<br />

<strong>and</strong> the weight was obtained using an analytical<br />

balance (Mettler H31, Mettler Instrument Corp.,<br />

Hightstown, NJ, USA). The density ðrÞ was obtained<br />

using the Helium Pycnometer (Accupyc 1330,<br />

Micromeritics Instrument Co., Norcross, GA, USA)<br />

after calculating the volume.<br />

The elastic modulus ðEÞ <strong>and</strong> Poisson’s ratio ðnÞ<br />

were determined by means <strong>of</strong> ultrasonic waves <strong>and</strong><br />

a computer program (ECALC <strong>and</strong> Sigview-F s<strong>of</strong>tware,<br />

Nuson Inc., Boalsburg, PA, USA) <strong>based</strong> on a<br />

set <strong>of</strong> equations that uses the time <strong>of</strong> flight,<br />

density, <strong>and</strong> thickness. Piezoelectric transducers<br />

(Ultran Laboratories Inc., Boalsburg, PA, USA) <strong>and</strong><br />

an ultrasonic pulse apparatus (Ultima 5100, Nuson<br />

Inc., Boalsburg, PA, USA) were used to determine<br />

the time <strong>of</strong> flight through the ceramic specimens.<br />

Longitudinal <strong>and</strong> shear (transverse) time <strong>of</strong> flight<br />

values were obtained <strong>and</strong> used to calculate n <strong>and</strong> E<br />

<strong>of</strong> the materials. 11 A hardness tester (Model MO<br />

Tukon Microhardness Tester, Wilson Instruments<br />

Inc., Binghampton, NY, USA) with a Vickers diamond<br />

was used to measure the hardness <strong>of</strong> these<br />

ceramic specimens. Each specimen was indented at<br />

four different locations under a load ðPÞ <strong>of</strong>


4<br />

9.8 N. The dimensions <strong>of</strong> the indentation diagonals<br />

were measured using an optical microscope with a<br />

filar eyepiece. The hardness ðHÞ values were<br />

calculated by<br />

H ¼½2P sinðu=2ÞŠ=a 2<br />

ð5Þ<br />

where P is the indentation load, u is the angle<br />

between opposite diamond faces (1368), <strong>and</strong> a is the<br />

mean diagonal length <strong>of</strong> the indentation.<br />

Flexural strength <strong>and</strong> K IC data were analyzed<br />

statistically using one-way ANOVA <strong>and</strong> Tukey’s HSD<br />

test. Weibull regression analysis was performed on<br />

the strength data as previously described 12 <strong>and</strong> as<br />

presented in a previous report. 11<br />

Results<br />

The mean values <strong>of</strong> density ðrÞ; elastic modulus ðEÞ;<br />

Poisson’s ratio ðnÞ; <strong>and</strong> Vickers hardness ðHÞ <strong>of</strong> the<br />

<strong>ceramics</strong> used in this study are summarized in<br />

Table 2.<br />

XRD analyses showed that high-exp<strong>and</strong>ing mineral,<br />

<strong>leucite</strong> (K2O·Al2O3·4SiO2), is the crystal phase<br />

in the IPS Empress core ceramic (E1), <strong>and</strong> the<br />

lithium <strong>disilicate</strong> (Li2O·2SiO2) is the crystal phase in<br />

the IPS Empress 2 (E2) <strong>and</strong> Experimental (ES) core<br />

<strong>ceramics</strong>. These crystal phases are shown in Fig. 2.<br />

The length <strong>of</strong> the elongated Li2Si2O5 crystals ranged<br />

from 0.5 to 4 mm for E2 <strong>and</strong> from 0.5 to 2 mm for ES.<br />

XRD analysis <strong>of</strong> the IPS Empress2 GV produced a<br />

typical amorphous glass plot. The absence <strong>of</strong> any<br />

crystal phase was confirmed by FTIR analysis.<br />

Representative photomicrographs <strong>of</strong> the <strong>ceramics</strong><br />

microstructure are presented in Fig. 2. As<br />

expected, there is a positive correlation between<br />

the elastic modulus ðEÞ; hardness ðHÞ; flexure<br />

strength ðsÞ; <strong>and</strong> fracture toughness ðK ICÞ for the<br />

four <strong>ceramics</strong> studied (E1, E2, ES, <strong>and</strong> GV).<br />

One-way ANOVA <strong>and</strong> Tukey’s test subsets<br />

revealed significant statistical differences for the<br />

mean s <strong>and</strong> K IC values between E1 <strong>and</strong> GV <strong>ceramics</strong><br />

ðp , 0:05Þ: The two <strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> <strong>ceramics</strong><br />

(E2 <strong>and</strong> ES) showed greater mean s <strong>and</strong> K IC values<br />

Table 2 Mean values <strong>of</strong> density ðrÞ; elastic modulus ðEÞ;<br />

Poisson’s ratio ðnÞ; <strong>and</strong> Vickers hardness ðHÞ <strong>of</strong> Empress<br />

<strong>ceramics</strong>.<br />

Material properties E1 E2 ES GV<br />

Density (r, g/cm 3 ) 2.47 2.51 2.56 2.53<br />

Elastic modulus (E, GPa) 86 96 96 65<br />

Poisson’s ratio ðnÞ 0.27 0.26 0.24 0.23<br />

Vickers hardness (H, GPa) 5.9 6.3 6.3 5.4<br />

ARTICLE IN PRESS<br />

compared with the <strong>leucite</strong>-<strong>based</strong> ceramic (E1) <strong>and</strong><br />

the GV (Table 3) <strong>and</strong> the differences in mean values<br />

were statistically significant ðp # 0:05Þ:<br />

All fractures initiated along the tensile surfaces<br />

between the inner load points. Corner flaws (CF),<br />

either cracks or voids, were identified as the<br />

fracture origin in 20–30% <strong>of</strong> the specimens (Fig. 3,<br />

Table 3).<br />

Discussion<br />

A. Della Bona et al.<br />

Structural failure associated with an applied load<br />

occurs when the induced stress is greater than the<br />

strength <strong>of</strong> the material. The failure location is<br />

dependent on the size <strong>of</strong> the initiating crack <strong>and</strong> the<br />

fracture toughness. 9 Usually, fracture <strong>of</strong> the ceramic<br />

originates from the most severe flaw. The<br />

large number <strong>of</strong> pre-existing ceramic cracks <strong>of</strong><br />

differing sizes, coupled with a low fracture toughness,<br />

limits the strength <strong>of</strong> <strong>ceramics</strong> <strong>and</strong> causes a<br />

large variability in strength <strong>and</strong> the time dependence<br />

<strong>of</strong> strength. The size <strong>and</strong> spatial distribution<br />

<strong>of</strong> flaws justify the necessity <strong>of</strong> a statistical<br />

approach to failure analysis 18 to determine the<br />

reliability <strong>of</strong> <strong>ceramics</strong>. 11,19<br />

Investigations <strong>of</strong> clinically failed all-ceramic<br />

restorations have shown that the fracture origin is<br />

typically located at the inner tensile surface.<br />

20 – 22<br />

However, for strength tests to yield relevant<br />

service information, the test <strong>and</strong> service environments<br />

must be similar <strong>and</strong> the strength-controlling<br />

flaw population must approximate that responsible<br />

for failure in clinical service. 19 The distribution <strong>of</strong><br />

flexural strengths under wet environmental conditions<br />

is a potential indication <strong>of</strong> ceramic performance.<br />

The greater mean s <strong>and</strong> K IC values <strong>of</strong> E2<br />

<strong>and</strong> ES core <strong>ceramics</strong> are indicative <strong>of</strong> potentially<br />

improved structural performance. However, resistance<br />

to stress corrosion processes <strong>and</strong> a small range<br />

<strong>of</strong> strength values are also factors in structural<br />

reliability. Thus, a combination <strong>of</strong> good fracture<br />

toughness, excellent resistance to stress corrosion,<br />

<strong>and</strong> resistance to loading are required for structural<br />

reliability. If one <strong>of</strong> these elements is missing, then<br />

there may be a false sense <strong>of</strong> security in designing<br />

ceramic prostheses made from these materials. For<br />

example, in the case <strong>of</strong> the Group GV, the Weibull<br />

modulus, m; is twice as large as that for each <strong>of</strong> the<br />

other groups <strong>and</strong> the flaw size is smaller than the<br />

flaw sizes <strong>of</strong> the other groups. This appears to be a<br />

positive result. However, because <strong>of</strong> the low value<br />

<strong>of</strong> fracture toughness, the strength is much less<br />

than that <strong>of</strong> the other groups <strong>and</strong> this, most likely,<br />

would not be a reliable material.


Under certain service <strong>and</strong>/or environmental<br />

conditions, stable crack extension or slow crack<br />

growth can occur at stress intensities that are less<br />

than the critical value, K C: In this study all fracture<br />

surfaces were produced from specimens immersed<br />

in 37 8C distilled water. The loading rate used<br />

limited, but did not eliminate, slow crack growth,<br />

when tested in water. 11 Thus, the strengths<br />

reported are less, but more conservative than<br />

those expected for tests conducted in air. Yet<br />

the calculated fracture toughness does not depend<br />

on the crack growth since the flaw size was<br />

measured. The strength values are correlated to<br />

the flaw size to calculate the value <strong>of</strong> toughness.<br />

ARTICLE IN PRESS<br />

<strong>Fracture</strong> Behavior <strong>of</strong> Ceramics 5<br />

Figure 2 SEM micrographs <strong>of</strong> lightly etched <strong>ceramics</strong>. (A) IPS Empress 2 GV revealing light <strong>and</strong> dark glass phases<br />

containing very similar compositions. (B) XRD analysis showed the presence <strong>of</strong> <strong>leucite</strong> crystals ( p ) in a glassy phase <strong>of</strong><br />

IPS Empress core ceramic (E1). XRD analysis showed the presence <strong>of</strong> lithium <strong>disilicate</strong> crystals ( p ) in a glassy phase <strong>of</strong><br />

(C) IPS Empress 2 core ceramic (E2) <strong>and</strong> (D) Experimental <strong>lithia</strong> <strong>disilicate</strong>-<strong>based</strong> core ceramic (ES).<br />

Previous reports on the fracture toughness using the<br />

single edge notched beam (SENB) method performed<br />

at room conditions showed KIC values<br />

comparable to those presented in this study.<br />

23 – 25<br />

<strong>Fracture</strong> surface analysis is well established as a<br />

means <strong>of</strong> failure analysis in the field <strong>of</strong> glasses <strong>and</strong><br />

<strong>ceramics</strong>. 3 It has been recognized as a powerful<br />

analytical tool in dentistry. 20,22 <strong>Fracture</strong> in glass<br />

<strong>and</strong> <strong>ceramics</strong> occurs when pre-existing cracks<br />

propagate under excessive tensile stresses.<br />

These cracks can be produced by mechanical<br />

means, e.g. grinding or polishing, by processing,<br />

or by intrinsic defects, e.g. imperfections in the<br />

structure. Fractographic analysis identified the type<br />

Table 3 Mean values <strong>and</strong> st<strong>and</strong>ard deviation (SD) <strong>of</strong> flexural strength ðs fÞ; critical crack size ðcÞ; fracture toughness ðK ICÞ;<br />

characteristic strength ðs 0Þ; Weibull modulus ðmÞ; <strong>and</strong> fracture origins for the four ceramic groups.<br />

Groups s f (SD) (MPa) s 0 (MPa) m c (SD) (mm) K IC (SD) (MPa m 1/2 ) <strong>Fracture</strong> origin a (%)<br />

E1 85 (15) B b<br />

91.3 5 110 (62) 1.3 (0.3) B b<br />

SF (70); CF (30)<br />

E2 215 (40) A 231 5 140 (56) 3.4 (0.6) A SF (70); CF (30)<br />

ES 239 (36) A 252 7 110 (25) 3.1 (0.4) A SF (80); CF (20)<br />

GV 63.8 (5.8) C 65.7 14 95 (35) 0.8 (0.1) C SF (80); CF (20)<br />

a SF, surface flaw (void or crack); CF, corner flaw (void or crack).<br />

b The values in each column with same letters do not differ significantly ðp , 0:05Þ:


6<br />

Figure 3 SEM micrographs <strong>of</strong> ceramic fracture surfaces showing representative critical flaws outlined by white arrows.<br />

(A) <strong>Fracture</strong> surface <strong>of</strong> E1; line from flaw corner, c ¼ 84 mm (500 £ ). (B) <strong>Fracture</strong> surface <strong>of</strong> E2; measured line<br />

represents the semiminor axis, a ¼ 44 mm (500 £ ). (C) <strong>Fracture</strong> surface <strong>of</strong> ES; measured line represents the semiminor<br />

axis, a ¼ 35 mm (600 £ ). (D) <strong>Fracture</strong> surface <strong>of</strong> GV; note the tailed fracture markings (top right) pointing toward the<br />

crack origin; measured line represents the semiminor axis, a ¼ 55 mm (500 £ ).<br />

<strong>of</strong> origins <strong>and</strong> confirmed the presence <strong>of</strong> characteristic<br />

markings <strong>of</strong> the fracture process (Fig. 3).<br />

Ceramic specimens tested in bending are very<br />

sensitive to edge or surface machining damage. 19<br />

Fractographic analysis has shown that all failures<br />

started from either a surface (Fig. 3B–D) or a corner<br />

flaw (Fig. 3A) located along the tensile surface <strong>of</strong><br />

the specimens. These results are consistent with a<br />

previous report, which suggested that surface<br />

failures started only at the tensile side <strong>of</strong> specimens<br />

tested in four-point bending. 26 However, the<br />

number <strong>of</strong> fractures originating from corner flaws<br />

(20–30%) in this study suggests that careful manual<br />

chamfer or rounding <strong>of</strong> specimen edges may<br />

improve the reproducibility <strong>of</strong> test results produced<br />

through ISO st<strong>and</strong>ard 6872 for dental <strong>ceramics</strong>.<br />

Although machine rounding <strong>of</strong> edges can produce<br />

even more defects, careful manual rounding <strong>of</strong> the<br />

edges is preferred, by some investigators, to<br />

machining 908 corners.<br />

The Weibull modulus ðmÞ is a measure <strong>of</strong> the<br />

distribution <strong>of</strong> critical flaws. As the Weibull moduli<br />

are similar for three <strong>of</strong> the four <strong>ceramics</strong>, <strong>and</strong> the<br />

crack sizes ðcÞ are comparable (Table 3),<br />

ARTICLE IN PRESS<br />

A. Della Bona et al.<br />

the differences in mean strength cannot be<br />

explained by a difference in crack sizes. In fact,<br />

E2 has the largest mean crack sizes, <strong>and</strong> yet, has<br />

one <strong>of</strong> the largest mean strength values. Since the<br />

fracture toughness is a constant (Eq. (1)), a large<br />

crack size would result in a low strength value for<br />

the same toughness. Thus, the differences in mean<br />

strength can only be explained by differences in<br />

fracture toughness that are related to processing<br />

<strong>and</strong> composition variables. The GV contains<br />

relatively few, if any, crystals <strong>and</strong> would be<br />

expected to have a toughness value the same as,<br />

or slightly greater than those <strong>of</strong> silicate glasses as<br />

shown in Table 3. The E1 core ceramic has crystals<br />

dispersed in a glassy matrix. However, the volume<br />

fraction <strong>of</strong> crystals is very small (Fig. 2) compared<br />

with those <strong>of</strong> E2 <strong>and</strong> ES <strong>ceramics</strong> <strong>and</strong> is expected to<br />

have a toughness value that lies between that <strong>of</strong> the<br />

GV <strong>and</strong> those <strong>of</strong> the E2 <strong>and</strong> ES <strong>ceramics</strong>. The fine<br />

dispersion <strong>of</strong> crystals in the E2 <strong>and</strong> ES <strong>ceramics</strong><br />

leads to the increased toughness values observed<br />

relative to either the GV or the E1 ceramic core<br />

material. The differences in the crystal sizes<br />

between E2 <strong>and</strong> ES <strong>ceramics</strong> are not great enough


within the sampling error to determine if these<br />

differences can affect the toughness values.<br />

However, the amount <strong>and</strong> sizes <strong>of</strong> crystals are<br />

enough to increase the magnitude <strong>of</strong> toughness to a<br />

greater level compared with GV <strong>and</strong> E1. Thus, the<br />

strength values are proportional to the toughness<br />

values as stated previously.<br />

Acknowledgements<br />

This paper is <strong>based</strong> on a dissertation submitted to<br />

the graduate faculty, University <strong>of</strong> Florida, in<br />

partial fulfillment <strong>of</strong> the requirements for the PhD<br />

degree.<br />

This study was partially supported by CNPq-Brazil,<br />

grant 300659/03-2, <strong>and</strong> NIH-NIDCR grant DE06672.<br />

The authors thank Dr C. Shen for performing the<br />

statistical analysis <strong>and</strong> Mr Robert B. Lee for his<br />

assistance in processing the <strong>ceramics</strong>. We also<br />

thank Ivoclar AG (Liechtenstein) for providing the<br />

ceramic materials used in this study.<br />

References<br />

1. Griffith AA. The phenomena <strong>of</strong> rupture <strong>and</strong> flow in solids.<br />

Philos Trans R Soc 1920;221:163—98.<br />

2. West JK, Mecholsky Jr JJ, Hench LL. The application <strong>of</strong><br />

fractal <strong>and</strong> quantum geometry to brittle fracture. J Non-<br />

Cryst Solids 1999;260:99—108.<br />

3. Mecholsky Jr JJ. Fractography: determining the sites <strong>of</strong><br />

fracture initiation. Dent Mater 1995;11:113—6.<br />

4. Anstis GR, Chantikul P, Lawn BR, Marshall DB. A critical<br />

evaluation <strong>of</strong> indentation techniques for measuring fracture<br />

toughness: I. Direct crack measurements. J Am Ceram Soc<br />

1981;64:533—8.<br />

5. Chantikul P, Anstis GR, Lawn BR, Marshall DB. A critical<br />

evaluation <strong>of</strong> indentation techniques for measuring fracture<br />

toughness: II. Strength method. J Am Ceram Soc 1981;64:<br />

539—43.<br />

6. Scherrer SS, Kelly JR, Quinn GD, Xu K. <strong>Fracture</strong> toughness<br />

ðK ICÞ <strong>of</strong> a dental porcelain determined by fractographic<br />

analysis. Dent Mater 1999;15:342—8.<br />

7. R<strong>and</strong>all PN. Plain strain crack toughness testing <strong>of</strong> high<br />

strength metallic materials. In: Brown WF Jr, Strawley JE,<br />

editors. ASTM STP 410. Philadelphia: American Society for<br />

Testing <strong>and</strong> Materials; 1966. p. 88—126.<br />

8. Mecholsky JrJJ, Freiman SW, Rice RW. Fractographic analysis<br />

<strong>of</strong> <strong>ceramics</strong>. In: Strauss BM, Cullen WH, editors. Fractography<br />

in failure analysis, ASTM STP 645. Philadelphia: American<br />

Society for Testing <strong>and</strong> Materials; 1978. p. 363—79.<br />

ARTICLE IN PRESS<br />

<strong>Fracture</strong> Behavior <strong>of</strong> Ceramics 7<br />

9. Mecholsky Jr JJ. <strong>Fracture</strong> mechanics principles. Dent Mater<br />

1995;11:111—2.<br />

10. Anusavice KJ. Phillips’ science <strong>of</strong> dental materials, 11th ed.<br />

St. Louis, MO: Saunders, an imprint <strong>of</strong> Elsevier Science;<br />

2003.<br />

11. Della Bona A, Anusavice KJ, DeH<strong>of</strong>f PH. Weibull analysis <strong>and</strong><br />

flexural strength <strong>of</strong> hot-pressed core <strong>and</strong> veneered ceramic<br />

structures. Dent Mater 2003;19:662—9.<br />

12. Ban S, Anusavice KJ. Influence <strong>of</strong> test method on failure<br />

stress <strong>of</strong> brittle dental materials. J Dent Res 1990;69:<br />

1791—9.<br />

13. Rice RW. Ceramic fracture features, observations, mechanisms,<br />

<strong>and</strong> uses. In: Mecholsky JJ Jr, Powell SR Jr, editors.<br />

Fractography <strong>of</strong> ceramic <strong>and</strong> metal failures, ASTM STP 827.<br />

Philadelphia: American Society for Testing <strong>and</strong> Materials;<br />

1984. p. 5—103.<br />

14. Fréchette VD, Advances in <strong>ceramics</strong>: failure analysis <strong>of</strong><br />

brittle materials, vol. 28. Westerville, OH: American<br />

Ceramic Society; 1990.<br />

15. Della Bona A, Anusavice KJ, Mecholsky Jr JJ. Failure analysis<br />

<strong>of</strong> resin composite bonded to ceramic. Dent Mater 2003;19:<br />

693—9.<br />

16. Irwin GR. Analysis <strong>of</strong> stresses <strong>and</strong> strains near the end <strong>of</strong><br />

crack transversing a plate. J Appl Mech 1957;24:361—4.<br />

17. Hertzberg RW. Deformation <strong>and</strong> fracture mechanics <strong>of</strong><br />

engineering materials, 4th ed. New York: Wiley; 1996.<br />

18. Weibull W. A statistical theory <strong>of</strong> the strength <strong>of</strong> materials.<br />

Ing Vetensk Akad Proc 1939;151:1—45.<br />

19. Ritter JE. Critique <strong>of</strong> test methods for lifetime predictions.<br />

Dent Mater 1995;11:147—51.<br />

20. Kelly JR, Campbell SD, Bowen HK. <strong>Fracture</strong>-surface analysis<br />

<strong>of</strong> dental <strong>ceramics</strong>. J Prosthet Dent 1989;62:536—41.<br />

21. Kelly JR, Giordano R, Pober R, Cima MJ. <strong>Fracture</strong> surface<br />

analysis <strong>of</strong> dental <strong>ceramics</strong>: clinically failed restorations. Int<br />

J Prosthodont 1990;3:430—40.<br />

22. Thompson JY, Anusavice KJ, Naman A, Morris HF. <strong>Fracture</strong><br />

surface characterization <strong>of</strong> clinically failed all-ceramic<br />

crowns. J Dent Res 1994;73:1824—32.<br />

23. Höl<strong>and</strong> W, Schweiger M, Frank M, Rheinberger V. A<br />

comparison <strong>of</strong> the microstructure <strong>and</strong> properties <strong>of</strong> the IPS<br />

Empress 2 <strong>and</strong> the IPS Empress glass—<strong>ceramics</strong>. J Biomed<br />

Mater Res 2000;53:297—303.<br />

24. Drummond JL, King TJ, Bapna MS, Koperski RD. Mechanical<br />

property evaluation <strong>of</strong> pressable restorative <strong>ceramics</strong>. Dent<br />

Mater 2000;16:226—33.<br />

25. Fischer H, Marx R. <strong>Fracture</strong> toughness <strong>of</strong> dental <strong>ceramics</strong>:<br />

comparison <strong>of</strong> bending <strong>and</strong> indentation methods. Dent Mater<br />

2002;18:12—19.<br />

26. Thompson GA. Influence <strong>of</strong> relative layer height <strong>and</strong><br />

testing method on the failure mode <strong>and</strong> origin in a<br />

bilayered dental ceramic composite. Dent Mater 2000;16:<br />

235—43.<br />

27. Mecholsky JrJJ. Quantitative fracture surface analysis <strong>of</strong><br />

glass materials. In: Simmons CJ, El-Bayoumi O, editors.<br />

Experimental techniques <strong>of</strong> glass science. Westerville, OH:<br />

American Ceramic Society; 1993. p. 483—520.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!