You are on page 1of 9

d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606

Available online at www.sciencedirect.com

ScienceDirect

journal homepage: www.intl.elsevierhealth.com/journals/dema

Mechanical properties and fracture behavior of


flowable fiber reinforced composite restorations

Lippo Lassila a,b , Filip Keulemans a , Eija Säilynoja c , Pekka K. Vallittu a,b,d ,
Sufyan Garoushi a,b,∗
a Department of Biomaterials Science, Institute of Dentistry, University of Turku, Finland
b Turku Clinical Biomaterial Center—TCBC, Institute of Dentistry, University of Turku, Turku, Finland
c Research Development and Production Department, Stick Tech Ltd—Member of GC Group, Turku, Finland
d City of Turku Welfare Division, Oral Health Care, Turku, Finland

a r t i c l e i n f o a b s t r a c t

Article history: Objective. The aim was to evaluate the effect of short glass-fiber/filler particles proportion
Received 30 August 2017 on fracture toughness (FT) and flexural strength (FS) of an experimental flowable fiber-
Received in revised form reinforced composite (Exp-SFRC) with two methacrylate resin formulations. In addition,
18 November 2017 we wanted to investigate how the fracture-behavior of composite restorations affected by
Accepted 8 January 2018 FT values of SFRC-substructure.
Methods. Exp-SFRC was prepared by mixing 50 wt% of dimethacrylate based resin matrix
(bisGMA or UDMA based) to 50 wt% of various weight fractions of glass-fiber/particulate
Keywords: filler (0:50, 10:40, 20:30, 30:20, 40:10, 50:0 wt%, respectively). FT and FS were determined for
Fiber reinforced flowable composite each experimental material following standards. Specimens (n = 8) were dry stored (37 ◦ C for
Fracture toughness 2 days) before they were tested. Four groups of posterior composite crowns (n = 6) composed
Posterior restoration of different Exp-SFRCs as substructure and surface layer of commercial particulate filler
composite were fabricated. Crowns were statically loaded until fracture. Failure modes were
visually examined. The results were statistically analysed using ANOVA followed by post hoc
Tukey’s test.
Results. ANOVA revealed that ratio of glass-fiber/particulate filler had significant effect
(p < 0.05) on tested mechanical properties of the Exp-SFRC with both monomer systems.
Exp-SFRC (50 wt%) had significantly higher FT (2.6 MPam1/2 ) and FS (175.5 MPa) (p < 0.05) com-
pared to non-reinforced material (1.3 MPam1/2 , 123 MPa). Failure mode analysis of crown
restorations revealed that FT value of the substructure directly influenced the failure mode.
Significance. This study shows that short glass-fibers can significantly reinforce flowable com-
posite resin and the FT value of SFRC-substructure has prior importance, as it influences
the crack arresting mechanism.
© 2018 The Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.


Corresponding author at: Department of Biomaterials Science, Institute of Dentistry and TCBC, University of Turku, Turku, Finland.
E-mail address: sufgar@utu.fi (S. Garoushi).
https://doi.org/10.1016/j.dental.2018.01.002
0109-5641/© 2018 The Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.
d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606 599

fibers lengths for the SFRC were measured between 0.3 mm


1. Introduction and 1.9 mm, which makes the lowest aspect ratio 18 and great-
est 112 [18].
Today, particulate filler composite resins (PFCs) are selected
Many of the properties of fiber composites are strongly
on a regular basis for direct and laboratory made poste-
dependent on microstructural parameters such as resin
rior restorations, as an extension to their original indication,
matrix, fiber diameter, fiber length, fiber orientation and fiber
which was limited to direct restorations in anterior teeth.
loading [18,19]. The intent of this study was to evaluate the
Their use has been widened not only to the posterior intra-
effect of microfibers loading and resin matrix type on the
coronal area, but also to extra-coronal restorations [1]. In
mechanical properties of flowable dental resin composite. To
addition, composite resins came into focus for the fabrica-
the authorś knowledge, very little research exists in this field.
tion of resin-bonded fixed dental prostheses (RBFDP) after the
Thus, the hypothesis evaluated was that a flowable dental
introduction of fiber-reinforced composites (FRC). However,
composite may be significantly reinforced by incorporating
insufficient material properties limited the success of com-
randomly distributed microfibers, i.e. fibers of smaller diame-
posite restorations in high stress bearing areas [2,3]. Fracture
ter and length than those used in the SFRC which is available
within the body (bulk) and margins of restorations and sec-
commercially (everX Posterior). Further it was hypothesized
ondary caries have been cited as major problems regarding
that, fracture behavior of biomimetic composite restorations
the failure of posterior composites [4]. The fracture related
can be affected by fracture toughness values of SFRC substruc-
material properties, such as fracture resistance, deforma-
ture (core material).
tion under occlusal load, and the marginal degradation of
materials have usually been evaluated by the determination
of the basic material parameters of flexural strength and 2. Materials and methods
fracture toughness [5]. Fracture toughness is a mechanical
property that describes the resistance of brittle materials to 2.1. Materials
the catastrophic propagation of flaws under an applied load,
and thus, it describes damage tolerance of the material and Bisphenol-A-glycidyl dimethacrylate (bis-GMA), triethylene
can be seen as a measure of fatigue resistance. Fracture tough- glycol dimethacrylate (TEGDMA), and polymethylmethacry-
ness values are dependent on the physical properties and late (PMMA) were purchased from Esstech Inc. (Essington, PA,
chemical composition of the individual component of restora- USA). Diurethane dimethacrylate (UDMA), camphoroquinone
tive material. From a biomimetic point of view, we strive to (CQ), and N,N -dimethyl aminoethyl methacrylate (DMAEMA)
replace lost tooth tissue by biomaterials with similar physi- were obtained from Sigma-Aldrich Co. (St Louis, MO, USA).
cal properties, especially with reference to flexural strength, Silaned (␥-methacryloxypropyl-tri-methoxy-silane) BaAlSiO2
modulus and thermal expansion coefficient [6]. A generally filler particles were received from Schott (UltraFine GM27884,
accepted biomimetic restorative approach advocates replac- Schott, Landshut, Germany). All of reagents were used without
ing enamel with feldspathic porcelain or glass ceramic and purification.
dentine by hybrid composites [6,7]. Although such approach
seems effective there are still relevant mechanical properties, 2.2. Production of flowable fiber reinforced composite
such as fracture toughness, not taken into account. Fracture
toughness of hybrid PFC is significantly lower than that of Experimental flowable short fiber reinforced composite
dentine [8]. Furthermore, the microstructure of hybrid com- (Exp-SFRC) was prepared by mixing 50 wt% of dimethacry-
posite does not match that of dentine. Hybrid PFC consists of late based resin matrix (bisGMA/TEGDMA = 60/40 or
filler particles embedded in a resin matrix while dentine con- UDMA/PMMA = 80/20 wt/wt respectively, with 0.7 wt% CQ
sists of collagen fibers embedded in a hydroxyapatite matrix. and 0.7 wt% DMAEMA as a photoinitiator system) to 50 wt% of
Therefore dentine should be rather seen as a fiber-reinforced various weight fractions of glass fiber/particulate filler (0:50,
composite instead of a particulate filler composite. For that 10:40, 20:30, 30:20, 40:10, 50:0 wt%, respectively) (Table 1). The
reason improvement could be found when taking advantage of discontinuous E-glass fibers (as-received silanized) having
a more dentine-like and high toughness composite as dentine length scale of 200–300 ␮m (Ø6 ␮m), so-called microfibers
replacement. A material which has high fracture toughness and silaned BaAlSiO2 filler particles (Ø7 ␮m) were added
has the ability to better resist crack initiation and propagation. gradually to the resin matrix. The average aspect ratio of
Consequently, the property of fracture toughness and flexural fibers was 41. The mixing was carried out by using a high
strength become important criterions in a dental materials’ speed mixing machine for 2 min (Hauschild Speed Mixer
longevity [9,10]. DAC 400.1, 3500 rpm). Temperatures during the mixing were
Till now, composite resins reinforced with millimetre scale monitored by an infrared thermometer.
short glass fibres (SFRC) are the most interesting materi-
als because of their close resemblance to dentine at the
2.3. Flexural strength
level of microstructure and mechanical properties [11–14].
This composite was previously reported to exhibit improved
Three-point bending test specimens (2 × 2 × 25 mm3 ) were
mechanical properties regarding strength, fracture toughness,
made from each tested composite. Bar-shaped specimens
fatigue resistance and polymerisation shrinkage and to show a
were made in a half-split stainless steel mold between trans-
more favorable (repairable) type of failure behavior in compar-
parent Mylar sheets. Polymerization of the composite was
ison to particulate filler composite [15–17]. The discontinuous
done using a hand light-curing unit (Elipar S10, 3 M ESPE,
600 d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606

Table 1 – Classification of the experimental flowable SFRC test groups used in the study according to their fiber weight
content.
Experimental groups Micro glass fiber (wt%) Resin matrix (wt%) Particulate fillers (wt%)
A (control) 0 50 50
B 10 50 40
C 20 50 30
D 30 50 20
E 40 50 10
F 50 50 0

St. Paul, MN, USA) for 20 s in five separate overlapping por-


tions from both sides of the metal mold. The wavelength of
the light was between 430 and 480 nm and light intensity
was 1600 mW/cm2 . The specimens from each group (n = 8)
were stored dry at 37 ◦ C temperature for 48 h before testing.
Three-point bending test was conducted according to the ISO
4049 (test span: 20 mm, cross-head speed: 1 mm/min, inden-
ter: 2 mm diameter). All specimens were loaded in material
testing machine (model LRX, Lloyd Instrument Ltd, Fareham, Fig. 1 – Schematic drawing of core-crown restoration and
England) and the load-deflection curves were recorded with the compression load test setup.
PC-computer software (Nexygen 4.0, Lloyd Instruments Ltd,
Fareham, England).
Flexural strength (o’f ) was calculated from the following timeters (cm), W is the specimen width (depth) in cm, x is
formula (ISO 1992): a geometrical function dependent on a/W and a is the crack
length in cm.
o’f = 3Fm I/2bh2
Work of fracture (the energy required to fracture the
where Fm is the applied load (N) at the highest point of load- specimen) was calculated from the area under the load-
deflection curve, I is the span length (20 mm), b is the width of displacement curve of single-edge-notched-beam specimens
test specimens and h is the thickness of test specimens. and reported in unit of Ncm.

2.4. Fracture toughness 2.5. Core and crown fabrication

Single-edge-notched-beam specimens (2.5 × 5 × 25 mm3 ) Full crown preparation of lower 1st molar was prepared on
according to adapted ISO 20795-2 standard method (ASTM Frasaco acrylic tooth (Frasaco GmbH, Tettnang, Germany).
2005) were prepared to determine the fracture toughness. Preparation was made with 2 mm of axial and occlusal reduc-
Custom-made stainless steel split mold was used, which tion. A transparent template matrix of an ideally contoured
enabled specimen’s removal without force. Accurately lower 1st molar crown and the prepared tooth structure (core)
designed slot was fabricated centrally in the mold extending was used to aid standardized core-crown restorations con-
until its mid-height, which enabled central location of the struction. Core-crown restorations were constructed via three
notch and optimization of the crack length (x) to be half of different approaches (6 groups) in order to simulate clinical
specimen’s height. The composite material was inserted into techniques (Fig. 1). Two commercial composite resins were
the mold placed over a Mylar-strip-covered glass slide in used in this study, G-aenial Anterior and everX Posterior
one increment. Before polymerization a sharp and centrally (Table 2).
located crack was produced by inserting a straight edged Approach A: group (control) made only from PFC (G-aenial
steel blade into the prefabricated slot. Polymerization of the Anterior) resin.
composite was carried out for 20 s in five separate overlapping Approach B: groups made with experimental flowable
portions. The upper side of the mold was covered with Mylar SFRCs (10, 20, 30, 40 wt% microfiber) as core materials and
strip and glass slide from both sides of the blade, before covered by 2 mm surface layer of PFC resin.
being exposed to the polymerization light. Upon the removal Approach C: group made with commercial SFRC resin
from the mold, each specimen was polymerized also on the (everX Posterior) as core and covered by 2 mm surface layer
opposite side. The specimens from each group (n = 8) were of PFC resin.
stored dry at 37 ◦ C temperature for 48 h before testing. The Human mandibular 1st molars of similar dimensions
specimens were tested in three-point bending mode, in a (±0.2 mm) (n = 6), which were extracted for periodontal rea-
universal material testing machine at a crosshead speed of sons and free of caries, restorations, abrasions and fractures
1.0 mm/min. were used also as control group.
The fracture toughness was calculated using the Equa- The crown restorations of each group (n = 6) were poly-
tion: Kmax = [P L/B W3/2] f(x), where: f(x) = 3/2 × 1/2 [1.99 − x merized with hand-light curing unit (Elipar S10, 3M ESPE,
(1 − x) (2.15 − 3.93x + 2.7 × 2)]/2(1 + 2x) (1 − x)3/2 and 0 < x < 1 St. Paul, MN, USA) for 40 s for increment (wavelength of the
with x = a/W. Here P is the maximum load in kilonewton (kN), light was between 430 and 480 nm and light intensity was
L is the span length (2 cm), B is the specimen thickness in cen- 1600 mW/cm2 ). Subsequently after polymerization, the crown
d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606 601

Table 2 – The commercial composite resins used and their composition.


Brand Manufacturer Type Composition
G-aenial Anterior GC Corp, Tokyo, Japan Hybrid microfilled UDMA, dimethacrylate co-monomers, prepolymrized
silica and strontium fluoride containing fillers 76 wt%
everX Posteior GC Corp, Tokyo, Japan Short fiber reinforced Bis-GMA, PMMA, TEGDMA, discontinuous E-glass fiber
8.6 wt%, Barium glass fillers 67.7 wt%
PMMA, polymethylmethacrylate; Bis-GMA, bisphenol-A-glycidyl dimethacrylate; TEGDMA, triethylene glycol dimethacrylate; UDMA, urethane
dimethacrylate; wt%, weight percentage.

was fixed to the flat metal base of testing device using double 3.2. Core-crown restorations
side tape before statically loaded (spherical Ø5 mm) (Fig. 1).
The static compressive fracture test was performed using a No statistical (p > 0.05) significance difference in fracture
universal testing machine (model LRX, Lloyd Instruments Ltd., resistance was recorded between tested crown restorations
Fareham, UK) at a speed of 1 mm/min, and data were recorded (Fig. 5). However, failure mode analysis revealed that fracture
using PC software (Nexygen Lloyd Instruments Ltd.). The load- toughness value of the substructure (core material) directly
ing event was registered until restoration fracture. Failure influence the failure mode (Fig. 6). Fracture patterns were ana-
patterns of each loaded restorations were visually analyzed lyzed visually, and showed three various types of fracture
and categorized into three typical fracture patterns: crushing, patterns distributed according to the type of core material:
delamination and ductile. fatal crushing, delamination and ductile fracture (Fig. 6).
Crown specimens having only PFC with no fiber reinforce-
ment showed a fatal crushing fracture pattern. All of the crown
2.6. Statistical analysis specimens that have reinforced core material of Exp-SFRC (10
and 20 wt% microfiber) revealed delaminating of PFC from the
The data were statistically analyzed with SPSS version 23 SFRC substructure layer. While in crown specimens that have
(SPSS, IBM Corp) using analysis of variance (ANOVA) at the core materials of Exp-SFRC (30 and 40 wt% microfiber) and
P < 0.05 significance level followed by a Tukey HSD post hoc commercial short fiber composite (everX Posterior) showed
test to determine the differences between the groups. delamination and ductile (cracking) fracture patterns, which
were similar to natural crown fracture patterns (Fig. 6).

3. Results
4. Discussion
3.1. Flexural strength and fracture toughness
Reports on the properties of currently used flowable PFCs
The mean values of flexural strength and fracture toughness, are thoroughly documented by several authors [20–22]. In
for tested experimental composite materials with standard general, the fracture toughness of all commercial flowable
deviations (SD) are summarized in Figs. 2 and 3. ANOVA PFCs is well below 1.3 MPa m1/2 , and flexural strength below
revealed that fraction of glass fiber/filler had significant effect 160 MPa [20–22]. Flowability of dental composite allows for
(p < 0.05) on the fracture toughness and flexural strength of extrusion through nozzle, simplifying placement and over-
the Exp-SFRC with both monomer systems. Exp-SFRC (50 wt% coming difficulties, namely technique sensitivity, time and
microfiber) had significantly higher mechanical performance expense. Based on this, the present study used microfibers as
of fracture toughness (2.6 MPam1/2 ) and flexural strength fillers in 200–300 ␮m length to maintain the resin flowability.
(175.5 MPa) (p < 0.05) compared to non-fiber reinforced mate- The results of this study support the hypothesis that
rial (1.3 MPam1/2 , 123 MPa). No statistical (p > 0.05) significance microfibers can still have a significant effect on flexural
difference in fracture toughness was recorded between the strength and fracture toughness together with the easy to use
two methacrylate resin formulations (Fig. 3). On the other properties of a flowable material.
hand, groups made of UDMA/PMMA resin matrix showed bet- The flowable Exp-SFRC demonstrated enhanced resistance
ter performance in flexural strength (p < 0.05) (Fig. 2). capability to crack propagation, that is fracture toughness and
The total energy release (work of fracture) of Exp-SFRCs flexural strength, which could be explained by the fiber and
and non-fiber reinforced materials was calculated from the matrix related properties of the material. Fibers in this flow-
area under the load-displacement curves. The contribution of able fibrous composite material are equal to the critical fiber
microfiber plays a major role in increasing the work of frac- length and thus, could efficiently accept the stress transferred
ture energy of Exp-SFRCs. A pronounced increase in the energy from the matrix. This is in accordance with Shouha et al.,
of fracture (from preload to break) as the percentage of glass study, which showed superior flexural properties of short fiber
microfiber was increased. reinforced flowable composite compared to conventional par-
SEM analysis of the tested Single-edge-notched-beam ticulate filler composites [23].
specimens showed protruded (pullout) fiber ends at fracture Aspect ratio, critical fiber length, fiber loading and fiber ori-
surface and microfibers orientation in polymer matrix (Fig. 4). entation are the main factors that could improve or impair
In addition, it presented the good adhesion and wettability of the mechanical properties of SFRC [18]. Aspect ratio is the
microfibers with the resin matrix. fiber length to fiber diameter ratio (l/d). It affects the tensile
602 d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606

200.0
UDMA+PMMA
bisGMA+TEGDMA

160.0

Flexural strength (MPa)


120.0

80.0

40.0

0.0
A 0% B 10% C 20% D 30% E 40% F 50%
Fiber wt%

UDMA 123 (15) 132 (11) 146 (8) 149 (18) 171 (17) 175 (13)
bisGMA 106 (8) 117 (19) 136 (15) 137 (16) 153 (15) 158 (15)
P-value p=0.622 p=0.999 p=0.988

Fig. 2 – Bar graph illustrating means flexural strength (MPa) and standard deviation (SD). Groups joined by a horizontal line
are not significantly difference (p > 0.05). p = 0.988.

3.5
UDMA+PMMA bisGMA+TEGDMA

3
Fracture toughness (MPam½)

2.5

1.5

0.5

0
A 0% B 10% C 20% D 30% E 40% F 50%
Fiber wt%

UDMA 1.3 (0.2) 1.9 (0.4) 2 (0.5) 2.4 (0.5) 2.6 (0.4) 2.6 (0.4)
bisGMA 1.3 (0.2) 1.6 (0.1) 2 (0.2) 2.4 (0.1) 2.6 (0.1) 2.6 (0.2)
P-value p=0.487 p=0.487

Fig. 3 – Bar graph illustrating means fracture toughness (KIC ) and standard deviation (SD). Groups joined by a horizontal line
are not significantly difference (p > 0.05).

strength, flexural modulus and the reinforcing efficiency of between the fibers and polymer matrix increase the criti-
the FRC [19]. Microfibers used in this study have aspect ratio cal fiber length [24]. Sufficient adhesion between fiber and
of more than 30. In order for a fiber to act as an effective matrix provides good load transfer between the two ingredi-
reinforcement for polymers, stress transfer from the poly- ents, which ensures that the load is transferred to the stronger
mer matrix to the fibers is essential [11,19]. This is achieved fiber and this is how the fiber actually works as reinforce-
by having a fiber length equal to or greater than the critical ment. However, if the adhesion is not strong and if any voids
fiber length and the given fiber aspect ratio in range of 30–94 appear between the fiber and the matrix, these voids may
[19]. It has been also concluded that for advanced FRCs, the act as initial fracture sites in the matrix and facilitate the
critical fiber length could be as much as 50 times the diam- breakdown of the material. That is why adhesion between
eter of the fiber. The diameter of glass fibers used in this the fibers and the resin matrix is significant for the mechani-
study is 6 ␮m and the critical fiber length should be, there- cal performance and the longevity of restorations [19]. Earlier
fore, around 300 ␮m. Deteriorated or initially poor adhesion formulations of SFRCs (Alert, Jeneric/Pentron; Nulite F, NSI
d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606 603

Fig. 4 – SEM photomircographs of fracture surface of the experimental flowable SFRC single-edge-notched-beam specimen
(×25 magnification; scale bar = 1 mm) showing pull-out of fibers (A). Random orientation of microfibers (×1000
magnification; scale bar = 10 ␮m) in the resin matrix of the flowable composite (B).

1600

1400

1200

1000
Load (N)

800

600

400

200

0
Gaenial A (PFC) PFC+10% SFRC PFC+20% SFRC PFC+30% SFRC PFC+40% SFRC PFC+everX Teeth

Mean 936 1051 993 994 946 1088 1164


SD 289 189 214 62 156 350 350

Fig. 5 – Compressive load bearing capacity of the crown-shaped specimens and standard deviation (SD). PFC refers to
particulate filler composite used as surface layer. SFRC refers to experimental flowable short fiber composite which was
used as core material. No statistical significance difference was recorded between tested groups (p = 0.829).

Ducle fracture Delaminaon Crushing

100

80

60
%

40

20

0
Gaenial A 10 % 20 % 30 % 40 % everX Teeth
Posterior
Core material

Fig. 6 – Percentage of various fracture patterns of the crown-shaped specimens plotted against to the type of the core
material after the compression load test.
604 d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606

dental) were already commercialized in the late 1990s as filler composite with no fiber reinforcement showed a fatal
packable composites. Although they exhibited acceptable crushing fracture pattern. It can be clearly seen that the brit-
mechanical properties, clinical problems were encountered tleness of the PFC caused the brittle fatal fracture. On the
concerning surface roughness and wear resistance [25,26]. In other hand, all of the crown specimens that have reinforced
addition, a medium-term clinical study by van Dijken et al., core material of Exp-SFRC (10 and 20 wt% microfiber) revealed
revealed a high failure rate for one of these SFRCs (Nulite delaminating of PFC from the SFRC substructure layer. While
F) due to secondary decay and bulk fracture [26]. Alert has crown specimens fabricated of core materials of Exp-SFRC (30
fiber length in micrometer scale (20–60 ␮m) and diameter of and 40 wt% microfiber) and commercial short fiber composite
7 ␮m, which is well below the critical fiber length [27]. This (everX Posterior) showed delamination and ductile (cracking)
explained the difference in fracture toughness values between fracture patterns, which retains the original shape of the
Alert (1.7 MPam1/2 ) and the flowable Exp-SFRC (2.6 MPam1/2 ) crown restoration despite the occurrence of multiple cracks.
resin [27]. Interestingly, these were similar to natural crown fracture pat-
In this Exp-SFRC microfibers were exceptionally well wet- terns (Fig. 6).
ted with the resin (Fig. 4B) and that the adhesion was may be Results indicated that SFRC substructure supports the PFC
better than with previously used fiber systems (everX Poste- layer and serve as a crack prevention and redirection layer.
rior). The better adhesion of the fibers most likely explained Previous studies demonstrated that mechanism of arresting
the good reinforcing effect although the aspect ratio was lower the crack propagation is greatly influenced by the distance
than that of everX Posterior. The toughening mechanisms pro- between the SFRC substructure and the surface where the
vided by the microfibers are result of their ability to deflect stress initiates [30,31]. Thus, highly important is how thick
crack propagation, to bridge, resist the opening and propaga- SFRC and PFC layers are applied. In vitro was observed that
tion of the crack, consequently inducing a closure force on optimal thickness of the veneering PFC composite over the
the crack [28]. Recent review article by Vallittu showed failure SFRC substructure is between 1–1.5 mm [30,31].
types of SFRC composite including cracking of the polymer The contemporary biomimetic restorative approach uses
matrix between fibers, deboning of the fiber and fracture of only three material properties, namely flexural strength, elas-
fiber [19]. tic modulus and thermal expansion coefficient, in order to
The fracture toughness of a material is a measure of how select restorative materials to replace lost tooth tissue [6,7].
well that material hinders the progress of a crack or flaw under Nevertheless, the results of the present study clearly shows
load. Fiber impedes the extension of a crack and develops that fracture toughness is of paramount importance, since
interlocking bridges behind the progressing crack dissipat- this property strongly influences the fracture behavior of den-
ing energy by fiber pullout resulting in graceful rather than tal materials in general and dental restorations in particular.
catastrophic failure (Fig. 4A). This might be due to the random Especially, biomimetic restorations fabricated with a dentine-
orientation of microfibers in resin matrix and forming fiber replacing flowable SFRC or commercial SFRC with a fracture
network (Fig. 4B), which seemed to have enhanced the abil- toughness close to that of dentine, were able to mimic a
ity of the material to resist the fracture propagation, as well more natural fracture behavior. Several earlier studies also
as to reduce the stress intensity at the crack tip from which a revealed that biomimetic or bilayered restorations made of
crack propagates in an unstable manner. As a consequence, an SFRC with a dentine-like fracture toughness exhibited a nat-
increase work of fracture energy and fracture toughness can ural failure behavior and less catastrophic failures [16,31–37].
be expected. Interestingly, addition of microfibers into resin Therefore the authors advocate including fracture toughness
matrix and forming fiber network did not effect on the flowa- as one of the major material properties, besides flexural
bility of the material. To authors’ knowledge, this type of fiber strength, elastic modulus and thermal expansion coefficient,
reinforcing system is not well documented in the literature. for the selection of tooth tissue-replacing materials from a
Experimental SFRC made of UDMA/PMMA resin matrix biomimetic point-of-view.
showed slightly better performance in flexural strength in The fracture resistance values determined by the vari-
comparison with bisGMA/TEGDMA resin matrix (Fig. 2). Which ous investigators were recorded under different measurement
could be due differences between polymer matrices of pure criteria. These criteria were either initial cracking that was
thermoset and semi-IPN. The linear polymer chains of PMMA interpreted as crack development or a reduction in the load
in the cross-linked matrix of UDMA plasticize the polymer by an absolute or relative amount [36,37]. For this study, the
matrix to some extend and increases the flexural strength of maximum static force on the final fracture was determined.
composite resin [29]. Stress applied to the teeth and dental restorations is generally
The crown restorations in this study were designed to eval- low and repetitive rather than being isolated and impactive
uate the failure mode and load-bearing capacity of mandibular in nature. However, because of a linear relationship between
first molar restored with biomimetic approach for the fab- fatigue and static loading, the compressive static test also
rication of direct composite overlay. This design reproduced gives valuable information concerning the fracture behavior
the scenario of major loss of tooth structure. In this series an and load-bearing capacity [37].
attempt was made by using experimental flowable SFRCs as The alteration of diameter of the loading tip i.e. ball size
bulk core material under surface layer of particulate filler com- changes the position of load applications, modifying the con-
posite resin, i.e. biomimetic composite restorations, in order centration of tension patterns [38]. Habekost et al., showed
to emulate a more natural fracture behavior of composite that force required to cause fracture with 10 mm diameter ball
resin restorations. Crown specimens having only particulate was greater than with the 3 mm diameter ball [38]. In this study
5 mm ball size was select to fit the dimension of our crown
d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606 605

restorations. Further research is needed for detailed informa- [13] Garoushi S, Säilynoja E, Vallittu P, Lassila L. Physical
tion on this flowable SFRC material and its influence on many properties and depth of cure of a new short fiber reinforced
clinical applications. composite. Dent Mater 2013;29:835–41.
[14] Lassila L, Garoushi S, Vallittu PK, Säilynoja E. Mechanical
properties of fiber reinforced restorative composite with two
distinguished fiber length distribution. J Mech Behav Biomed
5. Conclusion
Mater 2016;60:331–8.
[15] Garoushi S, Vallittu PK, Watts DC, Lassila LVJ. Polymerization
Based on the results of the present study, one could conclude shrinkage of experimental short glass fiber reinforced
that short glass fibers can significantly reinforce flowable com- composite with semi-inter penetrating polymer network
posite resin and the fracture toughness value of SFRC core matrix. Dent Mater 2008;24:211–5.
material has prior importance, as it influences the failure [16] Garoushi S, Hatem M, Lassila L, Vallittu PK. The effect of
short fiber composite base on microleakage and load
mode and the crack arresting mechanism.
bearing capacity of posterior restorations. Acta Biomater
Odontol Scand 2015;1:6–12.
[17] Bijelic-Donova J, Garoushi S, Vallittu PK, Lassila LVJ.
Acknowledgments Mechanical properties, fracture resistance, and fatigue
limits of short fiber reinforced dental composite resin. J
Testing materials were provided by the manufacturing com- Prosthet Dent 2016;115:95–102.
panies, which is greatly appreciated. This study belongs to the [18] Bijelic-Donova J, Garoushi S, Lassila L, Keulemans F, Vallittu
research activity of BioCity Turku Biomaterials Research Pro- PK. Mechanical and structural characterization of
gram (www.biomaterials.utu.fi) and it was supported by Stick discontinuous fiber-reinforced dental resin composite. J
Dent 2016;52:70–8.
Tech LTD — Member of the GC Group.
[19] Vallittu PK. High-aspect ratio fillers: fiber-reinforced
composites and their anisotropic properties. Dent Mater
references 2015;31:1–7.
[20] Leprince JG, Palin WM, Vanacker J, Sabbagh J, Devaux J,
Leloup G. Physico-mechanical characteristics of
commercially available bulk-fill composites. J Dent
[1] Fennis WM, Kuijs RH, Roeters FJ, Creugers NH, Kreulen CM. 2014;42:993–1000.
Randomized control trial of composite cuspal restorations: [21] Bonilla ED, Yashar M, Caputo AA. Fracture toughness of nine
five-year results. J Dent Res 2014;93:36–41. flowable resin composites. J Prosthet Dent 2003;89:261–7.
[2] Manhart J, Chen H, Hamm G, Hickel R. Buonocore Memorial [22] St-Georges AJ, Swift EJ, Thompson JY, Heymann HO.
Lecture. Review of the clinical survival of direct and indirect Irradiance effects on the mechanical properties of universal
restorations in posterior teeth of the permanent dentition. hybrid and flowable hybrid resin composites. Dent Mater
Oper Dent 2004;29:481–508. 2003;19:406–13.
[3] Wilder AD, May KN, Bayne SC, Taylor DF, Leinfelder KF. [23] Shouha P, Swain M, Ellakwa A. The effect of fiber aspect
Seventeen-year clinical study of ultraviolet-cured posterior ratio and volume loading on the flexural properties of
composite Class I and II restorations. J Esthet Dent flowable dental composite. Dent Mater 2014;30:1234–44.
1999;11:135–42. [24] Karacaer Ö, Polat TN, Tezvergil A, Lassila LVJ, Vallittu PK.
[4] van Dijken JW, Lindberg A. A 15-year randomized controlled The effect of length and concentration of glass fibers on the
study of a reduced shrinkage stress resin composite. Dent mechanical properties of an injection- and a
Mater 2015;31:1150–8. compression-molded denture base polymer. J Prosthet Dent
[5] Craig RG, editor. Restorative dental materials. 10th ed. St. 2003;90:385–93.
Louis, MO: Mosby Publishing Co; 1997. [25] Fagundes TC, Barata TJ, Carvalho CA, Franco EB, van Dijken
[6] Magne P, Belser U. Understanding the intact tooth and the JW, Navarro MF. Clinical evaluation of two packable posterior
biomimetic principle. In: Magne P, Belser U, editors. Bonded composites: a five-year follow-up. J Am Dent Assoc
porcelain restorations in the anterior dentition: a 2009;140:447–54.
biomimetic approach. Chicago: Quintessence Publishing Co; [26] van Dijken JW, Sunnegardh-Gronberg K. Fiber-reinforced
2002. p. 23–55. packable resin composites in Class II cavities. J Dent
[7] Magne P. Composite resins and bonded porcelain: the 2006;34:763–9.
postamalgam era? J Calif Dent Assoc 2006;34:135–47. [27] Garoushi S, Vallittu PK, Lassila L. Mechanical properties and
[8] Manhart J, Kunzelmann KH, Chen HY, Hickel R. Mechanical wear of five commercial fibre-reinforced filling materials.
properties and wear behavior of light-cured packable Chin J Dent Res 2017;20:137–43.
composite resins. Dent Mater 2000;16:33–40. [28] Kim SH, Watts DC. Effect of glass-fiber reinforcement and
[9] Kim KH, Okuno O. Microfracture behaviour of composite water storage on fracture toughness (KIC) of polymer-based
resins containing irregular-shaped fillers. J Oral Rehabil provisional crown and FPD material. Int J Prosthodont
2002;29:1153–9. 2004;17:318–22.
[10] Ruddell DE, Maloney MM, Thompson JY. Effect of novel filler [29] Lastumäki TM, Lassila LV, Vallittu PK. The
particles on the mechanical and wear properties of dental semi-interpenetrating polymer network matrix of
composites. Dent Mater 2002;18:72–80. fiber-reinforced composite and its effect on the surface
[11] Garoushi S, Vallittu PK, Lassila LVJ. Short glass fiber adhesive properties. J Mater Sci Mater Med 2003;14:803–9.
reinforced restorative composite resin with [30] Garoushi S, Vallittu PK, Lassila LV. Use of short
semi-interpenetrating polymer network matrix. Dent Mater fiber-reinforced composite with semi-interpenetrating
2007;23:1356–62. polymer network matrix in fixed partial dentures. J Dent
[12] Garoushi S, Tanner J, Vallittu PK, Lassila LVJ. Preliminary 2007;35:403–8.
clinical evaluation of short fiber-reinforced composite resin [31] Keulemans F, Lassila LVJ, Garoushi S, Vallittu PK, Kleverlaan
in posterior teeth: 12-months report. Open Dent J CJ, Feilzer AJ. The influence of framework design on the
2012;6:41–5.
606 d e n t a l m a t e r i a l s 3 4 ( 2 0 1 8 ) 598–606

load-bearing capacity of laboratory-made inlay-retained reinforced cusp-replacing composite restorations. Dent


fibre-reinforced composite fixed dental prostheses. J Mater 2005;21:565–72.
Biomech 2009;42:844–9. [36] Keulemans F, Van Dalen A, Kleverlaan CJ, Feilzer AJ. Static
[32] Garoushi S, Lassila LVJ, Vallittu PK. Fiber-reinforced and dynamic failure load of fiber-reinforced composite and
composite substructure: load bearing capacity of an onlay particulate filler composite cantilever resin-bonded fixed
restoration. Acta Odontol Scand 2006;64:281–5. dental prostheses. J Adhes Dent 2010;12:207–14.
[33] Fráter M, Forster A, Keresztúri M, Braunitzer G, Nagy K. [37] Garoushi S, Lassila LVJ, Tezvergil A, Vallittu PK. Static and
In vitro fracture resistance of molar teeth restored with a fatigue compression test for particulate filler composite
short fibre-reinforced composite material. J Dent resin with fiber-reinforced composite substructure. Dent
2014;42:1143–5. Mater 2007;23:17–23.
[34] Dere M, Ozcan M, Göhring TN. Marginal quality and fracture [38] Habeksot LV, Camacho GB, Pinto MB, Demarco FF. Fracture
strength of root-canal treated mandibular molars with resistance of premolars restored with partial ceramic
overlay restorations after thermocycling and mechanical restorations and submitted to two different loading stresses.
loading. J Adhes Dent 2010;12:287–94. Oper Dent 2006;31:204–11.
[35] Fennis WM, Tezvergil A, Kuijs RH, Lassila LV, Kreulen CM,
Creugers NH, et al. In vitro fracture resistance of fiber

You might also like