Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/299523119

Characterization of an experimental resin composite organic matrix based on


a tri-functional methacrylate monomer

Article  in  Dental Materials Journal · March 2016


DOI: 10.4012/dmj.2014-307

CITATIONS READS

14 308

6 authors, including:

Eduardo Moreira da Silva Jaime Dutra Noronha-Filho


Universidade Federal Fluminense Universidade Federal Fluminense
116 PUBLICATIONS   1,598 CITATIONS    9 PUBLICATIONS   69 CITATIONS   

SEE PROFILE SEE PROFILE

Cristiane Mariote Amaral Laiza Tatiana Poskus


Universidade Federal Fluminense Universidade Federal Fluminense
49 PUBLICATIONS   902 CITATIONS    39 PUBLICATIONS   815 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Dental Biomaterials Science - Research & Teaching View project

The influence of flaxseed (Linum usitatissimum) intake on Wistar rats body composition and bone quality View project

All content following this page was uploaded by Eduardo Moreira da Silva on 03 February 2017.

The user has requested enhancement of the downloaded file.


Dental Materials Journal 2016; 35(2): 159–165

Characterization of an experimental resin composite organic matrix based on a


tri-functional methacrylate monomer
Eduardo Moreira da SILVA, Luciana MIRAGAYA, Jaime Dutra NORONHA-FILHO, Cristiane Mariote AMARAL,
Laiza Tatiana POSKUS and José Guilherme Antunes GUIMARÃES

Analitical Laboratory of Restorative Biomaterials —LABiom-R, School of Dentistry, Federal Fluminense University, Niterói, Brazil, Rua Mário Santos
Braga, nº 30-Centro, Niterói, RJ, 24020-140, Brazil
Corresponding author, Eduardo Moreira da SILVA; E-mail: emsilva@vm.uff.br

This study evaluated the potential of a tri-functional monomer (trimethylolpropane trimethacrylate —TMPTMA) for inclusion in a
dental composite organic matrix. Initially, four ternary matrixes with different concentrations (wt%) of bi-functional monomers [Bis-
GMA (G), Bis-EMA (E) and TEGDMA (T)] were analyzed: GET523, GET532, EGT523 and EGT532 (the numbers (n) represent n×10
wt% of each monomer). The following properties were evaluated: degree of conversion, flexural strength, elastic modulus, hardness,
absorption, solubility, diffusion coefficient of water and crosslink density. Based on the best overall results obtained for EGT532,
all properties were re-evaluated in a matrix where TEGDMA (T) was replaced by a tri-functional monomer, TMPTMA (A)-EGA532.
EGA532 presented the best results for flexural strength, hardness, absorption and crosslink density. EGT523, EGT532 and EGA532
presented the lowest diffusion coefficients of water. The overall results indicated that TMPTMA could be useful in formulating
organic matrixes suitable for dental restorative composites.

Keywords: Resin composite, Organic matrix, TMPTMA, Chemical-mechanical properties

thereby increasing the filler incorporation and degree of


INTRODUCTION
conversion of the resin composite5).
Since the development of the resin composites by Rafael Other high molecular weight monomers were
BOWEN in the late 1950s, many efforts have been developed in an attempt to improve specific properties
made to improve the behavior of this class of restorative of commercially available resin composites10). Non-
materials1-3). In general, resin composites comprise three hydroxylated monomers, such as ethoxylated bisphenol
phases: a organic matrix formed by a mix of bi-functional A glycol dimethacrylate (Bis-EMA), are used to reduce
methacrylate monomers, inorganic filler particles the water absorption through the organic matrixes11).
and a silane-coupling agent that chemically bonds Moreover, the absence of -OH also makes this monomer
the organic matrix to the filler particles. These three less viscous than Bis-GMA.
phases strongly influence the materials’ properties4,5). The polymerization of resin composites involves
Irrespective of recent and important developments1,2), a free radical reaction by breaking aliphatic C=C
organic matrixes of resin composites are still composed bonds and establishing C-C covalent bonds between
of bi-functional methacrylate momomers6). 2,2-Bis[4- monomers, oligomers and adjacent polymer chains.
(2-hydroxy-3-methacrylyloxypropoxy)phenyl] propane Theoretically, a greater degree of C=C bond conversion
(Bis-GMA), the most used methacrylate monomer, is a will produce greater crosslinking in the rigid polymeric
high molecular weight molecule that has two aromatic matrix network. However, the literature shows that,
rings and two -OH groups in its backbone. On one even reaching a high degree of conversion, the polymeric
hand, these characteristics improve properties such network may present areas with different concentrations
as fracture toughness and elastic modulus7). However, of remaining C=C bonds12). This structural heterogeneity
these -OH groups may also establish inter- and intra- can lead to the reduction of crosslink density, increasing
molecular hydrogen bonding that strongly increases the plasticization of the polymeric network and
Bis-GMA viscosity5,8). This high viscosity negatively jeopardizing the chemical-mechanical properties of the
interferes with the degree of conversion9) and decreases resin composite13). In this field, some previous studies
the loading of high amounts of filler particles into the have shown that TEGDMA undergoes more primary
organic matrix. Moreover, these polar groups also cyclization reactions rather than effective crosslinking
increase the hydrophilicity of the Bis-GMA molecule, among polymeric chains14). This primary cyclization may
thereby increasing its water absorption. To overcome create areas of microgel formation, also increasing the
these shortcomings, low molecular weight monomers, polymer network heterogeneity.
such as triethyleneglycol dimethacrylate (TEGDMA), Trimethylolpropane trimethacrylate (TMPTMA) is
are incorporated to reduce the organic matrix viscosity, a low viscosity, tri-functional monomer that is able to
act as a crosslinking agent in polymeric matrixes. This
monomer was first used as isofiller, i.e., pre-polymerized
Color figures can be viewed in the online issue, which is avail-
able at J-STAGE.
Received Nov 7, 2014: Accepted May 19, 2015
doi:10.4012/dmj.2014-307 JOI JST.JSTAGE/dmj/2014-307
160 Dent Mater J 2016; 35(2): 159–165

filler, to improve wear resistance in experimental resolution of 4 cm−1. Afterwards, the increments were
microfilled composites15) and in experimental dental light-activated and the spectra were recorded exactly as
resins16). More recently, Ghasaban et al.17) showed that performed for the unpolymerized increments. The DC%
acting as a crosslinking agent, TMPTMA improved was calculated from the ratio between the integrated
the dynamic mechanical properties of cyanoacrylate area of ababsorption bands of the aliphatic C=C bond
adhesives suitable for bone fracture fixation. Although (1,638 cm−1) to the aromatic C=C bond (1,608 cm−1)
these studies showed promising results for this obtained from the polymerized and unpolymerized
monomer, there is still a lack of information about increments, using the following equation:
the use of TMPTMA for improving the chemical-
mechanical properties of organic matrixes suitable for DC%=100×[1−(Rpolymerized/Runpolymerized)],
(1)
producing resin composites. Therefore, the purpose of where R=area at 1,638 cm−1/area at 1,608 cm−1
the present investigation was to evaluate the effect of
replacing TEGDMA with the tri-functional monomer, Flexural strength and elastic modulus
TMPTMA, on the chemical-mechanical properties of an Bar-shaped specimens (n=10) were built up by filling
experimental resin composite organic matrix. The a steel split mold (10×2×1 mm) positioned over a
established hypothesis was that the replacement of polyester strip. After filling the mold to excess, the
TEGDMA by TMPTMA would improve the chemical- material surface was covered with a polyester strip and
mechanical properties of the experimental resin a glass slide and compressed with a device (500 g) for
composite organic matrix.

MATERIALS AND METHODS


Materials
The monomers Bis-GMA, Bis-EMA, TEGDMA and
TMPTMA (Fig. 1) were used as received (Essthec,
Essington, PA, USA). Initially, four ternary matrixes
based on TEGDMA were analyzed (Table 1). Afterwards,
a new matrix was produced, replacing TEGDMA
with TMPTMA in the formulation that presented the
best overall results in the first phase. In each matrix,
0.6 wt% of camphoroquinone and of 1.2 wt% of ethyl
N,N-dimethyl-4-aminobenzoate (EDMAB) (Aldrich
Chemical, Milwaukee, WI, USA) were introduced as
photosensitizers. The matrices were produced using a
dual centrifuge (SpeedMixer, Flaktek, Germany) with
three stages of 1 min at 1,500 rpm. All specimens in
this study were light-activated with a quartz-tungsten-
halogen unit (Optilux 501, Kerr, Danbury, CT, USA),
using an irradiance of 650 mW/cm2 for 30 s.

Degree of monomer conversion (DC%)


Increments of each matrix were inserted in a teflon
mold (0.785 mm3) that was positioned on an ATR
crystal of an FT-IR spectrometer (Alpha-P/Platinum
ATR Module, Bruker Optik, Ettlingen, Germany). The
spectra between 1,600 and 1,700 cm−1 were recorded
by the spectrometer operating with 40 scans and at a Fig. 1 Monomers used in this current study.

Table 1 Polymeric matrixes composition (monomers concentration-wt%)

Matrixes Bis-GMA (G) Bis-EMA (E) TEGDMA (T) TMPTMA (A)

GET523 50 20 30 0

GET532 50 30 20 0

EGT523 20 50 30 0

EGT532 30 50 20 0

EGA532 30 50 0 20
Dent Mater J 2016; 35(2): 159–165 161

20 s to avoid porosities. The specimens were then light was covered with a polyester strip and a glass slide,
activated from the top by two overlapping footprints. compressed with a device (500 g) for 20 s and light
After 24 h of dry storage at 37ºC, the specimens were activated (n=5). The discs were placed in a desiccator
submitted to a three-point bending test in a universal containing freshly dried silica gel, and transferred
testing machine, with an 8 mm span between the to an oven at 37ºC. After every 24 h period, the discs
supports, at a crosshead speed of 0.5 mm/min and were weighed (AUW 220, Shimadzu, Tokyo, Japan)
a 50 N load cell (DL 10000, Emic, Curitiba, Paraná, until a constant mass (m1) was attained, i.e., disc mass
Brazil). The elastic modulus (EM) was calculated from variation was less than ±0.01 mg in any 24 h period. The
the linear portion of the load/deflection curve and the thickness and the diameter of the discs were measured
flexural strength (FS) from the load at fracture, using using a digital caliper (MPI/E-101, Mitutoyo, Tokyo,
the following standard equations: Japan), and the volume (V ) was calculated in mm3.
The discs were then individually placed in plastic vials
l3F and immersed in 20 mL of distilled water at 37±1ºC.
EM= (MPa) (2)
4wh3d Subsequently, the discs were removed from the vials at
fixed time intervals, blotted dry with absorbent paper,
3lF weighed and returned to the same vials with 20 mL of
FS= (GPa), (3)
2wh2 distilled water. On the first day the discs were weighed
at time intervals of 5, 10, 15, 30, 60, 90, 120, 150 and
where l is the support span length (mm), F is the load 180 min. After the first day, the discs were weighed
(N) at fracture, w is the specimen width (mm), h is the daily until absorption equilibrium was reached [mass
specimen height and d is the deflection (mm) at load F. variation less than ±0.01 mg (m2)]. The discs were then
placed in a desiccator containing freshly dried silica gel,
Knoop hardness number (KHN) transferred to an oven at 37ºC and weighed daily until
Disc-shaped specimens (n=5) were built up by filling a the mass variation was less than ±0.01 mg (m3). The
Teflon split mold measuring 3 mm in diameter and 2 mm absorption (Sp) and the solubility (Sl) were calculated
in height. The mold was covered with a polyester strip using the following formulae (µg/mm3):
and a glass slide and the organic matrixes were light
activated. After 24 h of dry storage at 37ºC, the light- m2−m3
Sp= (5)
irradiated surfaces of the specimens were wet ground V
in a polishing machine (DPU-10, Struers, Copenhagen,
Denmark), using 1200 and 4000 grit SiC paper. Five m1−m3
Sl= (6)
Knoop indentations (50 g/15 s), spaced 500 µm apart, V
were made on the irradiated surfaces of each specimen
(Micromet 5104, Buehler, Lake Bluff, IL, USA). Diffusion coefficient of water (D)
For a solid bounded by two parallel planes, the Fick’s
Crosslink density (CLD) second law relates the diffusion coefficient of water
The crosslink density was estimated by evaluating the in one dimension (x) as a function of the time, and its
softening effect of 75% ethanol, according to the method solution is expressed as follows:
described by da Silva et al.12). Briefly, disc-shaped
specimens were built up by filling a Teflon split mold ∂C ∂2C
=D 2 (7)
measuring 3 mm in diameter and 2 mm in height (n=5). ∂t ∂x
After 24 h dry storage at 37°C, the irradiated surfaces
were wet polished with 1200 and 4000 grit SiC paper. Here, D is the diffusion coefficient of water and C is the
Afterwards, five Knoop indentations (50 g/15 s), spaced concentration (%) of the diffusing specimen at time t.
of 500 µm apart, were made on the polished surfaces of For longer times of diffusion, the solution to this
each specimen (Micromet 5104, Buehler). The specimens differential equation is expressed as follows:
were then stored in 75% ethanol for 24 h and the Knoop
hardness was reevaluated. The crosslink density (CLD) Mt 8 ∞ 1 (2n+1)2π2
=1− 2 Σ exp[− Dt] (8)
was estimated by using the difference between the mean M∞ π n=0 (2n+1)2 L2
values of Knoop diamond indentation depth (dk-μm)
after and before ethanol storage: Where Mt is the mass uptake (g) at time t, M ∞ is the
mass uptake (g) at equilibrium, and L is the specimen
CLD=dk final −dkinitial, (4) thickness (m).
However, for the initial stages of uptake (when
where a smaller difference indicates a higher CLD. Mt /M∞ ≤0.6), the above equation is reduced to:

Absorption, solubility and diffusion coefficient of water Mt 4 Dt 21


reduced to: = ( ) (9)
Disc-shaped specimens were built up by filling an M∞ L π
aluminum mold (6 mm in diameter and 1 mm in
thickness). After filling the mold to excess, the surface The diffusion coefficient of water D (10−13.m2.s−1)
162 Dent Mater J 2016; 35(2): 159–165

was obtained from the initial curve slope of the plot of Table 3 shows the results for Sp, Sl, D and CLD.
Mt /M ∞ against t1/2. With regard to absorption, the result was as follow: EGA
532<EGT532<EGT523<GET532<GET523, (p<0.05). The
Statistical analysis solubility values recorded for GET532, EGT523, EGT532
Statistical analysis was performed using Statgraphics and EGA532 were not statistically different from each
Centurion XVI software (Statpoint Technologies, other (p>0.05). The highest D was presented by GET532
VA, USA). The data for each variable was separately followed by GET523 (p<0.05). EGA532, EGT523 and
analyzed using one-way analysis of variance (ANOVA)
and Tukey’s HSD post hoc test. All analyses were
performed at a significance level of α=0.05.

RESULTS
The means and standard deviations for DC%, FS, FM
and KHN are presented in Table 2. No significant
difference was found for the DC% (p>0.05). On the
other hand, ANOVA revealed statistically significant
differences for FS, EM and KHN. The highest FS was
presented by EGA532 (p<0.05), followed by EGT532,
EGT523 and GET532, without differences among
them (p>0.05). The EM recorded for GET523, GET532,
EGT532 and EGA532 were not statistically different
(p>0.05). The lowest EM was presented by EGT523
(p<0.05), although this was not significantly different
from GET523, GET532 and EGT532. The highest KHN
was presented by EGA532 (p<0.05). The GET523,
GET532, EGT523 and EGT532 matrixes presented the Fig. 2 Plots of Mt /M∞ against the t1/2 for all polymeric
lowest and statistically similar KHN values (p>0.05). matrixes.

Table 2 Means (Standard Deviations) for degree of conversion, flexural strength, elastic modulus and Knoop hardness
number

Degree of conversion Flexural strength Elastic modulus Knoop hardness number


Matrices
(%) (MPa) (GPa) (HK)

GET523 58.8 (1.9)a 87.8 (5.2)a 2.0 (0.1)a,b 13.0 (1.0)a

GET532 55.8 (4.6)a 77.8 (2.9)b 1.8 (0.1)a,b 12.9 (0.8)a

EGT523 60.8 (3.2)a 78.2 (3.0)b 1.7 (0.2)a 13.4 (0.4)a

EGT532 58.2 (4.3)a 79 (2.0)b 1.8 (0.2)a,b 13.5 (0.3)a

EGA532 61.9 (2.3)a 102.8 (5.8)c 2.0 (0.1)b 15.8 (0.5)b

* For each property, means followed by different letters are significantly different (p<0.05).

Table 3 Means (Standard Deviations) for water absorption, solubility, diffusion coefficient and crosslink density

Water absorption Solubility Diffusion coefficient Crosslink density


Matrixes
(µg/mm3) (µg/mm3) (10−13×m2/s) (µm)

GET523 39.3 (1.2)a 10.6 (0.4)a 7.5 (0.4)a 2.9 (0.2)b

GET532 33.7 (1.6)b 9.9 (1.7)a,b 8.7 (0.4)b 2.5 (0.1)a

EGT523 28.8 (0.9)c 9.2 (0.5)a,b 6.0 (0.2)c 2.3 (0.1)a,c

EGT532 20.1 (1.1)d 8.7 (0.5)b 5.6 (0.4)c 2.1 (0.1)c

EGA532 17.6 (0.6)e 8.9 (0.6)b 5.5 (0.5)c 1.7 (0.3)d

* For each property, means followed by different letters are significantly different (p<0.05).
Dent Mater J 2016; 35(2): 159–165 163

EGT532 presented similar and lower D (p>0.05). The of methacrylate matrixes with initial low viscosity, the
best CLD was presented by EGA532 (p<0.05), followed onset of limiting the diffusion and auto-acceleration
by EGT532 and EGT523 with no statistically significant are delayed14). Moreover, due to the greater statistical
differences and by GET532 and GET523. likelihood of crosslink formation21), it is possible that,
The plots of Mt /M∞ against the t1/2 for all organic even during auto-deceleration (where the polymeric
matrixes are shown in Fig. 2. matrix reaches a high viscosity and consequently
a reduction in mobility), the physical structure of
DISCUSSION the TMPTMA monomer (Fig. 1) could favor the
establishment of crosslinking between close chains.
In the first phase of the present study, four ternary The synergism between these phenomena could have
matrixes were synthetized using bi-functional contributed to the highest CLD obtained with the
dimethacrylate monomers commonly used in EGA532 organic matrix. The increasing in TEGDMA
commercially available resin composites, so that the content from 20 to 30 wt% negatively affected the
concentration of TEGDMA in two of them was 20 or CLD (Table 3). This can be explained by the great
30 wt%. The concentrations of Bis-GMA and Bis-EMA tendency of TEGDMA to form primary cycles during the
were also maintained constant in two matrices (50 wt%) polymerization reaction.
and ranged in others (20 or 30 wt%). The rationale for In a clinical view, the flexural strength of a resin
this was to analyze the effect of each monomer on the composite should be high enough to support the stresses
chemical-mechanical behavior of the organic matrixes. derived from mastication, while its elastic modulus
In a second phase, all properties were re-evaluated in should be as close as possible to that of the surrounding
a new matrix produced by replacing TEGDMA with tissues, i.e., enamel and dentin, in order to avoid or
TMPTMA (EGA532) in the matrix that presented the diminish the stress transfer to the tooth-composite
best overall behavior in the first phase (EGT532). interface. In the present study, the EGA532 matrix
During the polymerization, organic matrixes of presented the highest flexural strength (Table 2). This
resin composites forms a tri-dimensional polymer can be explained by its better crosslink density. On the
network that strongly influences the chemical- other hand, its elastic modulus did not differ from those
mechanical properties of the material. The first aspect of GET523, GE532 and EGT532 (Table 2). Since highly
that dictates this network formation is the degree of crosslinked polymeric matrixes are brittle in nature,
conversion18), which is influenced by aspects such as thereby presenting a high elastic modulus, this result
monomer structure (bi- or poly-functionality), molecular was somewhat surprising. The findings of Viljanen et
weight and presence of polar (hydrogen-bonding) sites al.23) could explain this result. These authors showed
on monomer backbones19). In the present study, no that irrespective of the larger number of reactive C=C
differences were found in DC% among the matrixes bonds, a dendrimer/methylmethacrylate copolymer
(Table 2). Since the presence of stiff aromatic rings and presented low elastic modulus, and claimed that the
-OH pendant groups negatively affects the degree of plasticizing effect of pendant side chains influenced
conversion of Bis-GMA molecules20), it was a surprising this result. Thus, it is reasonable to assume that even
that the matrixes with 50 wt% of Bis-GMA (GET523 and presenting the highest CLD, the branched pendant
GET532) did not present a lower DC%. Thus, it is most chains of TMPTMA acted as a plasticizer, thereby
probable that the more flexible ether linkages and the avoiding an increase in the elastic modulus of EGA532.
high C=C bonds unit to volume ratio of Bis-EMA and The fact that GET523 and GET532 did not present
TEGDMA have overpassed the negative effect of Bis- a higher elastic modulus was also surprising because all
GMA on DC%20). In light of the former statements, it the matrixes presented a similar degree of conversion
is reasonable to assume that the chemical-mechanical (Table 2). Due to the presence of two -OH groups,
properties of the matrixes analyzed here were influenced Bis-GMA is more highly prone to establish strong
more by the structure of the monomers and less by the intermolecular interactions that decrease the flexibility
degree of conversion achieved after polymerization. of the polymer9). Theoretically, this should increase
On the other hand, the CLD was found to be the elastic modulus. One reasonable explanation for
significant (Table 3). This property plays a crucial role on this behavior could be based on a synergistic effect of
the chemical-mechanical behavior of resin composites primary cyclization of TEGDMA and the low cohesive
because it improves the packing density, reduces the energy of hydrogen bonding given by the -O- groups
free volume and decreases the polymer plasticization from Bis-EMA, which could have overshadowed the role
of the composite polymeric matrix12). These features of Bis-GMA on the elastic modulus of these matrixes.
may increase the values of flexural strength and elastic In dentistry, many studies have tried to correlate
modulus21), while decreasing the ability of solvents to the hardness with the wear of resin composites24-26).
dissolve into the polymeric network, thereby increasing However, due to the great diameter of the indenter used
its degradation resistance22). EGA532 presented the best (Knoop diamond), it is possible that these studies were
CLD (lower variation in Knoop indentation depths). not able to distinguish between the hardness of the
This result was probably influenced by the highest C=C polymeric matrixes to that of inorganic fillers, thereby
bonds unit to volume ratio and by the low viscosity of recording a great variability in the results that
the TMPTMA9,23). During the polymerization reaction could have impaired obtaining a correlation between
164 Dent Mater J 2016; 35(2): 159–165

composite hardness and wear. Differently, in the experimental organic matrix was accepted. It was
present study the hardness was measured exclusively in concluded that the tri-functional monomer TMPTMA
polymeric matrixes. Thus, the current results indicate has the potential to formulate organic matrixes to
a contribution of this phase to the overall hardness of a produce dental composites. Future studies analyzing
resin composite. new organic matrixes with different concentrations of
The hardness of EGA532 was the highest among TMPTMA must be performed.
the matrixes. This result is supported by the work of
Kawai et al.16) Those authors proved that experimental ACKNOWLEDGMENTS
binary methacrylate matrixes composed of TMPTMA/
TEGDMA were harder than others formed by Bis- This study was supported by FAPERJ (Grant
GMA, UDMA and TEGDMA. Moreover, they also found E-26/101.465/2010) and CNPq (Grant 474921/2009-3).
a negative correlation between hardness and wear of The authors thank ESSTHEC for donating all the
matrixes. According to those authors, their findings monomers.
were due to the 3-D polymer network produced by
TMPTMA. Thus, it is most probable that in the present REFERENCES
study the highly crosslinked structure of EGA532
positively influenced its hardness. 1) Cramer NB, Couch CL, Schreck KM, Carioscia JA, Boulden
The values of absorption in the present study JE, Stansbury JW, Bowman CN. Investigation of thiol-ene
and thiol-ene-methacrylate based resins as dental restorative
clearly show the role monomers play in this chemical materials. Dent Mater 2010; 26: 21-28.
property (Table 3). The two matrixes with 50 wt% of 2) Hsu SH, Chen RS, Chang YL, Chen MH, Cheng KC, Su WF.
Bis-GMA (GET523 and GET532) presented higher Biphenyl liquid crystalline epoxy resin as a low-shrinkage
absorption than those with 50 wt% of Bis-EMA resin-based dental restorative nanocomposite. Acta Biomater
(EGT523 and EGT532). These findings are supported 2012; 8: 4151-4161.
by previous studies and highlight the influence of polar, 3) Mankovskaia A, Levesque CM, Prakki A. Catechin-
incorporated dental copolymers inhibit growth of Streptococcus
hydrophilic -OH groups on absorption. Sideridou et al.11)
mutans. J Appl Oral Sci 2013; 21: 203-207.
showed that poly-Bis-GMA presented higher water 4) Goncalves L, Filho JD, Guimaraes JG, Poskus LT, Silva
absorption than did poly-Bis-EMA and claimed that EM. Solubility, salivary sorption and degree of conversion of
this was due to the higher cohesive energy density for dimethacrylate-based polymeric matrixes. J Biomed Mater
hydrogen bonding of -OH (2,980 J/cm3) when compared Res B Appl Biomater 2008; 85: 320-325.
to the -O- group present in Bis-EMA (881 J/cm3). 5) Ellakwa A, Cho N, Lee IB. The effect of resin matrix
composition on the polymerization shrinkage and rheological
Additionally, when analyzing three resin composites,
properties of experimental dental composites. Dent Mater
Örtengren et al.27) showed that the absorption of 2007; 23: 1229-1235.
composites with Bis-GMA and TEGDMA in their 6) Peutzfeldt A. Resin composite in dentistry: the monomer
matrixes was comparatively higher than a composite systems. Eur J Oral Sci 1997; 105: 97-116.
with Bis-EMA, i.e. an ethoxylated bisphenol A glycol 7) Ferracane JL, Greener EH. The effect of resin formulation on
dimethacrylate without hydrophilic -OH groups. The the degree of conversion and mechanical properties of dental
absorption recorded for EGA532 was the lowest. It is restorative resins. J Biomed Mater Res 1986; 20: 121-131.
8) Charton C, Falk V, Marchal P, Pla F, Colon P. Influence of Tg,
safe to assume that this behavior was influenced by viscosity and chemical structure of monomers on shrinkage
the better CLD achieved with this matrix. Another stress in light-cured dimethacrylate-based dental resins.
important aspect in this current study was the higher Dent Mater 2007; 23: 1447-1459.
D (the rate at which the water penetrated the polymer 9) Sideridou I, Tserki V, Papanastasiou G. Effect of
network) presented by the matrixes with a high content chemical structure on degree of conversion in light-cured
of Bis-GMA (Table 3/Fig. 2). dimethacrylate-based dental resins. Biomaterials 2002; 23:
1819-1829.
Different from absorption, the current results
10) Moszner N, Salz U. New developments of polymeric dental
did not show a remarkable influence of monomers on composites. Prog Polym Sci 2001; 26: 535-576.
solubility (Table 3). Most probably, this behavior was 11) Sideridou I, Tserki V, Papanastasiou G. Study of water
influenced by the similarity in DC% among matrixes sorption, solubility and modulus of elasticity of light-cured
(Table 2). In fact, previous studies have already dimethacrylate-based dental resins. Biomaterials 2003; 24:
shown that solubility correlates well with degree of 655-665.
12) da Silva EM, Poskus LT, Guimaraes JG, de Araujo Lima
conversion4, 28).
Barcellos A, Fellows CE. Influence of light polymerization
modes on degree of conversion and crosslink density of dental
CONCLUSION composites. J Mater Sci Mater Med 2008; 19: 1027-1032.
13) Filho JD, Poskus LT, Guimaraes JG, Barcellos AA, Silva EM.
The EGA523 matrix presented better behavior Degree of conversion and plasticization of dimethacrylate-
regarding flexural strength, hardness, absorption based polymeric matrices: influence of light-curing mode. J
and crosslink density. Moreover, this matrix behaved Oral Sci 2008; 50: 315-321.
14) Dickens SH, Stansbury JW, Choi KM, Floyd CJE.
similarly to the other matrices for the remaining
Photopolymerization kinetics of methacrylate dental resins.
analyzed properties. Thus, the formulated hypothesis Macromolecules 2003; 36: 6043-6053.
that the replacement of TEGDMA by TMPTMA would 15) Suzuki S, Leinfelder KF. An in vitro evaluation of a
improve the chemical-mechanical properties of the copolymerizable type of microfilled composite resin.
Dent Mater J 2016; 35(2): 159–165 165

Quintessence Int 1994; 25: 59-64. 22.


16) Kawai K, Iwami Y, Ebisu S. Effect of resin monomer 22) Ferracane JL. Hygroscopic and hydrolytic effects in dental
composition on toothbrush wear resistance. J Oral Rehabil polymer networks. Dent Mater 2006; 22: 211-222.
1998; 25: 264-268. 23) Viljanen EK, Lassila LV, Skrifvars M, Vallittu PK. Degree
17) Ghasaban S, Atai M, Imani M, Zandi M, Shokrgozar MA. of conversion and flexural properties of a dendrimer/methyl
Photo-crosslinkable cyanoacrylate bioadhesive: shrinkage methacrylate copolymer: design of experiments and statistical
kinetics, dynamic mechanical properties, and biocompatibility screening. Dent Mater 2005; 21: 172-177.
of adhesives containing TMPTMA and POSS nanostructures 24) Yap AU, Ong LF, Teoh SH, Hastings GW. Comparative
as crosslinking agents. J Biomed Mater Res A 2011; 99: 240- wear ranking of dental restoratives with the BIOMAT wear
248. simulator. Part II. An SEM evaluation. J Oral Rehabil 2000;
18) da Silva EM, Poskus LT, Guimaraes JG. Influence of light- 27: 52-59.
polymerization modes on the degree of conversion and 25) Harrison A, Draughn RA. Abrasive wear, tensile strength, and
mechanical properties of resin composites: a comparative hardness of dental composite resins —is there a relationship?
analysis between a hybrid and a nanofilled composite. Oper J Prosthet Dent 1976; 36: 395-398.
Dent 2008; 33: 287-293. 26) Bayne SC. Correlation of clinical performance with ‘in vitro
19) Leprince JG, Palin WM, Hadis MA, Devaux J, Leloup G. tests’ of restorative dental materials that use polymer-based
Progress in dimethacrylate-based dental composite technology matrices. Dent Mater 2012; 28: 52-71.
and curing efficiency. Dent Mater 2013; 29: 139-156. 27) Örtengren U, Andersson F, Elgh U, Terselius B, Karlsson S.
20) Pfeifer CS, Silva LR, Kawano Y, Braga RR. Bis-GMA co- Influence of pH and storage time on the sorption and solubility
polymerizations: influence on conversion, flexural properties, behaviour of three composite resin materials. J Dent 2001;
fracture toughness and susceptibility to ethanol degradation 29: 35-41.
of experimental composites. Dent Mater 2009; 25: 1136- 28) da Silva EM, Almeida GS, Poskus LT, Guimaraes JG.
1141. Relationship between the degree of conversion, solubility and
21) Stansbury JW. Dimethacrylate network formation and salivary sorption of a hybrid and a nanofilled resin composite.
polymer property evolution as determined by the selection of J Appl Oral Sci 2008; 16: 161-166.
monomers and curing conditions. Dent Mater 2012; 28: 13-

View publication stats

You might also like