You are on page 1of 9

G Model

JJOD 2645 No. of Pages 9

Journal of Dentistry xxx (2016) xxx–xxx

Contents lists available at ScienceDirect

Journal of Dentistry
journal homepage: www.intl.elsevierhealth.com/journals/jden

Mechanical and structural characterization of discontinuous


fiber-reinforced dental resin composite
Jasmina Bijelic-Donova, CDT DDSa,* , Sufyan Garoushia , Lippo V.J. Lassilaa ,
Filip Keulemansb , Pekka K. Vallittua,c
a
Department of Biomaterials Science and Turku Clinical Biomaterials Centre-TCBC, Institute of Dentistry, University of Turku, Itäinen Pitkäkatu 4 B, 20520
Turku, Finland
b
Dental Materials Science, Dental School, Ghent University, De Pintelaan 185/P8, B-9000 Gent, Belgium
c
City of Turku Welfare Division, Oral Health Care, Lemminkäisenkatu 2, 20520 Turku, Finland

A R T I C L E I N F O A B S T R A C T

Article history: Objectives: This study evaluated several fiber- and matrix related factors and investigated different
Received 14 March 2016 mechanical properties of discontinuous i.e. short fiber-reinforced composite (SFRC) (everX Posterior,
Received in revised form 18 July 2016 eXP). These were compared with three conventional composites, microfilled G-ænial Anterior (GA),
Accepted 20 July 2016
nanofilled Supreme XTE (SXTE) and bulk-fill Filtek Bulk-Fill (FBF).
Available online xxx
Methods: Fracture toughness (KIC), flexural strength (FS), flexural modulus (FM), compressive strength
(CS), diametral tensile strength (DTS), apparent horizontal shear strength (AHSS) and fracture work (Wf)
Keywords:
were determined for each composite (n = 8) stored dry or in water. SEM analysis of the fiber diameter (df)
Short fiber-reinforced composite
Critical fiber length
(n = 6) and orientation (n = 6) were performed. The theoretical critical fiber length (lfc) and the aspect
Volume fraction ratio (l/d) of SFRC were calculated and the volume fraction of discontinuous fibers (Vf%) and the fiber
Fiber orientation length (lf) of SFRC were evaluated. The results were statistically analyzed with two-way ANOVA
Toughening mechanism (a = 0.05).
Aspect ratio Results: The mechanical properties of SFRC (eXP) were generally superior (p < 0.05) compared with
conventional composites. GA had the highest FM (p > 0.05), whereas FBF had the highest AHSS (p < 0.05).
The fiber related properties Vf%, l/d, lf, lfc and df of eXP were 7.2%, 18–112, 0.3–1.9 mm, 0.85–1.09 mm and
17 mm respectively. SEM results suggested an explanation to several toughening mechanisms provided
by the discontinuous fibers, which were shown to arrest crack propagation and enable a ductile fracture.
Water exposure weakened the mechanical properties regardless of material type. Wf was unaffected by
the water storage.
Conclusion: The properties of this high aspect ratio SFRC were dependent on the fiber geometry (length
and orientation) and matrix ductility.
Clinical significance: The simultaneous actions of the toughening mechanisms provided by the short fibers
accounted for the enhanced toughness of this SFRC, which toughness value matched the toughness of
dentin. Hence, it could yield an inherently uniform distribution of stresses to the hard biological tissues.
ã 2016 Elsevier Ltd. All rights reserved.

1. Introduction high strength performance of the composite [1] and the fiber
length above its critical value [2]. Indeed, some SFRCs contain
Composites reinforced with discontinuous fibers are known as microfiber fillers (60–120 mm) [3,4], which are well below the
short fiber-reinforced composites (SFRC). Earlier attempts to critical fiber length (0.5–1.6 mm) [2]. Due to the insufficient fiber
improve the mechanical properties of composites by incorporating length, micrometer fibers act as microfiller particles instead of
short fibers had limited success. This was due to the difficulty of fiber filler. This gives the SFRC properties comparable with
maintaining their high aspect ratio (l/d), which is important for the particulate filler composites (PFCs) [5,6].
Interpretation of the properties of SFRCs is more difficult than is
for unidirectional fibers and composites reinforced with them. This
* Corresponding author at: TCBC, Institute of Dentistry, University of Turku, could be due to the fact that unidirectional fibers are assumed to be
Turku, Finland. Itäinen Pitkäkatu 4 B, FI-20520 Turku, Finland. straight, perfectly circular at the cross-section, to have same mean
E-mail addresses: jasmina.bijelic@utu.fi, jasminabijelic@gmail.com diameter [7] and to be closely packed [8]. In contrast, short fibers
(J. Bijelic-Donova).

http://dx.doi.org/10.1016/j.jdent.2016.07.009
0300-5712/ã 2016 Elsevier Ltd. All rights reserved.

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

2 J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx

and Vf, equally important is the length of the short fibers. Indeed,
Nomenclature
these factors together determine the fiber debonding from the
matrix during loading of the SFRC structure. As an example, the
eXP or SFRC Millimeter-scale short fiber-reinforced com- debonding for longer aligned fibers is greater than for random
posite or high aspect ratio short fiber-rein- ones, which undergo only limited debonding. In other words, the
forced composite load bearing capacity of SFRCs is more effective with short random
PFC Particulate filler composite than with longer aligned short fibers [15]. Furthermore, favorably
GA or MFR G-ænial Anterior, microfilled hybrid restorative oriented fibers would allow the bridging phenomenon, which is
composite material important fiber toughening mechanism. Fiber bridging is most
SXTE or NF Supreme XTE, nanofilled restorative composite effective for transversally aligned fibers, intermediate for random-
material ly orientated fibers and absent for non-reinforced systems. Fiber
FBF Filtek Bulk Fill, flowable bulk fill restorative breakage, on the other hand, is important as it introduces new
composite material length in the system, while the Vf remains unchanged [14].
KIC Fracture toughness Apparently, among the factors that determine the mechanical
FS Flexural strength performance of SFRC are the length [2,16], type, arrangement and
FM Flexural modulus distribution of the fibers [1]. Other factors, known as intrinsic
CS Compressive strength (fiber related) and extrinsic (matrix related) factors, also affect the
DTS Diametral tensile strength SFRC’s properties [17]. The fiber related factors of SFRCs could be
AHSS Apparent horizontal shear strength considered to be the lf, df, Vf, strength as well as the interfacial
Wf Fracture work fiber-matrix bonding, whereas the matrix related factor could be
SEM Scanning electron microscopy considered to be the FOD [17]. In addition, the random distribution
FOD Fiber orientation distribution of short fibers could be compromised or intentionally affected
Wt% Weight fraction during the manipulation process. If discontinuous fibers are long
Vol% Volume fraction enough, they behave as long fibers giving anisotropic mechanical
Vf% Fiber volume fraction properties [10,11,18]. To date, there exist some research on the
df Fiber diameter mechanical properties of recently introduced millimeter-scale
lf Fiber length discontinuous fiber-reinforced composite (SFRC) (everX Posterior,
lfc Critical fiber length GC Corp., Tokyo, Japan) [18–23], but none on its fiber related and
lfo Optimal fiber length matrix related factors. Based on its fiber l/d, this SFRC could be
l/d Aspect ratio classified as high aspect ratio SFRC i.e. millimeter-scale SFRC.
Consequently, the present study was designed to firstly
investigate several mechanical properties of this millimeter-scale
SRFC in comparison with three conventional PFCs (G-ænial
are often assumed to be randomly distributed, to have a cylindrical Anterior, GC; Supreme XTE, 3M Espe, MN, USA and Filtek Bulk Fill,
shape and an elliptical cross section [9]. Nevertheless, the short 3M Espe) (Table 1) and secondly to evaluate both fiber related and
fibers could also be transversally aligned (i.e. in plane). The matrix related factors of the SFRC. The selected properties were
mechanical properties of SFRCs essentially depend on the fiber fracture toughness (KIC), flexural strength (FS), flexural modulus
orientation distribution (FOD), aspect ratio (l/d) and their volume (FM), compressive strength (CS), diametral tensile strength (DTS),
fraction (Vf) [10,11]. There is, however, critical l/d for the apparent horizontal shear strength (AHSS) and fracture work (Wf).
transversally orientated short fibers above which further improve- The fiber related factors of the millimeter-scale SFRC were
ment of their mechanical properties declines [10]. For engineering analyzed on both a theoretical basis and experimentally. On a
SFRC this is between 6 and 9 [10], whereas the lowest l/d for dental theoretical basis, the volume fraction of the short fibers (Vf%) and
SFRC is 5.2 [12]. With increasing l/d they start behaving as long the theoretical critical fiber length (lfc) were calculated. Experi-
fibers [10,11], which macroscopically are transversally isotropic mentally determined fiber related factors were the fiber diameter
materials [13], although in an x-y system they show anisotropic (df), the fiber length (lf) and the aspect ratio (l/d), whereas the
properties [14]. Similarly, transversally aligned short fibers have experimentally determined matrix related factor was the fiber
also been described as macroscopically transversally isotropic orientation distribution (FOD). The null hypotheses were that the
materials [11]. By contrast, the properties of randomly orientated mechanical properties of the millimeter-scale SFRC will not differ
discontinuous fibers are mostly affected by their Vf [10]. Thus, from the PFCs’ properties and that short-term water storage would
these are macroscopically isotropic materials, with most effective not affect the mechanical properties of the investigated materials.
mechanical properties at Vf of 10% [11]. Consequently, if the fibers
are homogeneously distributed with random orientation, the SFRC 2. Materials and methods
is macroscopically homogenous and isotropic, whereas if the fibers
are homogeneously distributed with transverse (in plane) orien- FS, FM, DTS, CS, AHSS and KIC were evaluated. Eight specimens
tation, the SFRC is homogeneous and anisotropic i.e. transversally (n = 8) were prepared per restorative composite material listed in
isotropic. Furthermore, in order to evaluate various physicome- Table 1 for each test type and each storage condition. The storage
chanical properties of the SFRCs on macroscopic level, numerous conditions were dry and in water at 37  C for 7 and 30 days,
theoretical, computational based models have been proposed. respectively. Testing procedures were performed at room temper-
Such models are mechanical-based micromechanical numerical ature (23  1  C). The fiber diameter (n = 6) and orientation (n = 6)
model [14], energy-based homogenization model [15] and were determined immediately upon preparation of the specimens.
mesostructured with work-energy approach model [13]. These The fiber volume fraction, fiber length, critical fiber length and
models give an insight on what circumstances initiate fracture in aspect ratio were additionally computed.
the SFRC structure, what factors influence its progression, what FS and FM were evaluated by the three-point bending test.
causes the fiber debonding or breakage, how the FOD affects the Specimens were prepared, polymerized and tested following the
stiffness of the SFRC system and how FOD interacts with the protocol previously described [19].
fiber l/d. These models have shown that in addition to the FOD, l/d

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx 3

Table 1
Materials used in the study.

Brand Type Lot nos Filler Resin Filler load Wt%


*
everX Posterior Millimeter-scale SFRC Silicon dioxide (max. 5 wt%), Barium glass (max. 70 wt%) E-glass Bis-GMA, Average filler load
(GC Corp., short fiber- fiber (max. 15 wt%) TEGDMA, 67.7
Tokyo, Japan) reinforced PMMA Average glass fiber
composite load 8.6
G-ænial Anterior Microfilled resin MFR 1305311 Pre-polymerized: silica containing & strontium and lanthanoid UDMA, 70–80
(GC Corp., hybrid fluoride containing filler 16–17 mm, Silica filler >100 nm and dimethacrylate
Tokyo, Japan) Fumed silica filler <100 nm co-monomers
Filtek Nanofilled NF N649930 Silica: 5–20 nm nanoparticle Bis-GMA, Silica nanoparticle 8
Supreme XTE Zirconia/silica: 0.6–1.4 mm nanocluster UDMA, Bis- Zirconia/silica
(3M Espe, St. EMA, TEGDMA nanocluster 71
Paul, MN, USA)
Filtek Flowable bulk fill FBF N600679 Zirconia/silica: 0.01–3.5 mm Bis-GMA, 64.5
Bulk Fill Ytterbium trifluoride (YbF3): 0.1–5.0 mm UDMA, Bis-
(3M Espe, St. EMA, Procrylat
Paul, MN, USA)

Bis-GMA: bisphenol-A-diglycidyl dimethacrylate; TEGDMA: triethylene glycol dimethacrylate; PMMA: polymethyl methacrylate; UDMA: urethane dimethacrylate; Bis-EMA:
bisphenol-A-polyethylene-glycol-diether dimethacrylate.
*
Eight (8) batches of everX Posterior were used, 1301232, 1301252, 1306101, 1306262, 1308261, 1309191, 1307082 and 1308292.

Cylindrical specimens, 6 mm in height and 4 mm in diameter, the adapted ISO 20795-1 standard method [26], and as previously
were used to evaluate DTS and CS. The fabrication and testing described [19,20]. The SENB specimens were used for determining
procedures of the specimens are described elsewhere [19]. DTS in the total fracture work (Wf) in J/m2, by using Eq. (4):
megapascals (MPa) was calculated using Eq. (1) [24]:
Wf = U/[2B (H  a)] 1000 (4)
DTS= 2 F/p l D (1)
where Wf is the fracture work, U is the recorded area under the
where: DTS is the diametral tensile strength, F is the maximum load-deflection curve and represents the energy required to break
R
applied load in newtons (N); D is the diameter of the specimens in the whole specimen. U = PdD in newton millimeters (Nmm),
mm and l is the length of the specimen in mm. CS was calculated in where D is the measured deflection for load P in N; B is the
megapascals (MPa) using Eq. (2) [24]: specimen width in mm, H is the specimen height in mm, and a is
the crack length in mm.
CS = 4 F/p D2 (2)
Fiber volume fraction (Vf %) was calculated using Eq. (5) [27]:
where: CS is the compressive strength, F is the maximum applied
Wf =rf
load in N and D is the diameter of the specimen in mm. Vf % ¼  100 ð5Þ
Wf =rf þWm =rm
AHSS was determined by the short-beam method, according to
ASTM D 4475–96 [25]. A translucent mold with an inner diameter where W is the weight fraction (wt%), r is the density, f represents
of 3.6 mm and a length of 14.5 mm (0.2 mm), which is one the fiber and m the matrix. Wf = 8.6 wt% (Wf of the short E- glass
diameter greater than the test span was used to fabricate the rod fibers in everX Posterior) and rf = 2.56 g/cm3 [27]. The matrix
shaped specimens. The mold was placed on a Mylar-strip-covered comprises resin and particulate fillers with r(resin) = 1.23 g/cm3
glass slide and the restorative material was packed into the mold. [28], W(resin) = 23.7 wt%, r(particulate filler) = 2.8 g/cm3 and w(particulate
The top surface was flattened with Mylar strip and glass slide and filler) = 67.7 wt%.
each specimen was polymerized on both sides for 20 s by five The fiber length (lf) was measured using the direct technique, by
overlapping irradiations with a light-curing polymerization unit separating the fibers from the matrix, following the technique
(Elipar S10, 3M ESPE, St. Paul, MN, USA) at an intensity of previously described [20]. Eight batches of millimeter-scale SFRC
1600 mW/cm2 and wavelength ranging from 430 to 480 nm. The (everX Posterior) were evaluated. Fibers were deposited onto glass
irradiance of the light-curing polymerization unit was measured slide and examined with light microscopy (Leica; Leica Mycro-
with dental radiometer (Demetron L.E.D. Radiometer, Kerr, systems, Wild, Heerbrugg, Switzerland). The average fiber length
Middleton, WI, USA). Upon removal from the mold, each was then defined as LN = S Ni Li/S Ni, where LN is the number
longitudinal side was further polymerized for 20 s in five average of the fiber length, Ni is the number of fibers with length Li
overlapping places. The rod specimen was placed with its long [27]. Four thousand fibers were measured in total (500/batch).
axis perpendicular and its midpoint centered to the loading nose. When the fiber diameter (df), the ultimate tensile strength (sfu) of
The side supports were adjusted to a span length of 11 mm the glass fiber and the interfacial shear strength (ti) of the glass
(0.2 mm) and specimens were loaded by a universal material fiber reinforced composite are known, then the theoretical critical
testing machine (Lloyd model LRX, Lloyd Instruments Ltd., fiber length (lfc) can be calculated from lfc = sfudf/2 ti[29]. The
Fareham, UK) at a crosshead speed of 1.0 mm/min until failure. tensile strength of E-glass fiber is estimated at 2.0 GPa, whereas the
The diameter of each specimen was measured prior to testing with interfacial shear stress (ti) for a glass fiber reinforced composite
a digital caliper and the apparent horizontal shear strength in MPa varies between 17.6 and 33.8 MPa depending on the fiber
was calculated in accordance with Eq. (3): conditioning and the testing environment [30], but typically is
20 MPa [27]. The diameter of the individual glass fiber in our study
AHSS= 0.849 P/d2 (3)
was 17 mm, whereas the interfacial shear stress evaluated by
where AHSS is the apparent horizontal shear strength; P is the macroscopic short-beam test was 15.6 MPa. This experimentally
breaking load in N and d is the specimen diameter in meters (m). obtained result was then compared with the typical values
Fracture toughness (KIC) was evaluated using single-edge presented in the literature [27].
notched bend (SENB) specimens (2.5  5  25 mm) according to

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

4 J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx

In the present investigation the fiber orientation distribution Table 3


Fiber related properties of the millimeter-scale SFRC (everX Posterior) investigated
(FOD) was visually studied by sectioning the SENB specimens (12
in this study.
slices obtained from n = 6) along their long-axis plane with a high
performance precision cut-off saw at 300 rpm (Struers Secotom 50, Vf % df (mm) lfc (mm) lfc (mm) lfo lf (mm) l/d
(mm)
Struers, Ballerup, Denmark) and analyzing their images by
scanning electron microscopy (SEM, JSM 5500, Jeol Ltd., Tokyo, everX 7.2 17 0.85a 1.09b 1.0–1.3c 0.3–1.9 18–
Posterior 112
Japan). The specimens were ground wet (Stuers LabPol-21, Stuers
A/S, Copenhagen, Denmark) with silicon carbide papers of Vf%: volume fraction of short fibers in SFRC (everX Posterior); df: fiber diameter; lfc:
decreasing abrasiveness (1000-, 1200-, 4000-grit) and gold sputter critical fiber length; lfo: optimal fiber length; lf: fiber length and l/d (measured as lf/
df): aspect ratio.
coated before the SEM examination. After testing the specimens, a
Interfacial shear stress (ti) for a typical glass fiber reinforced polymer composite
their fracture surfaces were examined with SEM. is 20 MPa [27].
The fiber diameter (df) was also determined from SEM images, b
Interfacial shear stress (ti) evaluated by the macroscopic short-beam test used
which were processed with the Image J program (National in the present investigation.
c
lfo is the optimal fiber length measured according to lfo  1.2lfc [46].
Institutes of Health, USA). A cross section of a fiber bundle was
examined with SEM for this purpose. The fiber diameter was
determined as a mean value calculated from one hundred
0.3 mm and 1.9 mm; 71% of the fiber lengths varied between
measurements per specimen (n = 6).
0.4 mm and 1.0 mm, 11% were between 1.1 mm and 1.9 mm and 18%
The data were statistically analyzed with SPSS version 23 (SPSS,
were 0.3 mm short E-glass fibers. The average length of the short
IBM Corp.). Homogeneity of variance was analyzed by the Levene
fibers varied among the batches. The shortest average fiber length
test (p = 0.05). Two-way ANOVA (a = 0.05) followed by a Tukey post
was 0.7 mm and the longest average fiber length was 0.9 mm. The
hoc test was used to determine if the mechanical properties (FS,
critical fiber length was between 0.85 mm and 1.09 mm. The
FM, DTS, CS, AHSS, KIC, Wf) were different for the materials
theoretical value of 0.85 mm was obtained on a theoretical basis by
investigated in this study (SFRC, MFR, NF, FBF) and for the type of
using the value of 20 MPa for the interfacial shear stress reported in
storage condition (dry, water). Pearson correlation was used to
the literature. When the average value derived from our
analyze the correlation among the mechanical properties.
macroscopic AHSS test was used, a critical fiber length value
1.09 mm was obtained.
3. Results
The SEM investigation (Figs. 1 and 2) of the fiber orientation
distribution showed the orientation paths of the short fibers
Table 2 displays the results for the mechanical properties (FM,
around the crack notch, which are also presented schematically in
FS, DTS, CS, AHSS, KIC and Wf) investigated in this study. In general,
Fig. 3. The short fibers were aligned a. parallel to the crack notch,
water storage weakened the mechanical properties of all materials.
which is along to the long axis of the crack notch (perpendicular to
This reduction significantly affected the FS regardless of the
the pressing direction); b. normal to the crack notch; c. transverse
material type (p = 0.001), whereas the water effect on FM, DTS,
(perpendicular) to the long axis of the crack notch; and d. a radial
AHSS and KIC was material dependent. The material type also had a
fiber alignment was observed at the crack notch tip. The radial
major effect on the mechanical properties. An exception was the
alignment of the fibers was obtained as the paths of transversally
CS, which was not influenced by the material type (p = 0.947). The
and normally oriented fibers cross each other (Fig. 2). The SEM
millimeter-scale SFRC (everX Posterior) had significantly higher KIC
analysis of the tested SENB specimens (Fig. 4) showed that fibers
for both storage conditions (p = 0.001), compared to conventional
bridged the crack. Fiber pullout and breakage were also observed.
PFCs. KIC of the millimeter-scale SFRC insignificantly decreased
The pullout effect occurred only for the shorter segment of the
(p = 0.051) after water storage. The millimeter-scale SFRC had
embedded fiber length, which is typical in a well-bonded system.
significantly higher Wf (p = 0.001), for both storage conditions
The longer fibers, on the other hand, remained intact in the matrix.
(p = 0.001), than the conventional PFCs.
The SEM analysis of fractured SENB specimens (Fig. 5) showed
Strong, statistically significant correlation was observed only
pulled out and protruded fiber ends. Protruded fibers ends were
between the KIC and the Wf (r2 = 0.816), and moderate between KIC
covered with resin, which also indicates a well-bonded system. In
and FS (r2 = 0.475), KIC and DTS (r2 = 0.440) as well as between FM
addition, imprints of debonded fibers were observed in the resin
and AHSS (r2 = 0.568).
matrix.
Table 3 displays the results for the Vf%, df, lf, lfc, lfo (lfo is the
Fig. 6 shows typical load-deflection curves for the tested
optimal fiber length) and l/d of the short fibers within the
materials. For the millimeter-scale SFRC multiple cracking of the
millimeter-scale SFRC. The Vf% of the short fibers was 7.2%. The
matrix occurred during the non-elastic range, which is the portion
average fiber diameter was 17 mm. The fiber length varied between
between the first cracking and the peak. The brittle matrix of the

Table 2
Mechanical properties of investigated materials.
1
Material Storage FM (GPa) FS (MPa) DTS (MPa) CS (MPa) AHSS (MPa) KIC (MPa m /2) Wf (J/m2)
MFR dry 15.0a#(1.7) 90a*(7) 39a#(4) 233a*(32) 14.1a#(0.8) 0.9a#  0.1 55a*(9)
water 14.4a#(2.0) 78a^(6) 38a#(5) 174a^(39) 13.6a#(0.6) 0.8a#  0.1 164a^(35)
NF dry 3.9b#(1.0) 104b*(11) 37a*(3) 219a#(79) 18.4bc*(5.2) 1.1a#  0.1 53a#(7)
water 3.5b#(0.6) 64b^(8) 29b^(5) 191a#(60) 16.0b^(2.5) 0.9a#  0.1 146a#(18)
FBF dry 5.7b#(1.2) 119c*(13) 43ab#(8) 217a#(25) 19.8c#(2.1) 1.1a#  0.1 88a*(14)
water 4.7b#(0.9) 83a^(7) 39a#(4) 200a#(34) 19.9c#(1.1) 0.9a#  0.1 203a^(9)
SFRC dry 12.6a*(2.8) 119c*(5) 46b*(5) 235a#(33) 15.6ab#(1.3) 2.4b*  0.5 546b#(241)
water 9.8c^(1.0) 85a^(8) 39a^(3) 190a#(31) 14.7ab#(1.5) 1.7b^  0.2 580b#(102)

Same superscript letter and sign above the values indicates groups that were statistically similar (p > 0.05). Letters describe the statistical difference among the materials
across each storage medium, whereas the signs represent the statistical difference between dry and water storage mediums across each material category level.

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx 5

The effect of water was deterioration for all mechanical


properties evaluated at one month aging time for all investigated
composite materials, which is in good agreement with the
literature findings [31–33]. Firstly, the hydrolytic degradation in
water is presumably due to the hydrolysis of the silane coupling
agent on the surfaces of both E-glass fiber and particulate fillers.
Consequently, hydrolytic degradation of the polysiloxane network
between the E-glass and the matrix, as well as between the filler
and matrix occur [34]. Furthermore, water uptake is a slow, time-
dependent process, and thus the aging time of the composite in
water affects the saturation level. The period of one month could
be considered as a short-term aging. In short-term aging time, the
specimens do not sufficiently equilibrate with water. Water
absorption reduces the mechanical properties of the composite
material until complete water saturation of the composite material
has been reached, which is approximately 4–6 weeks [32]. This is
due to the plasticization effect of water [34]. After this, however,
further reduction stops and in the long term, the mechanical
properties remain unchanged [35].
The fracture work was the only property which increased
following water storage irrespectively of the material. Although it
cannot be fully explained, this could in part be due to the water
used as storage medium, which allows plasticization of the
polymer network. Indeed, the presence of small plasticizing
molecules, such as water, may induce some flexibility of the
composite matrix and consequently, it may allow pronounced
bending of the water stored specimens during testing. However,
the fracture work depends also on the microstructural filler
Fig. 1. SEM picture of the fiber orientation distribution. Black arrows indicate the
fiber orientated transverse (perpendicular) to the long axis of the crack notch, white parameters (size or shape) [36] that could be affected by the plastic
arrows indicate fiber orientation parallel to the crack notch and red arrows indicate deformation of the matrix around them.
fiber orientation in plane normal to the crack notch. (For interpretation of the Likewise, filler related properties such as type, size, shape and
references to colour in this figure legend, the reader is referred to the web version of load are important for the overall properties of the composite
this article.)
materials. Higher filler loads are generally associated with
improved properties. Nonetheless, critical filler level exists above
which any addition of filler could include more defects into the
conventional PFCs cracked at the first peak. This was followed by
resin matrix and might have an inverse effect on the mechanical
composite fracture.
properties. The critical filler load level has been reported to be
approximately 60 vol% (about 80 wt%) [37]. The optimal SFRC
4. Discussion
properties, on the other hand, are reached at fiber filler load level of
10 vol% [11]. Filtek Bulk Fill had lowest filler fraction among the
Investigated properties of the millimeter-scale SFRC differed
studied materials, whereas the inorganic fraction of the other
from the properties of conventional composites and all properties
restorative materials was at or approached the critical filler load
were affected by the storage medium, regardless of the material
level (Table 1).
type. Thus, both null hypotheses were rejected.

Fig. 2. SEM picture of the fiber orientation distribution around the crack notch. X-fiber orientation enabled by crossing paths of normally orientated (red arrow) and
transversely orientated (black arrow) fibers. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

6 J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx

The nanofilled composite in this study had the lowest FM,


comparable mechanical properties to the other investigated PFCs
and weaker mechanical properties compared to the millimeter-
scale SFRC. The water induced interfacial debonding was shown to
reduce the strength and the fracture resistance of the nanofilled
composite investigated here [38,39]. This is consistent with the
findings of the present study. Deterioration of the mechanical
properties of nanofilled (NF) composite (Supreme XTE) is due to
the loss of the structural integrity of the nanocluster, the increased
Fig. 3. Schematic drawing of the fiber orientation around the crack notch. Red line filler surfaces and the large intersurface area between filler and
shows fiber orientation in plane normal to the crack notch and green parallel to the organic matrix, which is the weakest link in the composite material
crack notch. Blue line shows the fiber orientated transverse (perpendicular) to the
long axis of the crack notch. Grey lines show the radial orientation of the fibers
and has high potential to hydrolytic degradation [40,41]. The
around the crack tip. (For interpretation of the references to colour in this figure properties of this nanofilled composite were shown to be related to
legend, the reader is referred to the web version of this article.) the inherent porosity and the inconsistency of the nanoclusters
[40].
Except for the FS, the bulk fill composite evaluated in this
investigation, revealed comparable mechanical properties to the
NF composite, but less optimal mechanical properties compared to
the millimeter-scale SFRC and the microfilled hybrid (MFR)
composite. This is in good accordance with other studies
[21,42,43], which showed that this could be due to the different
chemistry of the resin matrix of Filtek Bulk Fill that encompasses
Bis-EMA and UDMA monomer mixture [42] accompanied with the
presence of plastifying monomers [21]. Both constituents contrib-
ute to the reduction of the shrinkage stress [21,42]. However, the
lower filler fraction and the lower FM [43] could explain the
weaker mechanical properties of this bulk-fill composite.
The microfilled hybrid composite (MFR) in this study showed
good mechanical properties, which could be related to the
presence of pre-polymerized particle fillers, which are polymer-
ized and milled polymers heavily loaded with submicron sized
filler particles. This allows for improved resistance to fracture even
if the filler load of the composite is low, and is due to three effects
known as filler addition, stress relaxation and the branching effect
[44].
Fig. 4. SEM picture of tested SENB specimen. Fiber bridging phenomenon (black
Generally, the millimeter-scale SFRC performed better in
arrow), fiber breakage (red arrow) and fiber pull out (white arrow) were observed. comparison to the PFCs. High fracture toughness values were
(For interpretation of the references to colour in this figure legend, the reader is observed for this short fiber composite in this and other studies
referred to the web version of this article.) [19,20]. The millimeter-scale short fibers and the semi-IPN (semi
interpenetrating network) structure are the two main factors that
toughen this particular SFRC. The toughening mechanisms
provided by the millimeter fibers are a result of their ability to
stretch, to deflect crack propagation, to bridge, resist the opening
and propagation of the crack, consequently inducing a closure
force on the crack [45] as well as to reduce the stress intensity at
the crack tip. The stretching capacity of fibers may enable the crack
bridging and the crack blunting mechanisms to occur (Fig. 4).
During the crack-bridging, discontinuous fibers would probably
stretch between the edges of the propagating crack decreasing the
notch sensitivity and causing blunting of the initially sharp crack.
This would lower the stress concentration at the crack tip and
possibly slow down or impede the crack propagation. However, as
the millimeter-scale SFRC contains range of short fiber lengths, it is
expected for the shorter fiber segments to pull out and for the
longer segments to actively carry the bridging. In addition, the
pulled out fibers will have various lengths because the short fibers
have different embedded lengths, and the bridging will be
sustained only by the longer bridging fibers [17]. These longer
fibers will be responsible for increased fractured energy [46]. In
addition, the gradual fiber length might contribute to the
toughening effect. Indeed, recent study showed that dual fiber
Fig. 5. SEM picture of fractured surface of tested SENB specimen. Protruded fiber length distribution enhanced SFRC properties [47]. The silanization
covered with resin (black arrow), pulled out fiber ends (white arrow) and in plane of the fibers with methacrylate-based silane, compatible with both
normal to the crack notch orientated fiber (red arrow). (For interpretation of the methacrylates and dimethacrylates, and the presence of PMMA
references to colour in this figure legend, the reader is referred to the web version of
(polymethyl acrylate) within the semi-IPN structure, which
this article.)
reduces the stiffness of the cross-linked resin monomer, should

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx 7

Fig. 6. Typical load-deflection curves of investigated materials. 1: everX Posterior; 2: Supreme XTE; 3: G-ænial Anterior and 4:Filtek Bulk Fill.

be also mentioned as factors contributing to the toughening effect. findings [23,50]. Furthermore, the lower limit for the fiber aspect
Hence, it becomes clear that the different toughening mechanisms ratio in dental SFRC is 5.2 [12]. Consequently, it could be concluded
operate simultaneously and lead to improvement in the toughness that the upper limit for the length of the short fiber within SFRC
of this SFRC. In addition, the toughness of the investigated SFRC systems could be considered 2 mm, optimal fiber quantity could be
matches the toughness values reported for the dentin [48], which 10 Vf%, and lower limit for fibers aspect ratio 5.2. In the present
is important, because inherently the distribution of the stresses study, the gradual discontinuous fibers lengths for the everX
between the biological hard tissue and the restorative material Posterior were measured between 0.3 mm and 1.9 mm, which
substituting it would be uniform. Moreover, both have common makes the lowest aspect ratio 18 and greatest 112 (Table 3). The
toughening mechanisms such as crack deflection and the crack critical fiber length for the everX Posterior was measured to be
bridging phenomena [48]. Structural similarity between the dentin between 0.85 and 1.09 mm, which is in the range of the reported
and the millimeter-scale SFRC is also observed. The collagen fibers critical fiber length [2]. This critical length of the fibers is
provide the toughness of dentin similar to short fibers, which important, because short fibers tend to pullout. For the longer
provide the toughness of the millimeter-scale SFRC. embedded length, only the short segment will pull out, as seen in
Together, the fracture work and the fracture toughness, depict Fig. 4. Pulled out fibers indicate that the composite continues to
the fracture resistance ability of the material. As these properties carry the load, although its strength does not increase. Neverthe-
were significantly improved for the investigated SFRC, it can be less, the composite’s fracture toughness, at this stage, still
anticipated that the endurance of the material in static or fatigue continues to increase substantially [1]. If, however, the fiber
load could also be enhanced. However, it should be born on mind length is twice the critical length, then the fibers will break and the
that this fibrous material should clinically be protected with PFC bridging energy will decline. Thus, an optimum fiber length should
layer, and the fatigue endurance of the material combination will be maintained, in order to allow the bridging phenomenon [46].
significantly be affected by the thickness of this protective layer The optimum fiber length (lfo) for composites reinforced with short
[49]. fibers was determined to be approximately 1.2 times the critical
The reinforcement efficiency of SFRCs depend on the fiber l/d fiber length lfc (lfo  1.2lfc) [46]. The optimal lfo for everX Posterior
and the Vf. If the fiber l/d is low, the fiber ends create discontinuity would be approximately 1.15 mm (1.0–1.3 mm). It follows then,
in the resin matrix, act as stress concentration sides and weaken that the short fiber lengths within the investigated SFRC should be
the composite [12]. Experimental versions of high aspect ratio within the range between the theoretical fiber length (0.85 mm)
SFRCs indicated that fibers with lengths from 2 mm to 5 mm and the optimal fiber length (1.15 mm). When reporting a value for
provide almost equal reinforcement capacity, and that optimal short fiber length, a range of values rather than a fixed value should
physicomechanical properties were obtained with formulations be used, because of the variations of the average short fiber values,
between 8.5 and 10 Vf%. Other studies have reported comparable which seemed to be consignment dependent. This is important,

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

8 J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx

because these variations could explain discrepancies in the middle range will provide pull–out toughening mechanism (0.4–
properties reported for the millimeter-scale SFRC. Earlier, a short 0.5 mm) and the longer ones (0.6 mm–0.7 mm) would actively
fiber length between 1.3 and 2.0 mm was reported for the everX carry the loading; and 4) small portion (6%) of the fibers exceeding
Posterior [20], whereas recently two different ranges of short fiber the optimal fiber length could break or then behave as continuous
length values were reported for the same material, namely 0.3– fibers. The fibers with varying length distributions affect the FOD
1.5 mm [47] and 1.0–2.0 mm [51]. The difference between the and contribute to the toughening effect.
present and those studies [20,47,51] might be due to the processing Clinically, a material that has high probability of stopping an
operations of the millimeter-scale SFRC regulated by the manu- initiated fracture would preserve the tooth structure, because it
facturer. Indeed, this type of SFRC is subjected to the mixing will absorb the dissipated energy released during cracking and
method of the fabrication process, which comprises blending stop the progression of the fracture toward the tooth structure. The
reinforcing fibers, fillers, resin matrix and additives together. This material itself would chip, delaminate or crack, but all these are
process generates fiber fractures and results in a wider range of favorable fracture types. On contrary, a material with low or no
millimeter-scale fibers, which likely altered the mechanical ability of stopping the progression of the initiated fracture will
properties in the final SFRC [19,20,47,51]. fracture. This is also accompanied by fracture of the tooth
The presence of shorter and longer short fibers in the structure. These phenomena are more visible for restorations
millimeter-scale SFRC system is important also for the FOD, fiber where most of the tooth structure is already lost and replaced with
debonding and consequently reinforcement efficiency. Indeed, the a restoration. In such situations, the selection of the material to
longer short fibers will tend to align in plane providing anisotropic replace the missing tooth tissue is of paramount importance and a
reinforcement, and shorter fiber will be randomly distributed material with high fracture toughness should be preferred.
offering an isotropic reinforcement. However, the debonding for Although the results of this investigation cannot be used to infer
random fibers is limited, which allows them to enhance the load how the fiber filler volume or fiber filler size affect the fracture
bearing capacity of the SFRC better than longer aligned short fibers toughness, the present study ascertains that the short E-glass
[15]. fibers affect significantly the fracture toughness. The fiber
Measuring the fiber orientation distribution is often difficult. A orientation distribution obviously warrants further research and
convenient method for characterizing the fiber orientation the fiber orientation in various dental cavities prepared on teeth,
distribution of short fiber composite has not yet been developed, approximal box and mesial-occlusal-distal (MOD) cavity could be
which is why the SEM microstructural examination was conducted studied next.
here (Fig. 1). Fiber orientation analysis supports the hypothesis of
the importance of the fiber orientation in enhancing the material’s
5. Conclusion
toughness. The crossing paths of transversally and normally
oriented fibers enabled radial alignment of the short fibers. This
In conclusion, the fibrous composite comprising millimeter-
type of orientation of the short fibers is known as X-orientation
scale discontinuous E-glass fibers fillers with a semi-IPN matrix
and is important for arresting the crack propagation at the crack
showed improved tolerance to crack propagation (fracture
notch tip (Fig. 2). However, it seems that the fiber orientation
toughness) compared to the tested conventional composites
distribution is not only dependent on the lf, df and Vf, but also on
containing particulate fillers. The toughening mechanisms provid-
the mold size and the placement technique. A pressing technique
ed by the short fibers were the reduction of the stress intensity at
would produce a planar fiber orientation, implying that the short
the crack tip, crack blunting and bridging phenomena. The
fibers would adjust against the surface they are pressed to and
properties of this millimeter-scale SFRC are dependent on groups
adapt individually according to its shape.
of factors: 1) Fiber strength and structure related factors: breaking
Failure types for the millimeter-scale SFRC involved matrix
of the embedded short fibers during manufacturing process, and
cracking, fiber fracture and fiber pull out. They all seemed to be
orientation and length of the short fibers; 2) Matrix related factors:
related to factors such as FOD and lf. Indeed, post-failure
viscosity affecting the short fibers’ orientation and distribution,
examination of both the fractured and tested single-edge notched
and porosity of the matrix such as air bubbles inclusions, pores or
bend specimens’ surfaces suggests ductile behavior of the fibers
voids and 3) Interfacial factors: silane coupling agent type of the
with failure occurrence in the matrix. This was evident at the
fibers and the particulate fillers, and reaction with matrix.
protruded fiber ends covered with resin (Fig. 5) as well as in the
bridging phenomenon (Fig. 4), as the fibers bridge the matrix crack.
The finding is supported by the load-deflection curve (Fig. 6), Conflicts of interest
suggesting that this would be possible only if the pullout resistance
of fibers at the first crack is greater than the load at the first The authors report no conflicts of interests. Author PK Vallittu
cracking. It seems that owing to the presence of short fibers, SFRC consults Stick Tech Ltd., a member of the GC Group in RD and
continues to carry tensile stresses after the first cracking [1]. training.
Contrary to this, the conventional PFCs fracture immediately at the
first peak, which indicates matrix cracking. Fiber breakage in this
Acknowledgements
study was also observed, probably at those fibers in which length
surpassed the optimal fiber length (>1.15 mm). This finding could
The authors thank Keme Scientific for the English language
be however advantageous when repairing SFRC restorations. In
editing of this manuscript. The authors also thank 3M ESPE and
addition, the SEM investigation revealed uniform fiber distribu-
Stick Tech Ltd. (member of GC group), for generously providing
tion, which is firstly important for avoiding localized domination of
materials used in this investigation. This study is part of the BioCity
matrix properties (at regions with less fibers) and secondly for
Turku Biomaterials Research Program (www.biomaterials.utu.fi).
avoiding physical contact between fibers [27].
Mechanical testing (Table 2), the fiber properties characteriza-
tion (Table 3), and the SEM analysis (Figs. 1, 2, 4 and 5) suggest that References
1) longer discontinuous fibers (30%) would enable bridging; 2)
[1] S.M. Lee, Handbook of Composite, Reinforcements, first ed., VCH Publishers,
short discontinuous fibers (18%) would behave as particulate fillers
Inc, New York, USA, 1992.
instead of fiber fillers; 3) some of the discontinuous fibers in the

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009
G Model
JJOD 2645 No. of Pages 9

J. Bijelic-Donova et al. / Journal of Dentistry xxx (2016) xxx–xxx 9

[2] P. Vallittu, High-aspect ratio fillers: fiber-reinforced composites and their [27] D. Hull, T.W. Clyne, An Introduction to Composite Materials, second ed.,
anisotropic properties, Dent. Mater. 31 (2015) 1–7. Cambridge University Press, Cambridge, 1996.
[3] J.L. Drummond, L. Lin, K.J. Miescke, Evaluation of fracture toughness of a fiber [28] M. Obukuro, Y. Takahashi, H. Shimizu, Effect of diameter of glass fibers on
containing dental composite after flexural fatigue, Dent. Mater. 20 (2004) 591– flexural properties of fiber-reinforced composites, Dent. Mater. J. 27 (2008)
599. 541–548.
[4] J. Manhart, K.H. Kunzelmann, H.Y. Chen, R. Hickel, Mechanical properties of [29] D.A. Norman, R.E. Robertson, The effect of fiber orientation on the toughening
new composite restorative materials, J. Biomed. Mater. Res. 53 (2000) 353– of short fiber-reinforced polymers, J. Appl. Polym. Sci. 90 (2003) 2740–2751.
361. [30] W.G. McDonough, J.M. Antonucci, J.P. Dunkers, Interfacial shear strengths of
[5] J.W. van Dijken, K. Sunnegårdh-Grönberg, Fiber-reinforced packable resin dental resin-glass fibers by the microbond test, Dent. Mater. 17 (2001) 492–
composites in class II cavities, J. Dent. 34 (2006) 763–769. 498.
[6] T.C. Fagundes, T.J. Barata, C.A. Carvalho, E.B. Franco, J.W. van Dijken, M.F.L. [31] J.L. Ferracane, H.X. Berge, J.R. Condon, In vitro aging of dental composites in
Navarro, Clinical evaluation of two packable posterior composites: a five-year water—effect of degree of conversion filler volume, and filler/matrix coupling,
follow-up, J. Am. Dent. Assoc. 140 (2009) 447–454. J. Biomed. Mater. Res. 42 (1998) 465–472.
[7] S.W. Yurgartis, Measurement of small angle fiber misalignments in continuous [32] J.L. Ferracane, V.A. Marker, Solvent degradation and reduced fracture
fiber composites, Compos. Sci. Technol. 30 (1987) 279–293. toughness in aged composites, J. Dent. Res. 71 (1992) 13–19.
[8] A.P. Wilczynski, A basic theory of reinforcement for unidirectional fibrous [33] I.S. Medeiros, M.N. Gomes, A.D. Loquercio, L.E. Filho, Diametral tensile strength
composites, Compos. Sci. Technol. 38 (1990) 327–337. and Vickers hardness of a composite after storage in different solutions, J. Oral
[9] Y.T. Zhu, W.R. Blumenthal, T.C. Lowe, Determination on non-symmetric 3-D Sci. 49 (2007) 61–66.
fiber-orientation distribution and average fiber length in short-fiber [34] L.V. Lassila, T. Nohrström, P.K. Vallittu, The influence of short-term water
composites, J. Compos. Mater. 31 (1997) 1287–1301. storage on the flexural properties of unidirectional glass fiber-reinforced
[10] S. Kari, H. Berger, U. Gabbert, Numerical evaluation of effective material composites, Biomaterials 23 (2002) 2221–2229.
properties of randomly distributed short cylindrical fibre composites, Comput. [35] P.K. Vallittu, Effect of 10 years of in vitro aging on the flexural properties of
Mater. Sci. 39 (2007) 198–204. fiber-reinforced resin composites, Int. J. Prosthodont. 20 (2007) 43–45.
[11] H. Berger, S. Kari, U. Gabbert, R. Rodriguez-Ramos, J. Bravo-Castillero, R. [36] C. Baudin, R. Osorio, M. Toledano, S. de Aza, Work of fracture of a composite
Guinovart-Diaz, Evaluation of effective material properties of randomly resin: fracture-toughening mechanisms, J. Biomed. Mater. Res. 89A (2009)
distributed short cylindrical fiber composites using a numerical 751–758.
homogenization tenchnique, J. Mech. Mater. Struct. 2 (2007) 1561–1570. [37] N. Ilie, R. Hickel, Investigations on mechanical behaviour of dental composites,
[12] P. Shouha, M. Swain, A. Ellakwa, The effect of fiber aspect ratio and volume Clin. Oral Invest. 13 (2009) 427–438.
loading on the flexural properties of flowable dental composite, Dent. Mater. [38] M.B. Shan, J.L. Ferracane, J.J. Kruzic, R-curve behavior and toughening
30 (2014) 1234–1244. mechanisms of resin-based dental composites: effects of hydration and
[13] A.L. Kalamkarov, H.Q. Liu, A new model for the multiphase fiber-matrix post-cure heat treatment, Dent. Mater. 25 (2009) 760–770.
composite materials, Compos. Part B Eng. 29B (1998) 643–653. [39] A.R. Curtis, A.C. Shortall, P.M. Marquis, W.M. Palin, Water uptake and strength
[14] R. Brighenti, D. Scorza, Numerical modelling of the fracture behaviour of brittle characteristics of a nanofilled resin-based composite, J. Dent. 36 (2008) 186–
materials reinforced with unidirectional or randomly distributed fibres, Mech. 193.
Mater. 52 (2012) 12–27. [40] A.R. Curtis, W.M. Palin, G.J. Fleming, A.C. Shortall, P.M. Marquis, The
[15] R. Brighenti, D. Scorza, A micro-mechanical model for statistically mechanical properties of nanofilled resin-based composites: characterizing
unidirectional and randomly distributed fibre-reinforced solids, Math. discrete filler particles and agglomerates using a micromanipulation
Mech. Solids 17 (2012) 876–893. technique, Dent. Mater. 25 (2009) 180–187.
[16] R.C. Petersen, Discontinuous fiber-reinforced composites above critical length, [41] N. Ilie, R. Hickel, Macro-, mico- and nano-mechanical investigations on
J. Dent. Res. 84 (2005) 365–370. silorane and methacrylate-based composites, Dent. Mater. 25 (2009) 810–819.
[17] L.K. Jain, R.C. Wetherhold, Effect of fiber orientation on the fracture toughness [42] R.Z. Alshali, N. Silikas, J.D. Satterthwaite, Degree of conversion of bulk-fill
of brittle matrix composites, Acta Metall. Mater. 40 (1992) 1135–1143. compared to conventional resin-composites at two time intervals, Dent. Mater.
[18] S. Garoushi, L.V. Lassila, P.K. Vallittu, The effect of span length of flexural 29 (2013) e213–217.
testing on properties of short fiber reinforced composite, J. Mater. Sci. Mater. [43] H.M. El-Damanhoury, J. Platt, Polymerization shrinkage stress kinetics and
Med. 23 (2012) 325–328. related properties of bulk-fill resin composites, Oper. Dent. 39 (2014) 374–382.
[19] J. Bijelic-Donova, S. Garoushi, P.K. Vallittu, L.V. Lassila, Mechanical properties [44] K.H. Kim, Y.B. Kim, O. Okuno, Microfracture mechanisms of composite resins
fracture resistance, and fatigue limits of short fiber reinforced dental containing prepolymerized particle fillers, Dent. Mater. J. 19 (2000) 22–33.
composite resin, J. Prosthet. Dent. 115 (2016) 95–102. [45] S.H. Kim, D.C. Watts, Effect of glass-fiber reinforcement and water storage on
[20] S. Garoushi, E. Säilynoja, P.K. Vallittu, L. Lassila, Physical properties and depth fracture toughness (KIC) of polymer-based provisional crown and FPD
of cure of a new short fiber reinforced composite, Dent. Mater. 29 (2013) 835– material, Int. J. Prosthodont. 17 (2004) 318–322.
841. [46] R.C. Wetherhold, L.K. Jain, The toughness of brittle matrix composites
[21] J.G. Leprince, W.M. Palin, J. Vanacker, J. Sabbagh, J. Devaux, G. Leloup, Physico- reinforced with discontinuous fibers, Mater. Sci. Eng. 151A (1992) 169–177.
mechanical characteristics of commercially available bulk-fill composites, J. [47] L. Lassila, S. Garoushi, P.K. Vallittu, E. Säilynoja, Mechanical properties of fiber
Dent. 42 (2014) 993–1000. reinforced restorative composite with two distinguished fiber length
[22] C. Goracci, M. Cadenaro, L. Fontanive, G. Giangrosso, J. Juloski, A. Vichi, M. distribution, J. Mech. Behav. Biomed. Mater. 60 (2016) 331–338.
Ferrari, Polymerization efficiency and flexural strength of low-stress [48] R.K. Nalla, J.H. Kinney, R.O. Ritchie, Effect of orientation on the in vitro fracture
restorative composites, Dent. Mater. 30 (2014) 688–694. toughness of dentin: the role of toughening mechanisms, Biomaterials 24
[23] K. Al-Ahdal, N. Ilie, N. Silikas, D.C. Watts, Polymerization kinetics and impact of (2003) 3955–3968.
post polymerization on the degree of conversion of bulk-fill resin-composite at [49] G.T. Rocca, C.M. Saratti, M. Cattani-Lorente, A.J. Feilzer, S. Scherrer, I. Krejci, The
clinically relevant depth, Dent. Mater. 31 (2015) 1207–1213. effect of a fiber reinforced cavity configuration on load bearing capacity and
[24] International Organization for Standardisation, ISO 4104. Dental zinc failure mode of endodontically treated molars restored with CAD/CAM resin
Polycarboxylate Cements, ISO, Geneva, 1984. composite overlay restorations, J. Dent. 43 (2015) 1106–1115.
[25] Americal Society for Testing, Materials, ASTM D4475-96, Standard Test [50] R.B. Fonseca, A.S. Marques, O. Bernades Kde, H.L. Carlo, L.Z. Naves, Effect of
Method for Apparent Horizontal Shear Strength of Pultruded Reinforced glass fiber incorporation on flexural properties of experimental composites,
Plastic Rods by the Short-beam Method, ASTM, Philadelphia, 2008 Available Biomed. Res. Int. (2014) 2014:542678.
at: http://www.astm.org/DATABASE.CART/HISTORICAL/D4475-96.htm. [51] H. Abouelleil, N. Pradelle, C. Villat, N. Attik, P. Colon, B. Grosgogeat, Comparison
[26] International Organization for Standardisation. ISO 20795-1. Dentistry—Base of mechanical properties of a new fiber reinforced composite and bulk filling
polymers—Part 1: Denture base polymers. ISO, Geneva, 2013. Available at: composites, Restor. Dent. Endod. 40 (2015) 262–270.
http://www.iso.org/iso/store.htm.

Please cite this article in press as: J. Bijelic-Donova, et al., Mechanical and structural characterization of discontinuous fiber-reinforced dental
resin composite, Journal of Dentistry (2016), http://dx.doi.org/10.1016/j.jdent.2016.07.009

You might also like