You are on page 1of 12

journal of the mechanical behavior of biomedical materials 59 (2016) 279–290

Available online at www.sciencedirect.com

www.elsevier.com/locate/jmbbm

Research Paper

Fracture behavior of inlay and onlay fixed partial


dentures – An in-vitro experimental and XFEM
modeling study

Zhongpu Zhanga, Mark Thompsonb, Clarice Fielda, Wei Lia, Qing Lia,
Michael V. Swainb,n
a
School of Aerospace, Mechanical and Mechatronic Engineering, University of Sydney, NSW 2006, Australia
b
Discipline of Biomaterials, Faculty of Dentistry, University of Sydney, NSW 2010, Australia

art i cle i nfo ab st rac t

Article history: Objectives: This study aimed to explore the “sensitivity” of the fracture load and initiation
Received 11 September 2015 site to loading position on the central occlusal surface of a pontic tooth for both all-ceramic
Received in revised form inlay retained and onlay supported partial denture systems.
10 January 2016 Materials and methods: Three dimensional (3D) finite element (FE) inlay retained and onlay
Accepted 12 January 2016 supported partial denture models were established for simulating crack initiation and
Available online 6 February 2016 propagation by using the eXtended Finite Element Method (XFEM). The models were
Keywords: subjected to a mastication force up to 500 N on the central fossa of the pontic. The loading
All-ceramic position was varied to investigate its influence on fracture load and crack path.
Inlay retained Results: Small perturbation of the loading position caused the fracture load and crack
Onlay supported pattern to vary considerably. For the inlay fixed partial dentures (FPDs), the fracture origins
XFEM changed from the bucco-gingival aspect of the molar embrasure to the premolar
Crack embrasure when the indenter force location is slightly shifted from the mesial to distal
Fracture load side. In contrast, for onlay FPDs, cracking initiated from bucco-gingival aspect of the
premolar embrasure when the indenter is slightly shifted to the buccal side and from
molar embrasure when the indenter is shifted to the lingual side.
Conclusions: The fracture load and cracking path were found to be very sensitive to loading
position in the all-ceramic inlay and onlay FPDs. The study provides a basis for improved
understanding on the role of localized contact loading of the cusp surface in all-
ceramic FPDs.
& 2016 Elsevier Ltd. All rights reserved.

n
Correspondence to: Disciplines of Biomaterials, Faculty of Dentistry, The University of Sydney, Sydney Dental Hospital, 2 Chalmers
ST, Surry Hills, Sydney, NSW 2010 Australia. Tel.: þ61 2 9351 8375.
E-mail addresses: leo.zhang@sydney.edu.au (Z. Zhang), thompsons@iinet.net.au (M. Thompson),
cfie3540@uni.sydney.edu.au (C. Field), wei.li@sydney.edu.au (W. Li), qing.li@sydney.edu.au (Q. Li),
mswain@mail.usyd.edu.au (M.V. Swain).

http://dx.doi.org/10.1016/j.jmbbm.2016.01.035
1751-6161/& 2016 Elsevier Ltd. All rights reserved.
280 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

height and width in order to provide the highest possible


1. Introduction structural strength for enhancing restorative longevity. In
this regard, Oh et al. (2002) investigated the effect of bridge
In contrast to traditional metallic materials commonly used design on fracture resistance by testing a range of connector
in prosthetic dentistry, ceramics are considered superior in radii experimentally. More recently, Sailer et al. (2007)
translucency and tooth like appearance, which are particu- reported an in-vivo study to determine the success rate of
larly attractive to the patients who have a high esthetic three- to five-unit zirconia posterior FPDs over a 5 year
requirement (Denry and Kelly, 2008; Zarone et al., 2011). For clinical follow-up. An in-vitro study by Studart et al. (2007b)
this reason, ceramics have emerged as one of the major revealed that the minimum diameter of the connector should
dental biomaterials in the state-of-the-art prosthodontic be at least 4.2 mm, 5.6 mm and 6.4 mm due to clenching
clinic over the last two decades. Nevertheless, a major generating an intermittent critical loading (800 N) for the
concern for these sophisticated all-ceramic restorations is
three-, four- and five-unit FPDs, respectively, on the basis of
their limited loading capability due to the relatively low
evaluating the fatigue behaviors of prostheses. In their
fracture toughness (Guazzato et al., 2004) and time-
following study they further assessed the fatigue failure of
dependent strength decrease caused by progressive crack
zirconia components subjected to cyclic stresses under water
growth (van Dijken, 1999). Generally, catastrophic fracture
(Studart et al., 2007a). It was recommended that the mini-
of all-ceramic restorations and chipping of veneering materi-
mum area of the connector should be at least 5.7, 12.6 and
als are typical failure modes in the daily clinical practice,
18.8 mm2 for the three-, four- and five-unit FPDs, respectively.
which may necessitate repair or replacement (Swain, 2009;
Another in-vitro study by Kohorst et al. (2008) investigated
Wolfart et al., 2009). This has become a major barrier limiting
the influence of artificial aging on the load-bearing capacity
more extensive exploitation of ceramic materials to fully
of four-unit bridges. They found that the mean fracture load
replace metals in major dental restorative procedures such
of undamaged and non-aged zirconia FPDs was 1525 N,
as multi-unit fixed partial dentures (FPDs), where tensile
significantly higher than the artificially aged groups with
stresses may be substantial.
the means of 904 N, 924 N and 952 N. From a mechanical
The full zirconium dioxide dental restorations, now more
perspective, it would be preferable to increase the size of the
commonly generated by computer-aided design/computer-
connector as much as possible, but it may cause clinical
aided manufacturing (CAD/CAM) technique, have been
problems to the periodontium as it is difficult to implement
widely used in the posterior region, where the occlusal forces
adequate oral hygiene procedures with overly large connec-
are higher (Rafferty et al., 2010; Rekow et al., 2009). Although
tors of FPDs.
zirconia is somewhat a whitish opaque core material and
In addition to extensive experimental studies as to the
may not be entirely suitable for high esthetic requirements
fracture mechanisms and strength of all-ceramic prostheses,
(Baldissara et al., 2010), a number of in-vitro studies have
computational modeling (in-silico) approaches have been
evaluated its clinical appearance and strength. For exam-
drawing increasing attention for their proven predictive
ple, Perry et al. (2012) reported a 2-year in-vivo study on
performance of three- or four-unit (Lava) zirconia bridges. ability. To avoid having to manually prescribe initial cracks
They found that the outcomes were clinically acceptable in in the ceramic materials and structures, two numerical
terms of color stability and matching, marginal discoloration, techniques, namely continuum-to-discrete element method
incidence of caries, changes in restoration–tooth interface (CDEM) and eXtended Finite Element Method (XFEM), have
and surface texture, postoperative sensitivity, marginal integ- proven highly effective in predicting the location of initiation
rity, and maintenance of periodontal health. Beuer et al. and subsequent direction of cracking. Using CDEM, Ichim
(2012) evaluated single crowns with full-contour zirconia in et al. (2007b, 2007c) developed a 2D model for replicating non-
terms of esthetics, wear and load-bearing capacity; and they carious cervical lesions, which enabled an automatic transi-
found that milling zirconia to full-contour with a thinly tion from a continuum solid to discrete cracks at a highly-
glazed surface can be an alternative to traditional porcelain stressed region, thereby simulating dental abfraction lesion
bonded to zirconia based restorations. Another in-vitro study process effectively. Their further study (Ichim et al., 2007a; Li
by Tinschert et al. (2001) determined the fracture resistance et al., 2006) modeled crack initiation and propagation for a
of three-unit FPDs with a maxillary premolar and molar range of dental settings, such as failure of dentine, enamel, a
abutments, indicating that the FPDs cores made of sintered restored tooth, a three-unit all-ceramic FPD structure and
zirconia (DC-Zirkon) exhibited the highest fracture resistance contact-induced damage in the layered material systems. In
and were suitable for making highly loaded all-ceramic both these studies, correlation with experimental data was
prostheses. presented.
Note that all-ceramic FPDs exhibit different mechanical As a relatively new numerical technique, the eXtended
status and complex damage/fracture modes, which chal- Finite Element Method (XFEM) has also been employed to
lenges the design principles established with traditional model crack initiation and growth in traditional engineering
metal based FPDs. In the context of structural analysis, a fields (Bordas et al., 2007; Meschke and Dumstorff, 2007; Xu
number of in-vivo and in-vitro studies have been reported, et al., 2010). Recently, it has been successfully extended to
where restorative failure is a consequence of crack initiation fracture modeling of dental prostheses (Barani et al., 2012,
that begins at the gingival connector embrasure. Hence the 2011; Zhang et al., 2013, 2016). Our comparative study of an
connector shape of inlay retained and onlay supported partial all-ceramic dental bridge showed that XFEM did not only
dentures is a key concern; and its size should be adjusted in correlate to CDEM fracture well, but also enabled simulation
journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290 281

of fracture without remeshing, thereby better facilitating 3D onlay bridges were made from Filtek P60 composite resin for
crack modeling in sophisticated FPD structures. its similarity of Young's modulus to human dentine (Kikuti
Maximum bite force is a typical indicator to the functional et al., 2012). The Filtek P60 composite resin was not soaked in
state of the masticatory system and key mechanical factor of water for a week prior to cementation. In the experiment,
teeth. It can be measured unilaterally or bilaterally in either there was no significant swelling or shrinkage developed in
anterior or posterior teeth. The maximum forces developed in composite resin within the time frame. The adjacent teeth
the molar teeth are higher than those in incisors (Tortopidis were fabricated by using a sectioned silicone impression of a
et al., 1998). Gender and age differences in bite force have also typical human mandibular second molar and second premo-
been reported in literature. It was found that the mean bite lar. The support base was made from self-curing acrylic
force values were significantly higher in males than in (PMMA) with the alveolus of the abutment teeth (Kuehn
females. Varga et al. (2010) reported the mean maximum et al., 2005). The natural resilience of the periodontal liga-
biting force in subjects aged 18 years (males 777.7778.7 N and ment (PDL) of adjacent model teeth was constructed using a
females 481.67190.42 N) and aged 15 years old (males thin layer of poly vinyl siloxane (3M Imprint II). The abut-
522.37181.7 N and females 465.17243.55 N). Braun et al. ments were etched with 37% phosphoric acid for 20 s, in order
(1996) reported the bit force in the same age range (18–20 to produce the highest bond strengths to zirconia FPDs. The
years) was 176 N higher than those found in Varga et al.'s cavity preparation and the bridge components were treated
study, but their finding in an older age group (18–20 years) with an adhesive (3M Single Bond II) to ensure all compo-
was 738 N (Braun et al., 1995), which were close to the results nents of the specimens were fully bonded. A hardened
observed by Varga et al. (778 N) and Sasaki et al. (720 N) stainless steel ball was placed so as to achieve stable contact
(Sasaki et al., 1989). in the central fossa of the pontic, and then loaded at a cross
Despite these abovementioned in-vivo, in-vitro and in- head speed of 0.5 mm/min in a universal testing machine
silico studies, two major issues remain concerning FPD (Instron) until fracture took place.
fracture. First, previous researchers have largely concentrated
on clinical assessment in the performance and longevity of
2.2. Finite element models
various FPDs; and limited attention has been paid on the
fracture mechanics underlying different failure types, e.g.
The computational in-silico study was conducted using
contact-driven Hertzian cracking beneath the indenter
eXtended Finite Element Method (XFEM) for both inlay-
(Kohorst et al., 2008), crack initiated at a point close to the
retained and onlay supported FPDs. The finite element (FE)
gingival side of either mesial or distal connector (Sundh et al.,
models used in this study were constructed based on com-
2005), etc. Although some finite element analyses (FEA) were
puter tomography (CT) images of the right mandibular region.
conducted to help quantify stress concentration and distribu-
The anatomies of cortical bone, cancellous bone, enamel and
tion in an FPD, it does not inherently accommodate the
dentin were captured by using the image processing program
failure criteria from a fracture mechanics perspective. Sec-
ScanIP (Version 6.0, Simpleware Pty. Ltd., UK). The surface
ond, the in-vitro studies showed some diverse fracture
refinements of the dental structures were conducted in
patterns taking place even for comparable FPD structures
ScanIP þNURBS module. After a smoothing operation, all
with nominally similar loading. A major concern is whether a
intersecting vertices within the surface were fixed. Once
loading position has significant influence on the fracture path
surface refinement had been completed, the solid models
and strength in all-ceramic inlay and onlay bridge structures.
were exported as IGES files for further manipulation in
To address the above two issues, this paper adopts XFEM
SolidWorks. The PDL was created by offsetting the root of
to simulate crack initiation and propagation in both all-
the dentin surface, and an assumption of 0.3 mm uniform
ceramic inlay and onlay fixed partially denture (FPD), respec-
thickness was made (Nanci and Bosshardt, 2006; Nanci and
tively. The modeling validation is performed by comparing
Somerman, 2003). The PDL cushions the abutment teeth from
the XFEM simulation results with the in-vitro experiment.
direct contact with the cortical and cancellous bone that form
The hypothesis tested here was to test how loading position
would influence fracture path and failure load, in turn the bone sockets.
providing some insights into the design of FPD structures. The inlay and onlay preparations involved partial/full
cutting of the occlusal surfaces of the abutment teeth. In
the modeling, special considerations were taken for deter-
2. Materials and methods mining the taper angles within the crowns in the cutting
process to match clinical conditions. The geometry of abut-
2.1. Preparation of inlay-retained and onlay supported ment teeth were subtracted for the inlay and onlay prepara-
partial dentures tions from the dental structures. Consequently, the 3D inlay
retained FPDs at both abutment teeth consisted of an inlay
An in-vitro experiment was conducted for all-ceramic inlay- bridge, abutment teeth, periodontal ligament (PDL), cortical
retained and onlay supported FPDs in this study. The STL files bone, cancellous bone, and ball indenter as shown in Fig. 1a.
of the inlay and onlay supported FPD models, created in The 3D onlay model was constructed by the bridge, abutment
SolidWorks (Dassault Systèmes SolidWorks Corp., USA), were teeth, PDL, cortical bone, cancellous bone, and ball indenter
sent to a specialized ceramics laboratory that was able to as shown in Fig. 1b. Such FE models were exported to Abaqus
fabricate bridge replicas in partially sintered Y-TZP through a (Version 6.12, SIMULIA, Providence, RI, USA) for further pre-
CAD/CAM system. The abutment teeth beneath the inlay and processing and XFEM computation.
282 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

Fig. 1 – 3D FE models with dimensions, loading and boundary conditions.

The 3D FE models were meshed using 4-node linear 500 N was applied, which increased linearly during the pre-
tetrahedral elements. Unlike conventional FE studies where scribed loading time of 1 s. The loading position was chosen
total strain energy, stress and/or displacements were typi- as a variable here, and its variation is illustrated in Fig. 2.
cally adopted as convergence indicators (Li et al., 2004b, 2005), In the FEA study, loading position Δd stands for the
different measures are required for XFEM convergence. In this distance from the center of each loading case to the central
fracture study, the initial cracking site was considered appro- fossa of control case, as summarized in Table 1, which will be
priate to ensure that the mesh size had minimal effect on used to investigate sensitivity of loading position to the
fracture analysis. To do so, several different mesh sizes were fracture behaviors. The distance Δd between the central fossa
tested here. It was found that the element sizes of 1 mm for point A ¼(xc, yc, zc) of control case and the center point B ¼(xn,
cortical and cancellous bones, 0.3 mm for PDL, 0.36 mm for yn, zn) of indenter area acting on the occlusal surface is
teeth abutments and bridges were considered adequate. As calculated as:
such, the inlay and onlay models had a total number of qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
300,173 (185,346 degree of freedom (d.o.f.)) and 268,525 ele- Δd ¼ ðxn  xc Þ2 þ ðyn yc Þ2 þ ðzn  zc Þ2 ðn ¼ 1; 2; 3; 4; 5; 6; 7; 8Þ
ments (177,483 d.o.f.), respectively. ð1Þ

2.3. Boundary and loading conditions of the FE models

As shown in Fig. 1. the kinematic boundary conditions for the 2.4. Materials properties
inlay and onlay FE models were the same; namely the
sectioned bony faces were fully fixed (Field et al., 2012). The All the materials involved in this study were considered to be
force acting vertically through a cylindrical indenter of isotropic, homogeneous, and linearly elastic (Ichim et al.,
approximately 5 mm2 was applied to the central fossa of 2007a, 2007c; Zhang et al., 2016). The material properties of
the pontic. The mating surfaces of the inlay/onlay retainers ceramics, bone and tooth tissues are summarized in Table 2.
and abutments are at a distance of 0.005 mm apart for The fracture toughness of FPD ceramic made of partially
contacting. In the analysis, a ramped load to a maximum of sintered zirconia was determined from the single-edge V
journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290 283

Fig. 2 – Variation in loading position on the central fossa of pontic. The center coordinate of contact area for all loading cases
are given on the right.

Table 1 – Distance Δd between the centers of each load


case and that of the control case. Power law criterion as summarized in the following
equation (Giner et al., 2009),
To load control case Δd (mm)  α    
GI GII α GIII α
þ þ ¼1 ð3Þ
Load case 1 1.020 GIC GIIC GIIIC
Load case 2 0.680
Load case 3 1.101 where GI, GII, GIII represents the energy release rates of crack
Load case 4 0.818 extension modes I, II and III, respectively. GIC, GIIC, GIIIC are the
Load case 5 0.858 predefined critical energy release rates at modes I, II and III,
Load case 6 1.469 respectively. In this study, the power α¼ 1 was taken
Load case 7 1.020
(Thompson et al., 2013b; Zhang et al., 2013).
Load case 8 1.224

3. Results
notch beam test from our previous study (Thompson et al.,
2012). As a result, the fracture toughness of FPD ceramic was
3.1. Experimental validations of the inlay-retained and
0.456 MPa√m with a standard deviation of 0.076 MPa√m. The
onlay FPDs
critical energy release rate indicates a parameter related to
fracture toughness which is defined as the energy required
Fig. 3 compares the numerical with experimental force–displace-
for propagating the crack through a unit area of the material
ment curves for the fracture failure of the pre-sintered ceramic
under mode I fracture (Cherepanov, 1967).
inlay FPDs under controlled loading condition. In the experi-
mental force–displacement curve (in blue), the force suddenly
2.5. Fracture criteria for XFEM
dropped after it reached the ultimate tensile strength of the
ceramic and the sample began cracking. It can be seen that the
The XFEM was applied to simulate crack initiation and
inlay FPD model exhibited almost linear force–displacement
propagation of the inlay and onlay FPDs, respectively. The
behavior. Comparing the XFEM simulation with the experimen-
basis for damage initiation was per the criterion of the
tal test, correlation can be identified, in which both the experi-
maximum principal stress (s1) as indicated in Eq. (2). Cracking
mental Fig. 3a and simulated Fig. 3b crack initiation occurred
occurs when the maximum principal stress exceeds the
  from the gingival connector between the second molar and
tensile strength of the ceramic material smax (47 MPa)
e pontic, and then propagated towards the loading position on the
(Maiti and Smith, 1983), factor f refers to the stress ratio in
pontic. Furthermore, initiation of the secondary crack arose at
an element (e):
the occlusal surface of the pontic and then chipped off the cusp
se
fe ¼   1  ð2Þ of the pontic as shown in Fig. 3a (experiment) and c (simulation).
smax
In the XFEM simulation, the loading for the onset of the
Crack extension depends on the calculated stresses and dominant crack initiation and the secondary crack initiation
should be perpendicular to the direction of the maximum are indicated by arrows on the red loading curve (simulation) in
principal stress. The energy-based damage evolution criter- Fig. 3. Clearly, the XFEM results exhibited reasonable agreement
ion is applied in this study. It can be defined as a function of with the experimental results for both the force–displacement
the crack initiation mix-mode using a tabulated form of the curves and fracture patterns (initiation and path).
284 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

Table 2 – Material properties used for XFEM fracture analyses.

Material Elastic modulus (GPa) Poisson's ratio Tensile strength (MPa) Energy release rate (J/m2)

Abutment teeth 84.1 0.2


Cortical bone 12.2 0.26
Cancellous bone 1.22 0.31
Periodontal ligament 0.07 0.45
FPD partially sintered ceramic 15.6 0.28 47 13.3
Steel Indenter 200 0.3

Fig. 3 – Comparison of force–displacement curves for the Fig. 4 – Displacement vs force curves and fracture patterns
experimental and XFEM results as well as their fracture for the onlay FPD experiment and XFEM simulation:
patterns of the inlay FPD with the control (loading) case: (a) fracture pattern from the experimental test; (b) crack
(a) Cracking pattern from experimental test; (b) Cracking initiation from XFEM simulation; (c) XFEM fracture
initiation predicted from XFEM; (c) Secondary crack initiation propagation in the pontic. (For interpretation of the
occurred in the occlusal region of pontic from XFEM. (For references to color in this figure legend, the reader is
interpretation of the references to color in this figure legend, referred to the web version of this article.)
the reader is referred to the web version of this article.)

Fig. 4 compares the XFEM fracture with the experimental ceramic and to a lesser degree surface finish and grain size
results of the onlay FPD model under the controlled loading (approximately 0.2 mm). In the present XFEM analysis, the
condition. The simulated crack initiation occurred at the crack initiated at the bucco-gingival aspect of the molar
gingival connector between the second molar and pontic as connector or premolar connector where the highest first
seen in Fig. 4b. In the XFEM force–displacement curve (red), principal stresses happened to reach the fracture strength;
the load for the onset of crack initiation (Fig. 4b), and its and then the crack propagated obliquely through the body of
nearly complete extension toward the loading point on the the pontic. From Figs. 3 and 4, it was clearly seen that the
pontic (Fig. 4c) are indicated by the black arrows. Again, good XFEM crack origins and growth agreed well with the experi-
agreement between the XFEM and experimental results was mental fracture patterns.
obtained, indicating that the proposed damage and energy In Table 3, the XFEM predictions of fracture origins were
release model was able to the simulate crack initiation and marked and crack propagation directions were indicated by
propagation of the onlay FPD accurately. the black arrows for both the inlay and onlay models. From
In our previous in-vitro study (Thompson et al., 2013a), the the XFEM analysis it is clearly seen that loading position
fracture surfaces of all-ceramic inlay FPDs were examined via largely affects the resultant fracture origins and paths. This is
optical microscopy and scanning electron microscopy (SEM), consistent with the fractographic measurements, where the
enabling to determine the crack origin and propagation crack initiated at the bucco-gingival aspect of the molar
directions from the observed fractographic features such as embrasure or from premolar embrasure; and extended obli-
the hackle and arrest lines. The primary crack initiation site quely to the loading position of pontic in the onlay model.
at the bucco-gingival aspect of the molar embrasure or
premolar embrasure on the pontic had an occluso-lingual 3.2. Effects of loading position on the fracture path and
trajectory with a secondary crack origin occurring from the fracture loads
indenter contact. The critical flaw size of 3379 mm was
overserved from fractographic analysis (Thompson et al., Table 4 compared the fracture patterns of the inlay retained
2013a) which was largely related to the defect size of the FPDs for different loading positions. When the loading
journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290 285

Table 3 – Comparison of different XFEM fracture paths for the inlay and onlay FPDs models.

Model Fracture pattern Cracking origin Crack propagation direction

Pattern 1
(Control loading case)

Inlay
FPD

Pattern 2
(Loading case 6)

Pattern 1
(Control loading case)

Onlay
FPD

Pattern 2
(Loading case 5)

position was slightly shifted from the control (central) to the load-bearing capacities (209 N, 212 N, 216 N) than those at
distal side (loading cases 6 and 7), the crack initiated from the the mesial side (149 N, 160 N, 156 N). This indicates that the
bucco-gingival aspect of the premolar embrasure. It should fracture resistance of the inlay FPDs is higher when the
also be noted that an unstable fracture pattern (loading case loading position was shifted from the mesial to the
8) occurred when the loading position shifted to the linguo- distal side.
distal side, because the retainers are one of the weakest Fig. 5b illustrates the fracture loads at crack initiation for
aspects of inlay FPDs (Thompson et al., 2013a). The cracking the onlay FPD models at the different loading positions. In
patterns were fairly similar in cases 1–5 and control case, load case 7, the contact center of indenter was moved
indicating the similar location of critical stresses. 0.888 mm towards the distal side from the central fossa of
Table 5 shows the fracture patterns of the onlay FPDs pontic, which led to the highest fracture resistance (353 N) of
all loading cases. The failure mode under load case 7 was
under different loading positions. The fracture originates at
similar to load cases 1, 2, 4, 6, and control case, but differs
the bucco-gingival aspect of the molar embrasure for most of
from that under load cases 3, 5, 8, where the loading positions
the loading cases. When the loading position was shifted
were close to the lingual side. As a result, the crack origins
towards the lingual side, cracking initiated from the gingival
and propagation directions were different due to the slight
connector between the pontic and premolar abutment in
shifting of loading positions between these two groups. It
loading cases 3, 5 and 8. It should be noted that such a
should be pointed out that the onlay model is able to bear
pattern differs from that for the inlay FPDs. In the inlay
substantially higher fracture loading than the inlay model as
models, the fracture origin occurred at the premolar embra-
shown by comparing Fig. 5a and b.
sure when loading position shifted to the distal side.
When cracking occurred, the fracture loads (in N) for the
inlay models are graphically plotted in Fig. 5a for the different 4. Discussion
loading positions. Clearly, the load-bearing capacity varies
with loading positions. The models with loading position On the basis of this study, XFEM has demonstrated to be a
close to the distal side resulted in substantially higher powerful simulation tool for predicting both crack initiation
286 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

Table 4 – XFEM fracture patterns for different loading positions in the inlay FPDs model.

Case 1 Case 2 Case 3

Control
Case 4 Case 5
Case

Case 6 Case 7 Case 8

and propagation in the inlay and onlay FPDs with different quantifying the mechanical behaviors as well as the loading
occlusal loading positions. The previous studies investigated influence on fracture responses of the inlay and onlay FPDs.
the fracture origins and crack propagation in FPDs using From an overall fracture perspective, the XFEM analysis
different in-vitro tests (Thompson et al., 2012; Thompson provided a more complete basis for prediction of the failure
et al., 2013a). These diverse fracture patterns observed in the mode. The previous studies revealed that the gingival con-
in-vitro experiments often made it difficult to clarify and nector was a strong stress concentration site, thereby deter-
rationalize the fracture mechanism from the fractographic mining the fracture resistance of the FPD structures (Oh and
features. XFEM, on the other hand, facilitates a more succinct Anusavice, 2002). In this study, the fracture origins and crack
propagation directions from the in-vitro tests and XFEM
investigation into the mechanical causes for the apparent
simulations were consistent with the previous clinical and
inconsistency in fracture patterns. This in-silico study allows
other numerical studies (Mehl et al., 2010; Sundh et al., 2005;
us to explore the effects of loading position on fracture
Taskonak et al., 2008; Wolfart et al., 2007). In this regard, Oh
strength and cracking patterns for different all-ceramic FPDs.
et al. (2002) carried out an FE and fractographic study, and
The in-vitro experiments and the XFEM modeling showed
they found that fracture initiated at the gingival embrasure of
good correlation for both the force–displacement curves and
the connector, and then the crack propagated toward the
fracture patterns as shown in Figs. 3 and 4. The maximum
loading point. Li et al. (2006) showed the same results using
difference in the fracture load was less than 12% (187 N vs
so-called continuum-to-discrete element method (CDEM) and
211 N). This incongruity might be due to the discrepancy
found that the fracture of all-ceramic onlay FPDs originated
between the experimental specimens and XFEM models used from the connectors. In their study, the fracture path simply
in this study, which particularly included geometric error, connected the weakest connector area with the occlusal
difference in material properties, material defects, as well as loading point. Ichim et al.’s further CDEM study (Ichim
contact loading and boundary conditions specified. It must be et al., 2007a) showed a similar fracture path from the
noted that only a small number of specimens were used in connector to the loading point. However, the fracture from
the in-vitro test for obtaining the fracture loads in our in- the CDEM prediction did not consistently show the secondary
house experiments, which may not be sufficient statistically. vertical fracture as observed in the experimental results
XFEM, on the other hand, provided a feasible approach to (Thompson et al., 2012). This phenomenon could only be
journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290 287

Table 5 – XFEM fracture patterns for different loading positions in the onlay FPDs model.

Case 1 Case 2 Case 3

Control
Case 4 Case 5
Case

Case 6 Case 7 Case 8

explained by catastrophic fracture occurring in the pontic load of the inlay FPD observed here is much lower than those
before secondary fracture finally appeared. previously reported in the literature.
The determination of the fracture load is always of con- The XFEM studies presented here show a strong effect of
cern for prosthodontists, dental technicians and patients. In loading position on the fracture path and fracture load. The
this study, the maximum load 500 N was applied in a ramp influence of loading position on the resultant stress concen-
fashion in the XFEM fracture analysis. During the ramped tration was considered in the previous cantilever FPD struc-
loading process, the fracture load was quantified when the tures (Li et al., 2004a, 2004b). In Oh et al.'s experimental study
first cracking event took place. Such a loading is similar to on the onlay FPDs (Oh and Anusavice, 2002), the fracture path
other studies reported in the literature. For some in-vitro was oblique gingivo-occlusally through the molar connector
and pontic. Ichim et al.'s CDEM study (Ichim et al., 2007a)
studies, Luthy et al. (2005) tested four-unit bridges made of
showed the same path, whereas Li et al.'s CDEM simulation
IPS Empress 2 with a connector cross-sectional area of
(Li et al., 2006) predicted a different path, specifically from the
7.3 mm2 under a mastication force up to 500 N. Rezaei et al.
premolar connector to the loading point. From these three
(2011) used an even higher maximum bite force of 600 N to
studies, cracks developed either in the mesial area or in the
examine the effect of increased connector width on the stress
distal area of the pontic. The different locations for the
distribution in the posterior FPDs.
origins of cracking may be explained as the consequence of
The fracture load at crack initiation for the inlay FPD
occlusal contact variations, geometry of connectors and
(174 N for control loading case) is almost 50% lower than that
length of the moment arm relative to the contact areas. Such
of an onlay FPD (341 N). For the zirconia inlay retained FPD significant discrepancy necessitates further understanding of
the mean fracture load was reported to be 12487263 N when the associated loading induced fracture mechanics.
the interabutment distance was 10 mm (Kilicarslan et al., This study demonstrated that XFEM was suitable for
2004). Since the zirconia ceramic used in our study was not simulating crack initiation and propagation in the all-
fully sintered in order to ease of milling, accurate replication, ceramic fixed partial dentures with inlay and onlay config-
fully isotropic and homogeneous material behavior, the urations. Nevertheless, a number of limitations still need to
Young’s modulus and ultimate tensile strength of FPD cera- be addressed. First, some fracture paths did not completely
mic is much lower. This is also the reason why the fracture extend to the occlusal surface of the pontic tooth. This can be
288 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

Fig. 5 – Fracture loads at crack initiation under different loading positions: (a) Inlay bridge models; (b) Onlay bridge models.

explained as the complexity of 3D modeling adopted here stresses in the zirconia fixed partial dentures. It however
and the stipulated analysis convergence conditions became neglected the cyclic nature of chewing functions in oral
very difficult to satisfy in XFEM upon cracking to a certain environment. Sixth, it is assumed that the components were
extent. ABAQUS provides a viscous regularization capability fully bonded at the interface, so bonding strength of the
that could help improve the convergence when material cement layer between retainer and abutment was not speci-
models exhibiting stiffness degradation and softening beha- fically considered in the XFEM analysis. In future study,
vior (Liu and Zheng, 2010). Another approach to help with debonding of the FDPs due to early failure of the adhesive
such a convergence problem is to use automatic stabilization interface between retainer and cement layer needs to be
(Loehnert, 2014). Second, the assumption of PDL was made as addressed. Likely fracture would be caused by such unfavor-
linear elastic similarly to numerous studies in the literature able stress location in combination with direct exposure to
(Field et al., 2009; Sarrafpour et al., 2013; Zhang et al., 2016). As the oral environment. Seventh, the partially-sintered ceramic
a soft tissue, the PDL is non-linear material which may better was used in the in-vitro experiments for reducing fabrication-
buffer mechanical loading (Chen et al., 2014; Qian et al., 2009; induced discrepancy in material and structural geometry
Ziegler et al., 2005). Its role on fracture analysis of FPD should from in-silico models. While it may not affect the relative
be further studied. Third, the abutment tooth structures were sensitivity analysis of loading position, a study on fully-
sintered ceramic structures would be of more direct clinical
not modeled in detail. The enamel, dentine and pulp struc-
relevance. It should be pointed out that overcoming these
tures remaining in the abutments may affect the fracture
limitations could enhance the reliability of the numerical
resistance of the pontic (Li et al., 2006). Although the volume
simulation. From mechanical point of view, this study has
of the pulp is only a very small part of the entire abutment
established a feasible numerical approach to explore the
tooth, the lack of this feature in the in-vitro sample and the
fracture behavior of all-ceramic FPDs, potentially providing
FE model may not replicate the real clinical situation. Fourth,
a new tool for the design and fabrication of more reliable
like many other modeling analyses, material heterogeneity
prostheses.
and presence of defects in the ceramic samples were not
modeled in this XFEM study, which may to a certain extent
influence the results. Note that XFEM allows pre-inserting 5. Conclusions
micro-cracks and defects efficiently, which will enable us to
take into account these inhomogeneities in future modeling XFEM with its ability to simulate cracking initiation and
studies upon availability of the microscopic details of the propagation was demonstrated to be effective for fracture
specific ceramic materials investigated (Grogan et al., 2015). analysis of all-ceramic dental restorations. The XFEM crack
Fifth, the load-bearing capacities of the specimens were patterns and fracture loads were validated through the good
tested using monotonic load to fracture in the in-vitro study, agreement with the in-vitro experiments. The primary goal of
which simulates the bite force and predominantly causes the this paper was to explore the sensitivity of the fracture
journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290 289

behaviors to the loading position in the three-unit all-ceramic Ichim, I., Li, Q., Li, W., Swain, M.V., Kieser, J., 2007a. Modelling of
inlay and onlay FPDs. Within its limitations, the XFEM fracture behaviour in biomaterials. Biomaterials 28,
fracture analysis clearly indicated the strong association of 1317–1326.
Ichim, I., Li, Q., Loughran, J., Swain, M.V., Kieser, J., 2007b.
fracture patterns and failure loads with the loading position.
Restoration of non-carious cervical lesions – Part I. Modelling
The data generated from this study is considered indicative
of restorative fracture. Dent. Mater. 23, 1553–1561.
for better understanding of failure modes of different ceramic Ichim, I.P., Schmidlin, P.R., Li, Q., Kieser, J.A., Swain, M.V., 2007c.
FPDs, which is of implication for developing mechanically Restoration of non-carious cervical lesions – Part II. Restora-
more reliable dental prostheses. tive material selection to minimise fracture. Dent. Mater. 23,
1562–1569.
Kikuti, W.Y., Chaves, F.O., Di Hipolito, V., Rodrigues, F.P., Perlatti
Acknowledgment D’Alpino, P.H., 2012. Fracture resistance of teeth restored with
different resin-based restorative systems. Braz. Oral Res. 26,
275–281.
The support from Australian Research Council (DP1095140) is
Kilicarslan, M.A., Kedici, P.S., Kucukesmen, H.C., Uludag, B.C.,
grateful. 2004. In vitro fracture resistance of posterior metal-ceramic
and all-ceramic inlay-retained resin-bonded fixed partial
r e f e r e n c e s dentures. J. Prosthet. Dent. 92, 365–370.
Kohorst, P., Dittmer, M.P., Borchers, L., Stiesch-Scholz, M., 2008.
Influence of cyclic fatigue in water on the load-bearing
Baldissara, P., Llukacej, A., Ciocca, L., Valandro, F.L., Scotti, R., capacity of dental bridges made of zirconia. Acta Biomater. 4,
2010. Translucency of zirconia copings made with different 1440–1447.
CAD/CAM systems. J. Prosthet. Dent. 104, 6–12. Kuehn, K.D., Ege, W., Gopp, U., 2005. Acrylic bone cements:
Barani, A., Bush, M.B., Lawn, B.R., 2012. Effect of property composition and properties. Orthop. Clin. N. Am. 36, 17–28.
gradients on enamel fracture in human molar teeth. J. Mech. Li, Q., Ichim, I., Loughran, J., Li, W., Swain, M., Kieser, J., 2006.
Behav. Biomed. Mater. 15, 121–130. Numerical simulation of crack formation in all ceramic dental
Barani, A., Keown, A.J., Bush, M.B., Lee, J.J.W., Chai, H., Lawn, B.R., bridge. Key Eng. Mater. 312, 293–298.
2011. Mechanics of longitudinal cracks in tooth enamel. Acta Li, W., Swain, M.V., Li, Q., Ironside, J., Steven, G.P., 2004a. Fibre
Biomater. 7, 2285–2292. reinforced composite dental bridge. Part I: experimental
Beuer, F., Stimmelmayr, M., Gueth, J.F., Edelhoff, D., Naumann, M., investigation. Biomaterials 25, 4987–4993.
2012. In vitro performance of full-contour zirconia single Li, W., Swain, M.V., Li, Q., Ironside, J., Steven, G.P., 2004b. Fibre
crowns. Dent. Mater. 28, 449–456. reinforced composite dental bridge. Part II: numerical inves-
Bordas, S., Conley, J.G., Moran, B., Gray, J., Nichols, E., 2007. A tigation. Biomaterials 25, 4995–5001.
simulation-based design paradigm for complex cast compo- Li, W., Swain, M.V., Li, Q., Steven, G.P., 2005. Towards automated
nents. Eng. Comput. 23, 25–37. 3D finite element modeling of direct fiber reinforced compo-
Braun, S., Bantleon, H.P., Hnat, W.P., Freudenthaler, J.W., Marcotte, site dental bridge. J. Biomed. Mater. Res. Part B – Appl.
M.R., Johnson, B.E., 1995. A study of bite force, part 1: Biomater. 74B, 520–528.
relationship to various physical characteristics. Angle Orthod. Liu, P.F., Zheng, J.Y., 2010. Recent developments on damage
65, 367–372. modeling and finite element analysis for composite lami-
Braun, S., Hnat, W.P., Freudenthaler, J.W., Marcotte, M.R., Hönigle, nates: a review. Mater. Des. 31, 3825–3834.
K., Johnson, B.E., 1996. A study of maximum bite force during Loehnert, S., 2014. A stabilization technique for the regularization
growth and development. Angle Orthod. 66, 261–264. of nearly singular extended finite elements. Comput. Mech.
Chen, J., Li, W., Swain, M.V., Ali Darendeliler, M., Li, Q., 2014. A 54, 523–533.
periodontal ligament driven remodeling algorithm for ortho- Luthy, H., Filser, F., Loeffel, O., Schumacher, M., Gauckler, L.J.,
dontic tooth movement. J. Biomech. 47, 1689–1695.
Hammerle, C.H.F., 2005. Strength and reliability of four-unit
Cherepanov, G.P., 1967. Crack propagation in continuous media:
all-ceramic posterior bridges. Dent. Mater. 21, 930–937.
PMM vol. 31, no. 3, 1967, pp. 476–488. J. Appl. Math. Mech. 31,
Maiti, S.K., Smith, R.A., 1983. Comparison of the criteria for
503–512.
mixed-mode brittle-fracture based on the preinstability
Denry, I., Kelly, J.R., 2008. State of the art of zirconia for dental
stress-strain field. Part 1. slit and elliptical cracks under
applications. Dent. Mater. 24, 299–307.
uniaxial tensile loading. Int. J. Fract. 23, 281–295.
Field, C., Ichim, I., Swain, M.V., Chan, E., Darendeliler, M.A., Li, W.,
Mehl, C., Ludwig, K., Steiner, M., Kern, M., 2010. Fracture strength
Lig, Q., 2009. Mechanical responses to orthodontic loading: a
of prefabricated all-ceramic posterior inlay-retained fixed
3-dimensional finite element multi-tooth model. Am. J.
Orthod. Dentofac. Orthop. 135, 174–181. dental prostheses. Dent. Mater. 26, 67–75.
Field, C., Li, Q., Li, W., Thompson, M., Swain, M., 2012. A Meschke, G., Dumstorff, P., 2007. Energy-based modeling of
comparative mechanical and bone remodelling study of all- cohesive and cohesionless cracks via X-FEM. Comput. Meth-
ceramic posterior inlay and onlay fixed partial dentures. J. ods Appl. Mech. Eng. 196, 2338–2357.
Dent. 40, 48–56. Nanci, A., Bosshardt, D.D., 2006. Structure of periodontal tissues
Giner, E., Sukumar, N., Tarancón, J.E., Fuenmayor, F.J., 2009. An in health and disease. Periodontology 2000 (40), 11–28.
Abaqus implementation of the extended finite element Nanci, A., Somerman, M., 2003. In: Ten Cate Oral Histology:
method. Eng. Fract. Mech. 76, 347–368. Development, Structure, and Function. Harcourt Health
Grogan, D.M., Bradaigh, C.M.O., Leen, S.B., 2015. A combined Sciences, St. Louis.
XFEM and cohesive zone model for composite laminate Oh, W., Gotzen, N., Anusavice, K.J., 2002. Influence of connector
microcracking and permeability. Compos. Struct. 120, 246–261. design on fracture probability of ceramic fixed-partial den-
Guazzato, M., Albakry, M., Ringer, S.P., Swain, M.V., 2004. Strength, tures. J. Dent. Res. 81, 623–627.
fracture toughness and microstructure of a selection of all- Oh, W.S., Anusavice, K.J., 2002. Effect of connector design on the
ceramic materials. Part II. Zirconia-based dental ceramics. fracture resistance of all-ceramic fixed partial dentures.
Dent. Mater. 20, 449–456. J. Prosthet. Dent. 87, 536–542.
290 journal of the mechanical behavior of biomedical materials 59 (2016) 279 –290

Perry, R.D., Kugel, G., Sharma, S., Ferreira, S., Magnuson, B., 2012. approach for validating the finite element analysis. Aust.
Two-year evaluation indicates zirconia bridges acceptable Dent. J. 57, 23–30.
alternative to PFMs. Compend. Contin. Educ. Dent. (James- Thompson, M.C., Sornsuwan, T., Swain, M.V., 2013a. The all-
burg, N. J.: 1995) 33, e1–e5. ceramic, inlay supported fixed partial denture. Part 4. Fracture
Qian, L., Todo, M., Morita, Y., Matsushita, Y., Koyano, K., 2009. surface analyses of an experimental model, all-ceramic, inlay
Deformation analysis of the periodontium considering the supported fixed partial denture. Aust. Dent. J. 58, 141–147.
viscoelasticity of the periodontal ligament. Dent. Mater. 25, Thompson, M.C., Zhang, Z., Field, C.J., Li, Q., Swain, M.V., 2013b.
1285–1292. The all-ceramic, inlay supported fixed partial denture. Part 5.
Rafferty, B.T., Janal, M.N., Zavanelli, R.A., Silva, N.R.F.A., Rekow, E.D., Extended finite element analysis validation. Aust. Dent. J. 58,
Thompson, V.P., Coelho, P.G., 2010. Design features of a three- 434–441.
dimensional molar crown and related maximum principal Tinschert, J., Natt, G., Mautsch, W., Augthun, M., Spiekermann, H.,
stress. A finite element model study. Dent. Mater. 26, 156–163. 2001. Fracture resistance of lithium disilicate-, alumina-, and
Rekow, E.D., Zhang, G.M., Van, T., Kim, J.W., Coehlo, P., Zhang, Y., zirconia-based three-unit fixed partial dentures: a laboratory
2009. Effects of geometry on fracture initiation and propaga- study. Int. J. Prosthod. 14, 231–238.
tion in all-ceramic crowns. J. Biomed. Mater. Res. Part B – Appl. Tortopidis, D., Lyons, M.F., Baxendale, R.H., Gilmour, W.H., 1998.
Biomater. 88B, 436–446. The variability of bite force measurement between sessions,
Rezaei, S.M.M., Heidarifar, H., Arezodar, F.F., Azary, A., Mokhtar- in different positions within the dental arch. J. Oral Rehabil.
ykhoee, S., 2011. Influence of connector width on the stress 25, 681–686.
distribution of posterior bridges under loading. J. Dent. 8, van Dijken, J.W., 1999. All-ceramic restorations: classification and
67–74. clinical evaluations. Compend. Contin. Educ. Dent. (James-
Sailer, I., Feher, A., Filser, F., Gauckler, L.J., Luethy, H., Hammerle, burg, N. J.: 1995) 20, 1115–1124 1126 passim; quiz 1136.
C.H.F., 2007. Five-year clinical results of zirconia frameworks Varga, S., Spalj, S., Lapter Varga, M., Anic Milosevic, S., Mestrovic,
for posterior fixed partial dentures. Int. J. Prosthod. 20, S., Slaj, M., 2010. Maximum voluntary molar bite force in
383–388. subjects with normal occlusion. Eur. J. Orthod. 33 (4), 427–433.
Sarrafpour, B., Swain, M., Li, Q., Zoellner, H., 2013. Tooth eruption Wolfart, S., Harder, S., Eschbach, S., Lehmann, F., Kern, M., 2009.
results from bone remodelling driven by bite forces sensed by Four-year clinical results of fixed dental prostheses with
soft tissue dental follicles: a finite element analysis. PLoS One, 8. zirconia substructures (Cercon): end abutments vs. cantilever
Sasaki, K., Hannam, A.G., Wood, W.W., 1989. Relationships design. Eur. J. Oral Sci. 117, 741–749.
between the size, position, and angulation of human jaw Wolfart, S., Ludwig, K., Uphaus, A., Kern, M., 2007. Fracture
muscles and unilateral first molar bite force. J. Dent. Res. 68, strength of all-ceramic posterior inlay-retained fixed partial
499–503. dentures. Dent. Mater. 23, 1513–1520.
Studart, A.R., Filser, F., Kocher, P., Gauckler, L.J., 2007a. Fatigue of Xu, J., Li, Y., Chen, X., Yan, Y., Ge, D., Zhu, M., Liu, B., 2010.
zirconia under cyclic loading in water and its implications for Characteristics of windshield cracking upon low-speed
the design of dental bridges. Dent. Mater. 23, 106–114. impact: numerical simulation based on the extended finite
Studart, A.R., Filser, F., Kocher, P., Gauckler, L.J., 2007b. In vitro element method. Comput. Mater. Sci. 48, 582–588.
lifetime of dental ceramics under cyclic loading in water. Zarone, F., Russo, S., Sorrentino, R., 2011. From porcelain-fused-
Biomaterials 28, 2695–2705. to-metal to zirconia: clinical and experimental considerations.
Sundh, A., Molin, M., Sjogren, G., 2005. Fracture resistance of Dent. Mater. 27, 83–96.
yttrium oxide partially-stabilized zirconia all-ceramic bridges Zhang, Z., Guazzato, M., Sornsuwan, T., Scherrer, S.S., Rungsiya-
after veneering and mechanical fatigue testing. Dent. Mater. kull, C., Li, W., Swain, M.V., Li, Q., 2013. Thermally induced
21, 476–482. fracture for core-veneered dental ceramic structures. Acta
Swain, M.V., 2009. Unstable cracking (chipping) of veneering Biomater. 9, 8394–8402.
porcelain on all-ceramic dental crowns and fixed partial Zhang, Z., Zheng, K., Li, E., Li, W., Li, Q., Swain, M.V., 2016.
dentures. Acta Biomater. 5, 1668–1677. Mechanical benefits of conservative restoration for dental
Taskonak, B., Yan, J., Mecholsky Jr., J.J., Sertgoez, A., Kocak, A., fissure caries. J. Mech. Behav. Biomed. Mater. 53, 11–20.
2008. Fractographic analyses of zirconia-based fixed partial Ziegler, A., Keilig, L., Kawarizadeh, A., Jager, A., Bourauel, C., 2005.
dentures. Dent. Mater. 24, 1077–1082. Numerical simulation of the biomechanical behaviour of
Thompson, M.C., Field, C.J., Swain, M.V., 2012. The all-ceramic, multi-rooted teeth. Eur. J. Orthod. 27, 333–339.
inlay supported fixed partial denture. Part 3. Experimental

You might also like