Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Flexural strength and fracture toughness of dental core ceramics

Handan Yilmaz, DDS, PhD,a Cemal Aydin, DDS, PhD,b and Basak E. Gul, DDSc Faculty of Dentistry, Gazi University, Ankara, Turkey
Statement of problem. Many different strengthened all-ceramic core materials are available. In vitro study of their mechanical properties, such as flexural strength and fracture toughness, is necessary before they are used clinically. Purpose. The purpose of this study was to evaluate and compare the mechanical properties of 6 commonly used allceramic core materials using biaxial flexural strength and indentation fracture toughness tests. Material and methods. Specimens of 6 ceramic core materials (Finesse, Cergo, IPS Empress, In-Ceram Alumina, In-Ceram Zirconia, and Cercon Zirconia) were fabricated (n=25) with a diameter of 15 mm and width of 1.2 0.2 mm. For each group, the specimens were tested to compare their biaxial flexural strength (piston on 3 balls) (n=15), Weibull modulus, and indentation fracture toughness (n=10) (IF method). The data were analyzed with 1-way ANOVA test (=.05). The Tamhane multiple comparison test was used for post hoc analysis. Results. Mean (SD) of biaxial flexural strength values (MPa) and Weibull modulus (m) results were: Finesse (F): 88.04 (31.61), m=3.17; Cergo (C): 94.97 (13.62), m=7.94; IPS Empress (E): 101.18 (13.49), m=10.13; In-Ceram Alumina (ICA): 341.80 (61.13), m=6.96; In-Ceram Zirconia (ICZ): 541.80 (61.10), m=10.17; and Cercon Zirconia (CZ): 1140.89 (121.33), m=13.26. The indentation fracture toughness results showed that there were significant differences between the tested ceramics. The highest fracture toughness values (MPa x m0.5) were obtained with the zirconiabased ceramic core materials. Conclusions. Significant differences were found in strength and toughness values of the materials evaluated. Cercon Zirconia core material showed high values of biaxial flexural strength and indentation fracture toughness when compared to the other ceramics studied. (J Prosthet Dent 2007; 98: 120-128.)

Clinical Implications

The results of this in vitro investigation provide the clinician with data regarding strength and fracture toughness of a broad range of currently available core ceramics. The properties of core materials can affect the clinical performance of all-ceramic restorations.
Several all-ceramic materials and processing techniques have been introduced in the past decade. The most commonly used systems can be classified according to the laboratory processing procedure (pressable, slipcasting, milling, or sintering)1 and chemical composition (feldspar: high leucite and low leucite; glass ceramic: lithium disilicate and mica; core reinforced: alumina, magnesia, and zirconia).2,3 All-ceramic processing mechanisms include the formation of surface layers containing compressive stresses,4 controlled crystallization5 (Dicor; Dentsply Intl, York, Pa, and Cerapearl; Kyocera, San Diego, Ca-

This study was supported by the Scientific Research Project of the Rectorship of Gazi University, Ankara, Turkey. Professor, Department of Prosthodontics. Professor, Department of Prosthodontics. c Research Assistant, Department of Prosthodontics.
a b

The Journal of Prosthetic Dentistry

Yilmaz et al

August 2007
lif ), dispersion strengthening with alumina6 (Hi-Ceram; Vita Zahnfabrik, Bad Sackingen, Germany), and dispersion strengthening with leucite7 (Optec; Jeneric/Pentron, Wallingford, Conn and IPS Empress; Ivoclar Vivadent, Schaan, Liechtenstein). Glass infiltrated sintered alumina8 (In-Ceram; Vita Zahnfabrik) and high-purity alumina9 (Procera AllCeram; Nobel Biocare AB, Goteborg, Sweden) are strengthened with structures consisting of 2 interpenetrating network phases and high-density grains, respectively. Glass-infiltrated alumina/ zirconium oxide10,11 (In-Ceram Zirconia; Vita Zahnfabrik) and transformation-toughened polycrystalline zirconia11 (Cercon; Dentsply DeguDent, Lava; 3M ESPE, St. Paul, Minn and YZ; Vita Zahnfabrik) are other all-ceramic processing mechanisms. The mechanical properties of highperformance alumina- and zirconiabased ceramics make them attractive as potential materials for all-ceramic restorations in high stress-bearing areas.12 In-Ceram Alumina (Vita Zahnfabrik) is an alumina-reinforced ceramic that has been used as a core material for crowns and anterior 3unit FPDs since the early 1990s.13 InCeram Zirconia (Vita Zahnfabrik) was developed by adding 33 wt% of partially stabilized zirconia to the initial compound, In-Ceram Alumina (Vita Zahnfabrik), to provide a stronger and tougher core material.14 The Cercon Zirconia (Dentsply DeguDent) system is a partially stabilized zirconia ceramic. The introduction of partially stabilized zirconia ceramics has increased the reliability of all-ceramic fixed prosthodontic restorations, possibly due to the occurrence of a transformation-toughening mechanism.15,16 The process of transformation induces compressive stresses in the crack wake and shear strains that counteract the tensile field generated at the tip of a propagating crack.17-19 The addition of a stabilizer, namely yttria, stabilizes the transforming zirconia system in the tetragonal phase and retains the layer of compressive stresses, resulting in the formation of a yttria-stabilized tetragonal zirconia polycrystalline (YTZP) ceramic and inducing high fracture strength and fracture toughness characteristic of Y-TZP ceramics.3,20,21 The Cercon system (Dentsply DeguDent) requires conventional waxing techniques for designing the Y-TZPbased infrastructure.22 The zirconia frameworks are machined by the Cercon-smart ceramics system using Cercon blanks (Degudent, Hanau, Germany) made of 3 mol% yttria-stabilized zirconia (tetragonal zirconia polycrystals, TZP).23 First, the shape of the wax pattern of the framework was digitized by laser scanning. Then the enlarged shape of the framework was milled out of a homogeneous porous prefabricated Cercon blank (Degudent) and sintered to full density for 2 hours at 1350oC.23 During firing, the framework shrinks to the desired definitive dimensions.22 The milling process is faster and the abrasion of hardware is less than the milling from a fully sintered blank.22,24 The average masticatory force has been reported to vary between 11 and 150 N, whereas force peaks were reported to be 200 N in the anterior, 350 N in the posterior, and 1000 N in subjects with parafunctional habits.25-27 A dental restoration is routinely subjected to masticatory loads in excess of 200 N, while the most extreme clenching and bruxism forces in patients can exert loads of up to 1221 N with masticatory forces routinely experienced by fixed prosthodontic restorations ranging from 150 to 665 N.28,29 Restorations in the posterior areas should withstand these forces. Thus, it is important to evaluate the mechanical properties of ceramic core materials in vitro. Flexural strength and fracture toughness characterize the responses of materials, such as brittle dental ceramics, to loading forces and crack propagation, respectively.30 Strength is an important mechanical property that can assist in predicting the performance of brittle materials.31 Strength is defined as the ultimate stress that is necessary to cause fracture or plastic deformation and is strongly affected by the size of flaws and defects present on the surface of a tested material.32 However, microcracks and defects that grow inherently during thermal and mechanical processes can significantly influence strength measurement.33 Several studies have used 3- and 4-point flexural strength tests to examine the mechanical strength of brittle materials such as ceramics.14,15,34-38 In other studies, the biaxial flexural strength test method has been used to predict the performance of all-ceramic materials.10,30,39-43 The 4-point and biaxial flexure tests develop lower levels of shear in the test section as compared to the 3point flexure tests.30 The stress state in the 4-point and biaxial tests are, therefore, closer to pure bending.30 Furthermore, the biaxial tests are less sensitive to the edge effects than the 3- or 4-point flexural tests and are less sensitive to the surface imperfections resulting from specimen preparation.30,44 In addition, the biaxial test probes for the largest flaws oriented over a wider range of angles, while the 3- and 4-point bend tests are most sensitive to flaws nearly perpendicular to the beam axis of the specimen.30 It is impossible to eliminate all flaws, and because fracture often initiates at the edges, large variations in strength have been recorded.44 The biaxial flexure test eliminates the effect of edges because they are not directly loaded. Therefore, the biaxial test should produce less variation in data for strength determination.39 In the biaxial flexural test, a disk-shaped specimen is supported from below by 3 ball bearings distributed in a circular pattern. The load is applied from above by use of a piston concentric with the supporting ball bearings (Fig. 1). The variability of the strength and, consequently, the homogeneity of the materials, can then be appraised through a calculation of the Weibull modulus (m), which is related to the flaw size distribution.10 A high m value

121

Yilmaz et al

122

Volume 98 Issue 2
MATERIAL AND METHODS
Six different all-ceramic core materials were selected for this study (Table I). CZ (Cercon Zirconia; Dentsply DeguDent) and ICZ (Vita Zahnfabrik) specimens were provided by the manufacturers. The other specimens were provided by authorized dental laboratories and prepared according to the manufacturers recommendations. Strength was measured with the biaxial flexural strength test (piston on 3 balls) as described in the ISO standard 6872 for dental ceramics.52 Fifteen (n = 90) disk-shaped specimens for each material with a diameter of 15 mm and width of 1.2 0.2 mm were prepared. The parallelism and the flatness of opposing surfaces of each specimen were verified with a micrometer (CD-15CW; Mitutoyo Corp, Kawasaki, Japan) to a tolerance within 0.05 mm. In the biaxial flexural strength test, the disk specimens were positioned on 3 balls and loaded in the middle. A tension-compression test machine, manufactured at the METU (Middle East Technical University, Department of Metallurgy, Ankara, Turkey), with a crosshead speed of 0.15 mm/min, was used. To support the test specimen, 3 hardened steel balls with a diameter of 3.2 mm (Turkish Aerospace Industries Inc, Ankara, Turkey) were positioned 120 degrees apart on a support circle with a diameter of 10 mm (Fig. 1). The diskshaped specimens were positioned concentrically on these supports and the load was applied at the center of the specimen with a flat punch 1.4 mm in diameter (Fig. 1). The load at the point of fracture was recorded, and the biaxial flexural strength for each specimen was calculated with the following equation: 52 S= -0.2387 P (X-Y)/d2 where S is the maximum center tensile stress (MPa) (the flexural strength at fracture); P is the total load causing fracture (N):
X= (1+)ln(r2/r3)2 + [(1-)/2](r2/r3)2 : Y= (1+)[1+ln(r1/r3)2]+(1-)(r1/r3)2:

1 Biaxial flexural strength test machine with specimen.

indicates a close grouping of fracture stress values, while a low m value suggests a wide distribution with a long tail at low stress levels.45 The Weibull m values are subject to specimen size, finish, and test environment because of the possible influences on subcritical flaw growth.46 Hardness is one of the most frequently measured properties of a ceramic. Its value helps to characterize resistance to deformation, densification, and fracture. Hardness is usually measured with conventional microhardness instruments such as Knoop or Vickers diamond indenters. These instruments make impressions for which the diagonal size is measured with an attached optical microscope.47 Fracture toughness is defined as the critical stress intensity level at which a given flaw starts growing, and indicates the ability of a material to resist rapid crack propagation and its consequent catastrophic failure.48 One important feature of fracture toughness is its ability to indicate the serviceability of a material intraorally.49 The application of the indentation fracture toughness technique (IF) in studying the behavior and properties of brittle materials is specifically appropriate because only small di-

mensional specimens are required, and the crack growth parameter is similar to those cracks expected in clinical conditions.50 Using this technique, an approximate determination of relative fracture toughness is assessed by introducing flaws of controlled size, shape, and location, followed by direct measurement of the radial cracks. It is a simple technique and requires only a few specimens for testing, however, it results in high dispersion because of difficulties inherent with the accurate measurement of crack length and subcritical slow crack growth.48 Despite the fact that mechanical properties are not enough to make a valid prediction of all-ceramic core material performance or long-term success, it is hypothesized that materials with improved fracture toughness should have better clinical success and longevity.51 The purpose of this study was to evaluate and compare the mechanical properties of 6 different, commonly used, all-ceramic core materials using a biaxial flexural strength test and an indentation fracture toughness test.

The Journal of Prosthetic Dentistry

Yilmaz et al

August 2007
in which is Poissons ratio (if the value for the ceramic concerned is not known, Poissons ratio of 0.25 is used); r1 is the radius of the support circle, r2 is the radius of the loaded area (mm), r3 is the radius of the specimen (mm), d is the specimen thickness at the origin of fracture (mm). For this study, = 0.25, r1 = 0.5 mm, and r2 = 0.7 mm. Data were analyzed with a 1-way analysis of variance (ANOVA) test (=.05), followed by the Tamhane multiple comparison test (=.05). In addition, a Weibull regression analysis was performed on the strength data to determine the Weibull modulus (m value) and the strength levels at a 1% and 5% probability of failure for assessing the reliability of each material. The description of the Weibull distribution is given by the formula41: P () = 1 exp (- / o) m Where, P() is the fracture probability, is the fracture strength, o is the characteristic strength at the fracture probability of 63.21%, and m is the Weibull modulus, which is the slope of the ln (ln 1/1-P) vs in plots. Fracture toughness was determined by the indentation technique proposed by Anstis et al.53 Indentations were performed with a standard Vickers diamond pyramid on a hardness testing machine (Vickers; Fritz Heckert, Leipzig, Germany). Each specimen was indented at 3 equidistant locations (n=10, 3 readings for each specimen), such that each was approximately 4.5 mm away from the disk center, with an angle of 120 degrees between them. The crack length, c (as measured from the center of the indent), should be at least equal to or greater than the diagonal length (2a), according to Anstis et al53 (Fig. 2). The standard Vickers loads of the hardness instrument (Fritz Heckert) were evaluated first to determine the optimum load to meet the criterion of

123

Table I. All-ceramic core materials evaluated


Product
Finesse (F) Cergo (C) IPS Empress (E) In-Ceram Alumina (ICA) In-Ceram Zirconia (ICZ) Cercon Zirconia (CZ) Dentsply DeguDent GmbH 555001

Manufacturer
Dentsply Ceramco, York, Pa Dentsply DeguDent GmbH, Hanau-Wolfgang Germany Ivoclar Vivadent, Schaan, Lichtenstein Vita Zahnfabrick, Bad Sackingen, Germany Vita Zahnfabrick

Batch Number
413001 413264 6500 0001 CE0128 TC1 HSORALV2 HSORZV2

Technique
Leucite-reinforced glass ceramic Pressable all-ceramic Leucite-reinforced glass ceramic Alumina-based glass infiltrated ceramic Zirconia-reinforced glass infiltrated ceramic Yttria-stabilized zirconia

2 Schematic diagram of Vickers Indenter and measurement of indentation cracking; c = crack length (from center of indent), a = length of half diagonal, c 2a.

Yilmaz et al

124
c/a 2. Therefore, it was determined by trial and error that a load of 98.1 N for specimen groups of C, F, and E, 490 N for ICA and ICZ, and 588 N for the CZ group, should be used to produce a c/a ratio of 2. Thirty acceptable readings were recorded for each material under the microscope of the indentation machine and the fracture toughness was calculated as follows53: K= 0.016 x (E/H)1/2 x P/c3/2 where, H=P/2a2, in which K is the fracture toughness of the material (MPa x m1/2), E is the elastic modulus, P is the load applied (N), a is the indent half diagonal (m), c is the crack length measured from the center of the indent (m). Differences between materials were analyzed with a 1-way ANOVA at =.05. In order to compare specimen properties, the Tamhane multiple comparison test was used.

Volume 98 Issue 2

1400

Biaxial Flexural Strength (MPa)

1200 1000 800 600 400 200 F C E ICA ICZ CZ

Core Materials 3 Mean values of biaxial flexural strength (MPa). Lines at top of bars represent standard error.

Table II. One-way ANOVA statistical results of biaxial flexural strength test
Sum of Squares
Between groups 12,760,026.7

RESULTS
Biaxial flexural strength values for the 6 all-ceramic core materials are presented in Figure 3. Core material F had the lowest mean (SD) strength value, 88.04 (31.61) MPa, and CZ had the highest strength value, 1140.89 (121.33) MPa. The other mean (SD) strength values in MPa were: 94.97 (13.62) for C, 101.18 (13.49) for E, 341.80 (61.13) for ICA, and 541.80 (61.10) for ICZ. The results of the present study revealed that biaxial flexure strengths (piston on 3 ball test) of zirconia-based (CZ and ICZ) and alumina-based (ICA) ceramic systems are significantly different from other tested ceramic systems (F, C, E, and ICA, F = 649.917, P<.001) (Table II). The mean flexural strength (1140.89 MPa) for the CZ in this study is significantly higher (P<.001) than the mean flexural strength (541.80 MPa) recorded for the other zirconia-based ceramic system, ICZ. There were no significant differences between F, C, and E core materials. Significant differences (P<.001) were recorded in the biaxial flexural strength values be-

df
5

Mean Square
2,552,005.3

649.9 Within groups 329,843.7 84 3,926.7

<.001

Table III. Weibull statistical results


Brand
Finesse Cergo IPS Empress In-Ceram Alumina In-Ceram Zirconia Cercon Zirconia

m Value
3.17 7.94 10.13 6.96 10.17 13.26

0.05
38.7 69.5 79.5 239.5 424.4 951.0

0.01
23.1 56.8 67.8 189.5 288.3 844.0

R2 (%)
94.48 89.43 86.24 79.44 92.89 93.86

Coefficient of Variation (%)


35.9 14.3 13.3 17.9 11.3 10.6

m value = Weibull modulus; 0.05 = stress levels at 5% probability of failure; 0.01 = stress levels at 1% probability of failure; R 2 = regression coefficient

The Journal of Prosthetic Dentistry

Yilmaz et al

August 2007
tween ICA, ICZ, and CZ core materials. The Weibull statistical analysis of the biaxial flexural strength data are summarized in Table III. Weibull modulus (m) results were 13.26 for CZ, 10.17 for ICZ, 10.13 for E, 7.94 for C, 6.96 for ICA, and 3.17 for F core materials. High Weibull modulus values were obtained for the CZ, ICZ, and E groups. The Weibull modulus values for flexural strength for the C and ICA groups were similar. F had the lowest m value compared with the other test groups. The higher Weibull modulus values of the CZ, ICZ, and E groups indicate that they were uniform materials and more reliable than C, ICA, and F. The Weibull moduli of ICZ (10.17) and E (10.13) were close. For C (7.94) and ICA (6.96), results demonstrated similar Weibull m values and the F ceramic was found to have the lowest m value. The failure strength corresponding to a 1% or 5% probability of failure may be more clinically relevant when designing prostheses than mean strength values.41 Data for CZ, ICZ, and ICA had high 1% and 5% probability of failure values. The lowest 1% and 5% probability of failure values were recorded for F and comparable values for C and E. Highest 1% and 5% probability of failure values were obtained with CZ core material (844 and 951, respectively) and are in agreement with the flexural strength results. Indentation fracture toughness values for the 6 all-ceramic core materials are presented in Figure 4. Material C had the lowest mean toughness (SD) value at 1.73 MPa x m0.5 (0.02), and CZ had the highest toughness value at 6.27 MPa x m0.5 (0.05). The other mean toughness values (SD) in MPa x m0.5 were: 1.88 (0.07) for F, 2.40 (0.09) for E, 4.78 (0.18) for ICA, and 5.56 (0.18) for ICZ. Significant differences (P<.001) were recorded in the indentation fracture toughness values of each core material (Table IV). In the present study, CZ core material had the highest toughness value (6.27 MPa x m0.5) among the studied core materials (P<.001). The fracture toughness value of the other zirconiabased ceramic system ICZ was 5.56 MPa x m0.5. Although significant differences existed among F, C, E, and ICA (1.73-4.78 MPa x m0.5, P<.001), these materials were all significantly lower than zirconia-based ceramic systems (CZ and ICZ, P<.001).

125

Indentation Fracture Toughness (MPa x m1/2)

6 5 4 3 2 1 C F E ICA ICZ CZ

Core Materials 4 Mean values of indentation fracture toughness test results (MPa x m1/2). Lines at top of bars represent standard error.

Table IV. One-way ANOVA statistical results of indentation fracture toughness test
Sum of Squares
Between groups 601.6

df
5

Mean Square
120.3

8279.5 Within groups 2.53 174 .0145

<.001

Yilmaz et al

126
DISCUSSION
Although the relationship between the mechanical properties of a ceramic and its clinical performance is influenced by many variables, some of these properties, namely strength and fracture toughness, have often been the first parameters investigated to determine the clinical potential and limits of a dental ceramic. The standard for testing the strength of dental ceramics has been the 3-point flexural test, but one problem has been the sensitivity of the test to flaws along the specimen edges.39,44 The 3-point bending test is largely dependent on the surface finish of the edges of the specimen.35 Biaxial flexure testing is recognized as a reliable technique and the method of choice (ISO 6872)52 for studying brittle materials since the maximum tensile stress occurs within the central loading area and edge failures are eliminated.44 The method adapted was the one recommended by the International Standard Organization52 because the test standardizes specimen thickness, diameter, shape, and roughness. Wagner and Chu39 studied the biaxial flexural strength of In-Ceram (IC) and Empress (E) with the same method as the current study. The authors found the strength values (SD) of IC and E to be 352 (11) MPa and 134 (22) MPa, respectively. The Weibull moduli were reported as 1.61 and 3.65 for IC and E, respectively. Wen et al,30 using the same method, found the glass infiltrated sintered alumina strength value to be 433 (90) MPa and the leucite-reinforced glass strength value to be 115 (24) MPa. The Weibull moduli (SD) for these materials were 5.37 (0.52) and 5.64 (0.27), respectively. The mean biaxial flexural strength values (SD) for ICA (341.80 (61.13)) and E (101.18 (13.49)) achieved in the current study are in agreement with these results. The Weibull moduli for ICA (6.96) and E (10.13) in the present study were found to be higher than in these studies. This may be because the prepared specimens were reliable and indicate a different flaw size distribution. Kosmac et al42 examined the effect of surface grinding and airborne-particle abrasion on the biaxial flexural strength and reliability of 2 yttria-fully stabilized tetragonal zirconia (Y-TZP) ceramics. The mean flexural strength (SD) for the sintered, fine-grained YTZP ceramic was 1021 (89.5) MPa, and for coarse-grained Y-TZP, the strength value was 914 (58.3) MPa; these values are similar to the CZ values in the present study. According to the results of Weibull statistical analysis of the biaxial flexural strength, the highest m values (SD) were obtained with the sintered fine-grained (10.7 (3.2)) and coarse-grained (14.9 (4.5)) Y-TZP ceramic. Guazzato et al10 investigated the biaxial flexural strength of zirconiabased ceramics and reported that the mean strength values (SD) of ICA and ICZ were 600 (60) MPa and 620 (61) MPa, respectively. The strength values of ICZ reported by Guazzato et al are consistent with the results of the present study. The strength values of ICA that were found in the present study are relatively lower than those found in the Guazzato et al10 study. In the present study, the Weibull modulus value of ICA was lower (6.96 < 10.5), so this may be the reason for differences in the strength results. The Weibull moduli of ICA and ICZ reported by Guazzato et al10 were 10.5 and 10.4, respectively. The Weibull modulus values of ICZ and CZ achieved in the current study were close to those values achieved in the studies of Guazzatto et al10 and Kosmac et al.42 The close Weibull values indicate that the flaws and stress distribution in these ceramics were similar. In the present study, the highest m value was found in CZ and the lowest value was obtained in F. In another study, the mean flexural strength of ICZ was reported as 497 35 MPa. The mean biaxial flexural strength value of ICZ (541.80 61.10) achieved in the current study was relatively higher than the mean

Volume 98 Issue 2
value achieved by Itinoche et al.43 Although the testing methods and the specimen sizes (15-mm diameter, 1.2-mm width) were the same, variables such as specimen preparation may have contributed to the higher flexural strength found in the present study compared to the study by Itinoche et al.43 Restorations placed in the posterior region must withstand occlusal forces ranging from 150 to 665 N.28,29 Core ceramic materials must be veneered with porcelain before clinical use. However, veneer porcelain must be supported by a high strength core material to be used in stress-bearing areas.41 The higher failure loads and the transformation-toughening mechanism of zirconia-reinforced core ceramics are useful for highly loaded allceramic restorations in the posterior dentition. These systems may reduce clinical failure or enlarge the range of clinical applications of the all-ceramic core materials. However, the performance of these materials when layered with porcelain must be evaluated further. Fracture toughness is an important material characterization value. Its value characterizes the resistance of a material against a propagating crack. The lower the fracture toughness is, the lower the clinical reliability of the ceramic restoration, because the KIC value defines the critical stress intensity level at which catastrophic failure occurs due to a (critical) microdefect. Therefore, every new dental ceramic material should be tested, not only with respect to its flexural strength, but also with respect to its fracture toughness, before being introduced commercially.51 Different test methods have been established to evaluate the fracture toughness values. In this study, the indentation technique was used, according to Anstis et al,53 because it is a simple procedure and economical in its use of material. The greatest asset of the technique is the use of a small amount of specimen area. Wagner et al39 and Fischer et al51 used the same

The Journal of Prosthetic Dentistry

Yilmaz et al

August 2007
methodology used in the present study, according to Anstis et al.53 Guazzato et al15 investigated fracture toughness of 9 all-ceramic materials and compared DC Zircon, an experimental yttria-partially stabilized zirconia, ICZ slip, and ICZ dry-pressed ceramics. Fracture toughness values were 7.4 MPa x m0.5 for DC Zircon, 5.5 MPa x m0.5 for Y-ZTP, 4.8 MPa xm0.5 for ICZ, and 4.9 MPa x m0.5 for ICZ dry-pressed ceramics. Both ICZ values were slightly lower than in the present study. This small difference could be related to the differences in the specimen preparation. The fracture toughness value of the Y-ZTP ceramic that was used in this study (CZ) was between the DC Zircon and Y-ZTP toughness values. In 2002, Guazzato et al10 also studied the mechanical properties of ICA and ICZ and calculated the fracture toughness according to both indentation strength (IS) and IF techniques. According to the IF method, the fracture toughness of ICA and ICZ was 2.7 MPa x m0.5 and 3.0 MPa x m0.5, respectively. The fracture toughness values of ICA and ICZ in the present study are relatively high when compared to the results of Guazzato et al.10 This could be due to the intrinsic difficulty in obtaining an objective crack length measurement and, more importantly, to the residual stress fields associated with crack production.53 Residual stresses are responsible for slow crack growth, which occurred immediately after unloading the indenter.53 Therefore, depending on the interval between the indentation and the crack measurement, the fracture toughness measured could be somewhat less than the critical value. In another study, Guazzato et al37 demonstrated fracture toughness of pressable and alumina glass-infiltrated ceramics and calculated the fracture toughness of Empress as 2.9 MPa x m0.5, ICA dry-pressed as 3.6 MPa x m0.5, and ICA slip as 4.4 MPa x m0.5. The E and ICA fracture toughness values from the present study are in agreement with these results. Wagner et al39 investigated the toughness values of Empress, In-Ceram, and Procera All-Ceram core materials. The toughness values for InCeram ceramic (4.49 MPa x m0.5) and Empress ceramic (1.74 MPa x m0.5) were within the same range as found in the present study. The improvement of the in-service reliability of a material can be achieved by increasing its fracture toughness. CZ (6.27 0.05), ICZ (5.56 0.18), and ICA (4.78 0.18) showed high toughness values, respectively, among the tested materials. The present study has several limitations, making it difficult to compare results directly with the clinical situations. The limitations include that this was an in vitro study and the influence of fatigue in the oral cavity was not considered. It is important to note that mechanical tests, used in this study, are only a first step toward predicting clinical performance. The method used was the one recommended by International Standard Organization,52 and the thickness of the specimens tested was 1.2 mm, because the test requires standardization of the specimen thickness, diameter, and shape. Future studies that reflect clinical conditions are necessary for better characterization of new dental ceramics. a significantly different indentation fracture toughness value.

127

REFERENCES
1. Anusavice KJ. Reducing the failure potential of ceramic-based restorations. Part 2: Ceramic inlays, crowns, veneers and bridges. Gen Dent 1997;45:30-5. 2. OBrien WJ. Dental materials and their selection. 3rd ed. Chicago: Quintessence; 2002. p. 210-24. 3. Piconi C, Maccauro G. Zirconia as a ceramic biomaterial. Biomaterials 1999;20:125. 4. Anusavice KJ, Shen C, Lee RB. Strengthening of feldspathic porcelain by ion exchange and tempering. J Dent Res 1992;71:11348. 5. Hobo S, Iwata T. Castable apatite ceramics as a new biocompatible restorative material. I. Theoretical considerations. Quintessence Int 1985;16:135-41. 6. McLean JW. The nature of dental ceramics. In: McLean JW, Hubbard JR, Kedge MI, editors. The science and art of dental ceramics. 1st ed. Chicago: Quintessence; 1979. p. 23-112. 7. Dong JK, Luthy H, Wohlwend A, Scharer P. Heat-pressed ceramics: technology and strength. Int J Prosthodont 1992;5:9-16. 8. Probster L. Compressive strength of two modern all-ceramic crowns. Int J Prosthodont 1992; 5:409-14. 9. Andersson M, Oden A. A new all-ceramic crown - a dense-sintered, high-purity alumina coping with porcelain. Acta Odontol Scand 1993;51:59-64. 10.Guazzato M, Albakry M, Swain MV, Ironside J. Mechanical properties of In-Ceram Alumina and In-Ceram Zirconia. Int J Prosthodont 2002;15:339-46. 11.Kelly JR. Dental ceramics: current thinking and trends. Dent Clin North Am 2004;48:513-30. 12.Seghi RR, Sorensen JA. Relative flexural strength of six new ceramic materials. Int J Prosthodont 1995;8:239-46 13.Probster L, Diehl J. Slip-casting alumina ceramics for crowns and bridge restorations. Quintessence Int 1992;23:25-31 14.Guazzato M, Albakry M, Quach L, Swain MV. Influence of surface and heat treatments on the flexural strength of a glass-infiltrated alumina/zirconia-reinforced dental ceramic. Dent Mater 2005;21:454-63. 15.Guazzato M, Albakry M, Ringer SP, Swain MV. Strength, fracture toughness and microstructure of a selection of all-ceramic materials. Part II: Zirconia-based dental ceramics. Dent Mater 2004;20:449-56. 16.Rizkalla AS, Jones DW. Mechanical properties of commercial high strength ceramic core materials. Dent Mater 2004;20:20712. 17.Chevalier J, Olagnon C, Fantozzi G. Subcritical crack propagation in 3Y-TZP ceramics: static and cyclic fatigue. J Am Ceram Soc 1999;82:3129-38. 18.Srdic VV, Radonjic L. Transformation toughening in sol-gel derived aluminazirconia composites. J Am Ceram Soc 1997;80:2056-60.

CONCLUSIONS
Within the limitations of this in vitro study, the following conclusions were drawn: 1. According to the biaxial flexural strength test, the CZ core material had the highest strength value and the F core material had the lowest strength value among the core materials tested. The biaxial flexural strength values of CZ, ICZ, and ICA core materials were significantly higher than the other core materials. 2. According to the indentation fracture toughness test, CZ core material had the highest and C core material had the lowest fracture toughness value. Each ceramic core material had

Yilmaz et al

128
19.De Aza AH, Chevalier J, Fantozzi G, Schehl M, Torrecillas R. Crack growth resistance of alumina, zirconia and zirconia toughened alumina ceramics for joint prostheses. Biomaterials 2002;23:937-45. 20.Swain MV. Limitation of maximum strength of zirconia-toughened ceramics by transformation toughening increment. J Am Ceram Soc 1985;68:C97-9. 21.Guazzato M, Albakry M, Quach L, Swain MV. Influence of grinding, sandblasting, polishing and heat treatment on the flexural strength of a glass-infiltrated aluminareinforced dental ceramic. Biomaterials 2004;25:2153-60. 22.Filser F, Kocher P, Weibel F, Luthy H, Scharer P, Gauckler LJ. Reliability and strength of all-ceramic dental restorations fabricated by direct ceramic machining (DCM). Int J Comput Dent 2001;4:89-106. 23.Luthy H, Filser F, Loeffel O, Schumacher M, Gauckler LJ, Hammerle CH. Strength and reliability of four-unit all-ceramic posterior bridges. Dent Mater 2005;21:930-7. 24.Suttor D, Bunke K, Hoescheler S, Hauptmann H, Hertlein G. LAVA- the system for all-ceramic ZrO2 crown and bridge frameworks. Int J Comput Dent 2001;4:195-206. 25.De Long R, Douglas WH. An artificial oral environment for testing dental materials. IEEE Trans Biomed Eng 1991;38:339-45. 26.Hidaka O, Iwasaki M, Saito M, Morimoto T. Influence of clenching intensity on bite force balance, occlusal contact area, and average bite pressure. J Dent Res 1999;78:1336-44. 27.Korioth TW, Waldron TW, Versluis A, Schulte JK. Forces and moments generated at the dental incisors during forceful biting in humans. J Biomech 1997;30:631-3. 28.Jemt T, Karlsson S, Hedegard B. Mandibular movement of young adults recorded by intraorally placed light-emitting diodes. J Prosthet Dent 1979;42:669-73. 29.Ferrario VF, Sforza C, Zanotti G, Tartaglia GM. Maximal bite forces in healthy young adults as predicted by surface electromyography. J Dent 2004;32:451-7. 30.Wen MY, Mueller HJ, Chai J, Wozniak WT. Comparative mechanical property characterization of 3 all-ceramic core materials. Int J Prosthodont 1999;12:534-41. 31.Zeng K, Oden A, Rowcliffe D. Evaluation of mechanical properties of dental ceramic core materials in combination with porcelains. Int J Prosthodont 1998;11:183-9. 32.Mecholsky JJ Jr. Fracture mechanics principles. Dent Mater 1995;11:111-2. 33.Kelly JR. Perspectives on strength. Dent Mater 1995;11:103-10. 34.Drummond JL, King TJ, Bapna MS, Koperski RD. Mechanical property evaluation of pressable restorative ceramics. Dent Mater 2000;16:226-33. 35.Zeng K, Oden A, Rowcliffe D. Flexure tests on dental ceramics. Int J Prosthodont 1996; 9:434-9. 36.Apholt W, Bindl A, Luthy H, Mormann WH. Flexural strength of Cerec 2 machined and jointed InCeram Alumina and InCeram Zirconia bars. Dent Mater 2001;17:260-7. 37.Guazzato M, Albakry M, Ringer SP, Swain MV. Strength, fracture toughness and microstructure of a selection of all-ceramic materials. Part I: Pressable and alumina glass-infiltrated ceramics. Dent Mater 2004;20:441-8. 38.Chong KH, Chai J, Takahashi Y, Wozniak W. Flexural strength of In-Ceram Alumina and In-Ceram Zirconia core materials. Int J Prosthodont 2002;15:183-8. 39.Wagner WC, Chu TM. Biaxial flexural strength and indentation fracture toughness of three new dental core ceramics. J Prosthet Dent 1996;76:140-4. 40.Albakry M, Guazzato M, Swain MV. Biaxial flexural strength, elastic moduli, and x-ray diffraction characterization of three pressable all-ceramic materials. J Prosthet Dent 2003; 89:374-80. 41.Cattell MJ, Clarke RL, Lynch EJ. The biaxial flexural strength and reliability of four dental ceramics- Part II. J Dent 1997;25:40914. 42.Kosmac T, Oblak C, Jevnikar P, Funduk N, Marion L. The effect of surface grinding and sandblasting on flexural strength and reliability of Y-TZP zirconia ceramic. Dent Mater 1999; 15:426-33. 43.Itinoche KM, Ozcan M, Bottino MA, Oyafuso D. Effect of mechanical cycling on the flexural strength of densely sintered ceram-

Volume 98 Issue 2
ics. Dent Mater 2006;22:1029-34. 44.Ban S, Anusavice KJ. Influence of test method on the failure stress of brittle dental materials. J Dent Res 1990;69:1791-9. 45.Chandrasekhar BK. Estimation of Weibull parameter with a modified weight function. J Mater Res 1997;12:2638-42. 46.Gee MG, Morrell R. Fracture mechanics and microstructure. In: Bradt RC, Evans AG, Hasselman DP, Lange FF, editors. Fracture mechanics of ceramics: microstructure, methods design, and fatigue. Norwell: Kluwer Academic Publishers; 1986. p. 17-8. 47.Quinn GD. Hardness testing of ceramics. Adv Mater Process 1998;154: 23-8. 48.Scherrer SS, Denry IL, Wiskott HW. Comparison of three fracture toughness testing techniques using a dental glass and a dental ceramic. Dent Mater 1998;14:246-55. 49.Uctasli S, Wilson HJ. Influence of layer and stain firing on the fracture strength of heat-pressed ceramics. J Oral Rehabil 1996;23:170-4. 50.Seghi RR, Denry IL, Rosenstiel SF. Relative fracture toughness and hardness of new dental ceramics. J Prosthet Dent 1995;74:145-50. 51.Fischer H, Marx R. Fracture toughness of dental ceramics: comparison of bending and indentation method. Dent Mater 2002;18:12-9. 52.International Organization for Standardization. ISO 6872:1995, Dental ceramic. Geneva: ISO; 1995. 53.Anstis GR, Chantikul P, Lawn BR, Marshall DB. A critical evaluation of indentation techniques for measuring fracture toughness: I, Direct crack measurements. J Am Ceram Soc 1981;64:533-8. Corresponding author: Dr Handan Yilmaz 8. Cadde, 82. Sokak 06510, Emek Ankara TURKEY Fax: 00903122239226 E-mail: hozkula@gmail.com Copyright 2007 by the Editorial Council of The Journal of Prosthetic Dentistry.

New Product News


The January and July Issues of the Journal carry information regarding new products of interest to prosthodontists. Product information should be sent 1 month prior to ad closing date to: Dr Carol A. Lefebvre, Editor, The Journal of Prosthetic Dentistry, School of Dentistry, AD-1112, Medical College of Georgia, Augusta, GA 30912-1255. Product information may be accepted in whole or in part at the discretion of the Editor and is subject to editing. A black-andwhite glossy photo may be submitted to accompany product information. Information and products reported are based on information provided by the manufacturer. No endorsement is intended or implied by the Editorial Council for The Journal of Prosthetic Dentistry, the Editor, or the publisher.

The Journal of Prosthetic Dentistry

Yilmaz et al

You might also like