Anda di halaman 1dari 14

This article was downloaded by: [University of Auckland Library]

On: 26 October 2014, At: 15:12


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Journal of Biomaterials Science,


Polymer Edition
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tbsp20

Dynamic Mechanical Analysis of


a Hybrid and a Nanohybrid Light-
Cured Dental Resin Composite
a b
Irini D. Sideridou , Maria M. Karabela & Crysa S.
c
Spyroudi
a
Laboratory of Organic Chemical Technology,
Department of Chemistry, Aristotle University of
Thessaloniki, Thessaloniki GR-54124, Greece
b
Laboratory of Organic Chemical Technology,
Department of Chemistry, Aristotle University of
Thessaloniki, Thessaloniki GR-54124, Greece
c
Laboratory of Organic Chemical Technology,
Department of Chemistry, Aristotle University of
Thessaloniki, Thessaloniki GR-54124, Greece
Published online: 02 Apr 2012.

To cite this article: Irini D. Sideridou , Maria M. Karabela & Crysa S. Spyroudi (2009)
Dynamic Mechanical Analysis of a Hybrid and a Nanohybrid Light-Cured Dental Resin
Composite, Journal of Biomaterials Science, Polymer Edition, 20:12, 1797-1808, DOI:
10.1163/156856208X386408

To link to this article: http://dx.doi.org/10.1163/156856208X386408

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information
(the Content) contained in the publications on our platform. However, Taylor
& Francis, our agents, and our licensors make no representations or warranties
whatsoever as to the accuracy, completeness, or suitability for any purpose
of the Content. Any opinions and views expressed in this publication are the
opinions and views of the authors, and are not the views of or endorsed by
Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and
Francis shall not be liable for any losses, actions, claims, proceedings, demands,
costs, expenses, damages, and other liabilities whatsoever or howsoever caused
arising directly or indirectly in connection with, in relation to or arising out of the
use of the Content.

This article may be used for research, teaching, and private study purposes.
Any substantial or systematic reproduction, redistribution, reselling, loan, sub-
licensing, systematic supply, or distribution in any form to anyone is expressly
forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Downloaded by [University of Auckland Library] at 15:12 26 October 2014
Journal of Biomaterials Science 20 (2009) 17971808
www.brill.nl/jbs

Dynamic Mechanical Analysis of a Hybrid and a


Nanohybrid Light-Cured Dental Resin Composite

Irini D. Sideridou , Maria M. Karabela and Crysa S. Spyroudi


Laboratory of Organic Chemical Technology, Department of Chemistry,
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

Aristotle University of Thessaloniki, Thessaloniki GR-54124, Greece


Received 3 June 2008; accepted 17 October 2008

Abstract
This work was aimed at the study of the viscoelastic properties of two current light-cured dental resin
composites, Rok (hybrid) and Ice (nanohybrid), by dynamic mechanical analysis, under the influence of
variable temperatures and frequencies 1 h after curing or after storage at 37 C either in air or distilled water
for 1, 7 or 30 days. They both contain about the same amount of filler, which is strontium aluminosilicate
particles, but they have different size distribution, 40 nm2.5 m for Rok and 10 nm1 m for Ice. The resin
matrix of Rok consists of UDMA, while that of Ice consists of UDMA, Bis-EMA and TEGDMA. During
dynamic testing elastic modulus (E  ), viscous modulus (E  ) and loss tangent (tan ) were determined over
a temperature range from 25 to 185 C at a frequency of 0.1, 1.0 or 10.0 Hz. The elastic modulus (8.40
0.61 GPa) and tan (0.114) of Rok at 37 C and 1 Hz was slightly higher than that of Ice (E  = 7.58
0.50 GPa and tan = 0.110) which, however, were not statistically different (P  0.05). Both the light-
cured composites showed two Tg values, at 58 and 121 C for Rok and at 63 and 117.0 C for Ice. The first
Tg has an apparent value, since inside the glass transition region an additional curing reaction occurred of
the unreacted C=C methacrylate groups; the second Tg is due to the new material formed during the DMA
scan. During the storage of composites in air or water also an additional curing reaction took place. The
nanohybrid Ice showed a slight lower elastic modulus and tan than the hybrid Rok at 2550 C, which,
however, were not statistically different (P  0.05).
Koninklijke Brill NV, Leiden, 2009

Keywords
Dental resin composites, dynamic mechanical properties, Rok, Ice

1. Introduction

Dental composites are complex, tooth-colored filler materials composed of syn-


thetic polymers, particulate ceramic reinforcing fillers and silane coupling agents,

* To whom correspondence should be addressed. Fax: (30-210) 2310-997769;


e-mail: siderid@chem.auth.gr

Koninklijke Brill NV, Leiden, 2009 DOI:10.1163/156856208X386408


1798 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808

which bond the fillers to the polymer matrix. They also contain molecules which
promote or modify the polymerization reaction that produces the cross-linked poly-
mer matrix from the dimethacrylate monomers. Each component of the composite
is critical to the success of the final dental restoration. However, the most sig-
nificant developments in the evolution commercial composites to date have been
a direct result of modifications to the filler component [1]. These changes have
prompted the periodic development of classification systems for dental compos-
ites based upon filler size and volume fraction. Therefore, dental composites have
been classified in 1983 according to the particle size of the filler [2]. Although
this classification is 24 years old it is still valid for modern type dental composites
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

[1, 3]. This system includes the macrofilled or traditional composites (avg. parti-
cle size = 115 m), hybrids containing a mixture of macrofiller and microfiller
particles (avg. particle sizes = 5 m and 0.04 m, respectively) and microfilled
(avg. particle size = 0.040.1 m). The only update to this classification is that
macrofilled composites are significantly less frequently used nowdays, because of
esthetic concerns. Macrofiller particles are purely inorganic, usually splinter-shaped
and today have an average particle size between 0.2 and 5.0 m. Microfillers consist
of spherical primary particles with an average particle size of approx. 5100 nm.
Therefore, the term nanofiller is also used for such particles [3, 4].
The mechanical properties of dental composites traditionally have been investi-
gated using static methods. However, these focus only on the elastic component of
the composite. Dental composites exhibit viscoelastic behavior due to the presence
of polymer matrix, which is a viscoelastic material. Dynamic tests such as dynamic
mechanical analysis (DMA), on the other hand, are particularly well suited for vis-
coelastic materials, since they determine both the elastic and viscous response of
material. Also dynamic tests mimic better the cyclic masticatory loading to which
dental composites are clinically subjected [5]. This might be extremely valuable
to predict the clinical performance of biomaterials when working under the cyclic
solicitations generated by the human bodys physiological movements. The vis-
coelastic properties of several contemporary dental resins have been studied using
DMA [612].
This work was aimed at the DMA analysis of two commercial light-cured dental
composites, one of which is characterized as hybrid (Rok) and the other as nanohy-
brid (Ice). They both contain about the same amount of filler, which is strontium
aluminosilicate particles, but they have different size distribution, 40 nm2.5 m
for Rok and 10 nm1 m for Ice. Also they have a different resin matrix; in Rok
the resin matrix consists only of UDMA units, while in Ice the resin matrix con-
sists of UDMA, Bis-EMA and TEGDMA monomer units. This work aimed also at
the study of the change of viscoelastic properties of composites after aging in air
or water for 1, 7 or 30 days. The wet condition was used to simulate the moisture
in the oral environment and the dry condition was taken for comparison to reveal
changes in the properties due to water sorption.
I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808 1799

2. Materials and Methods

2.1. Materials

Rok (hybrid) and the Ice (nanohybrid) are light-curable dental resin composites pre-
pared by SDI (Melbourne, Vic, Australia). Their specifications are listed in Table 1.

2.2. Preparation of Specimens

Bar specimens were prepared by filling a Teflon mold (2 mm 2 mm 40 mm)


with unpolymerized material, taking care to minimize entrapped air. The upper and
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

lower surface of the mold was overlaid with glass slides covered with a Mylar sheet
to avoid adhesion with the unpolymerized material. The completed assembly was
held together with spring clips and irradiated using a XL 3000 dental photocuring
unit (3M-ESPE, Seefeld, Germany). This unit was used without the light guide on
contact with the glass slides. The samples were cured by a series of overlapping
light exposures of upper and lower surfaces of 60 s duration on each side. Then,
the mould was dismantled and the composite was carefully removed by flexing the
Teflon mould. A total of 21 bar-shaped specimens were prepared for each com-
posite. These samples were equally distributed into seven groups. One group was
tested immediately after preparation; three other groups were conditioned in air at

Table 1.
Data for composites studied in this work

Composite Composite type Batch No./shade Resin Filler type and loading
compo-
nents

Rok Hybrid LOT 030544/B2 UDMA Strontium aluminosilicate


(Sr3 Al10 SiO20 )
(7782 wt%, 63 vol%,
40 nm2.5 m)
Posterior
Core built up
Splinting
Ice Nanohybrid LOT 0505134/A1 UDMA, Strontium aluminosilicate
Bis-EMA, (Sr3 Al10 SiO20 )
TEGDMA (77 wt%, 61 vol%,
10 nm1 m)
Anterior/posterior
Veneers
Inlays/Onlays
Core built up

Information provided by the manufacturer.


1800 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808

37 1 C for either 1, 7 or 30 days, and the last three groups were stored in distilled
water at 37 1 C, also either for 1, 7 or 30 days. After the storage period, each
specimen was tested in a dynamic mechanical analyzer.
2.3. Dynamic Mechanical Analysis (DMA)
DMA tests were performed on a Diamond DMA dynamic mechanical analyzer
(Perkin-Elmer, Waltham, MA, USA) in bending dual cantilever clamped mode.
In DMA the viscoelastic properties of a material are characterized by applying a
sinusoidal deformation to the material at a single or at multiple frequencies and
monitoring the response of the material. Since polymers are viscoelastic materi-
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

als, i.e., they exhibit solid-like and liquid-like properties, they are by definition
time-dependent. This means that the response of a viscoelastic material to an im-
posed deformation will depend on how fast or slow the deformation was applied
to the sample. When characterizing a material by DMA, the time of the defor-
mation is the frequency (ies), as frequency is the inverse of time. In our exper-
iments a frequency of 0.1, 1.0 or 10 Hz was applied and amplitude of 10 m.
These frequencies were chosen because they represented a range of strain rates
from close to static testing (0.1 Hz) to the upper limit of normal chewing fre-
quency (10 Hz) [13]; the frequency of 1 Hz is the approximately average mas-
ticatory chewing rate. A heating rate of 2 C/min and a temperature range of
25185 C were selected to cover mouth temperature and the materials likely
glass-transition temperature (Tg ). Elastic modulus (E  ), viscous modulus (E  ) and
loss tangent (tan ) were plotted against temperature over this range. This method
was used for each of the samples and the mean values calculated. For the group
with dry samples, after the DMA run was complete, the sample was allowed
to cool at room temperature and then the process was repeated with the speci-
men.
2.4. Statistical Analysis
The values reported in the tables and figures represent mean values standard de-
viation of replicates. One-way analysis of variance (ANOVA) test, followed by
Tukeys test, for multiple comparisons between means to determine significant
differences was used at a significant level set at P  0.05. This was performed
separately for each of one of the independent variables: material, temperature, fre-
quency and storage time.

3. Results
Figure 1 displays the elastic modulus, viscous modulus and loss tangent versus
temperature curves obtained for light-cured Rok at a constant frequency of 1 Hz.
Analogous curves were obtained for Ice. After the first scan the specimen was
cooled down to 25 C and a second scan was recorded from 25 to 185 C. The elastic
modulus, viscous modulus and loss tangent versus temperature curves obtained for
I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808 1801
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

Figure 1. Plot of elastic modulus (E  ), viscous modulus (E  ) and loss tangent (tan ) as a function of
temperature at 1 Hz during the first and the second run in DMA for Rok (above) and Ice (below). The
composites were tested immediately after light-curing.

both composites at a constant frequency of 1 Hz during the first and the second run
in DMA are shown in Fig. 1.
In Fig. 2 is shown the dependence of frequency on the viscoelastic properties of
composites. The elastic modulus consistently increased with increases in frequency
(Fig. 2a). For the same temperature the storage modulus is higher at higher fre-
quency. The viscous modulus seems not to be affected by the frequency up to about
110 C and then increased with increases in frequency (Fig. 2b). In Fig. 2c is shown
the dependence of the loss tangent on the frequency. The first peak is observed at the
same temperature but with a smaller height at higher frequency, while the second
peak shifts at a higher temperature at a higher frequency.
In Table 2 are shown the Tg values of composites during the first and second run
in DMA at three frequencies. In Table 3 are compared the values of elastic modulus,
viscous modulus and loss tangent of composites at a range of temperatures and at a
constant frequency of 1.0 Hz.
The effect of composites storage in air or water for 1, 7 or 30 days on the vis-
coelastic properties is shown for Rok in Fig. 3 and for Ice in Fig. 4. The Tg of
composites after storage in air or water is shown in Table 4.
1802 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

Figure 2. Plot of elastic modulus (E  ) (a), viscous modulus (E  ) (b) and loss tangent (tan ) (c) as a
function of temperature at three frequencies for Rok (above) and Ice (below). The composites were
tested immediately after light-curing.

Table 2.
Glass-transition temperatures (Tg ) of composites during the first and second run in DMA at three
frequencies (mean (SD), n = 4)

Composite Frequency First run Second run


(Hz)
Tg1 ( C) Tg2 ( C) Tg1 ( C) Tg2 ( C)

Rok 0.1 58.7 (2.2) a,A 109.5 (2.5)D 121.9 (2.3)


1.0 57.9 (2.5) a,B 121.4 (1.2) 129.5 (2.1)
10 60.0 (1.6) a,C 130.7 (0.0) 139.7 (4.6)
Ice 0.1 63.3 (6.7)b,A 105.8 (2.3)D 113.0 (0.1)
1.0 62.6 (7.5)b,B 117.0 (1.3) 122.4 (2.8)
10 64.5 (6.0)b,C 126.1 (0.4) 131.8 (1.7)

Common corresponding letters indicate no significant difference (P  0.05).

4. Discussion
The log E  vs. temperature curve for both composites showed two abrupt decreases,
one at low temperatures (region I) and the other at high temperatures (region III)
I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808 1803

Table 3.
Elastic modulus (E  ), viscous modulus (E  ) and loss tangent (tan ) of composites at a range of
temperatures at constant frequency 1.0 Hz (mean (SD), n = 4)

Tempera- E  (GPa) E  (GPa) tan (103 )


ture ( C)
Rok Ice Rok Ice Rok Ice

25 9.35 (0.44)a,A 8.87 (0.27)a 0.94 (0.03)I 0.77 (0.05)K 102 (7)d,L 88 (10)d,N
37 8.40 (0.61)b,A,B 7.58 (0.50)b 0.94 (0.02)I 0.80 (0.05)K 114 (7)e,L 110 (14)e,N
50 7.21 (0.78)c,B,C 6.54 (0.26)c 0.94 (0.04)I 0.84 (0.04)K 133 (12)f,M 130 (5)f
70 6.27 (0.63)C,D 5.22 (0.28) 0.85 (0.05)I 0.75 (0.03)K 136 (7)g,M 143 (2)g,O
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

90 5.71 (0.58)D,E 4.60 (0.18) 0.72 (0.06)J 0.64 (0.03) 126 (2)M 140 (3),O
110 4.68 (0.53)E 3.29 (0.21) 0.61 (0.06)J 0.53 (0.02) 131 (4)M 163 (4)
130 3.50 (0.40)F 2.11 (0.16) 0.43 (0.05) 0.32 (0.02) 126 (12)M 146 (3)O
150 2.82 (0.30)F,G 1.64 (0.11)H 0.23 (0.04) 0.13 (0.02) 81 (5)h 81 (3)h
170 2.60 (0.23)G 1.54 (0.09)H 0.14 (0.02) 0.08 (0.01) 55 (3)i 51 (1)i

Common corresponding lowercase letters in a given row indicate no significant difference (P 


0.05). Common corresponding uppercase letters in a given column indicate no significant difference
(P  0.05).

Figure 3. Rok: Plot of elastic modulus (E  ), viscous modulus (E  ) and loss tangent (tan ) as a
function of temperature at 1 Hz, after aging in air (above) or water (below) for 1, 7 or 30 days.
1804 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

Figure 4. Ice: Plot of elastic modulus (E  ), viscous modulus (E  ) and loss tangent (tan ) as a function
of temperature at 1 Hz, after aging in air (above) or water (below) for 1, 7 or 30 days.

Table 4.
Glass-transition temperatures (Tg ) of composites after storage in air or water for 1, 7 or 30 days;
frequency = 1.0 Hz (mean (SD), n = 4)

Composite Storage time Tg in air Tg in water


(days) ( C) ( C)

Rok 0 121.4 (1.2)


1 116.2 (1.0) 105.2 (3.0) B
7 111.0 (2.7) 96.1 (1.9) C
30 102.4 (3.5)A 89.1 (1.9) D
Ice 0 117.0 (1.3)
1 108.0 (4.4) 108.4 (2.9) B
7 99.8 (0.3)a 94.5 (1.9) C
30 99.2 (0.9)a,A 90.9 (0.5) D

Common corresponding letters indicate no significant difference (P  0.05).

(Fig. 1a). The abrupt decrease in the elastic modulus as a function of temperature is
attributed to the glass transition region of the material. Region I for Rok is between
25 and 61 C, and for Ice between 25 and 58 C. Region III for Rok is 96148 C
I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808 1805

and for Ice 92144 C. Region II, between 6196 C for Rok and 5892 C for Ice,
is due to the curing reaction occurring during the test, because of the temperature
increase. The dashed curve in Fig. 1a was drawn assuming that there was no curing
reaction during the test. If this is the case, the Tg of the light-cured composite could
be defined with accuracy from the corresponding tan peak. However, the Tg value
of 58 C for Rok and 63 C for Ice, obtained from the tan peak at low temperature
(Fig. 1c), must underestimate the actual Tg of the light-cured sample, because of the
curing reaction occurring inside the glass transition region. Therefore, the determi-
nation of the Tg using the DMA as the peak of tan is not perfectly applicable to the
light-cured samples at room temperature. Above approx. 145 C the elastic modu-
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

lus of both composites stopped decreasing, forming the rubbery plateau region IV
(Fig. 1a). In this region the polymer network is completely softened and the storage
modulus is restrained from falling by the stable cross-linked network.
The curve of viscous modulus E  for both composites showed two different
decreasing regimes: the first between 60 and 100 C with a smaller slope and the
second between 100175 C with a greater slope (Fig. 1b).
Each of the two maximums in the tan versus temperature curves corresponds
to one of the two drops in the storage modulus curves and represent two Tg values.
The first Tg is the apparent one of the light-cured specimen, which still contains
unreacted methacrylate C=C bonds within the vitrified network, whereas the sec-
ond Tg is due to the specimen-cured during the DMA test. Two Tg values were also
observed in the tan curve of UDMA/TEGDMA resin [14], a Bis-GMA resin [15]
and in two commercial light-cured resin composites [8, 16]; in all these cases the
two Tg values were attributed to the additional thermal cure occurring in the middle
of the first glass transition.
In Fig. 1 are also shown the curves obtained during the second run. The dou-
ble decrease of elastic modulus became a single decrease and the values of elastic
modulus were higher than the corresponding ones recorded during the first run. The
values of viscous modulus in the second run remained almost the same. The loss
tangent peak at low temperature was disappeared, while that at higher temperature
remained also the same. This behavior confirmed the fact that additional curing
reaction took place during the first run.
The elastic modulus of both composites sloped upward towards higher frequency
(Fig. 2a) what is typical for a viscoelastic material [17, 18]. Since these materials
simultaneously exhibit solid-like and liquid-like properties, they are by definition
time-dependent. This means that the response of a viscoelastic material to an im-
posed deformation will depend on how fast or slow the deformation was applied to
the sample. When characterizing a material by DMA the time of the deformation is
defined by the frequency, as frequency is the inverse of time. Therefore, high fre-
quencies are analogous to short times and low frequencies to long times. The low
frequency range is where the viscous-like behavior predominates, since the material
will have enough time to flow. As frequency rises the material act in a more elastic
way, thus increasing the modulus.
1806 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808

The elastic modulus of Rok at the usual mouth temperature (37 C) was found
to be 7.34 0.47 GPa, 8.40 0.61 GPa or 9.51 0.71 GPa at frequency 0.1 Hz,
1.0 Hz or 10 Hz, correspondingly. The storage modulus of Ice at 37 C is slightly
lower, 6.69 0.53 GPa, 7.58 0.50 GPa or 8.60 0.41 GPa, at frequency 0.1 Hz,
1.0 Hz or 10 Hz. These values for elastic modulus are much lower than the values of
dynamic modulus of dentine at 37 C, which are 15.16 GPa (0.1 Hz) 15.54 (1.0 Hz)
or 15.85 (10 Hz) [19].
The increase of frequency reduced the loss tangent for both composites, due
to the corresponding increase of storage modulus E  (tan = E  /E  ). However,
at higher temperatures after the second peak of tan , the value of tan is higher
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

at higher frequency, because of the increase of the viscous modulus E  with the
increase of the frequency. For Rok the first tan peak occurred at about 59 C and
is not changed with the frequency, which indicates that it is not a true relaxation
(Tg ) peak. The second tan peak increases from 110 C to 131 C by increasing
the frequency from 0.1 to 10 Hz. This peak is shifted in the second run at higher
temperatures, 122 C and 140 C, correspondingly. The second tan peak increases
from 110 C to 131 C by increasing the frequency from 0.1 to 10 Hz (Table 2),
confirming that additional curing reaction occurred during the first run. It is known
that the higher the cross-linking density of the polymer network, the higher the Tg of
the network. For Ice the first tan appeared at about 63 C which is not statistically
different from that of Rok. The second tan peak increases from 106 C to 126 C
by increasing the frequency from 0.1 to 10 Hz. Also this peak shifted to higher
temperatures, 113 C and 132 C, during the second run. The values of Tg2 for Ice
are lower than the corresponding ones determined for Rok. This most probably is
due to the different composition of the polymer matrix. In Rok the polymer matrix
consists of UDMA monomer units, while in Ice it consists of UDMA, Bis-EMA
and TEGDMA monomer units and it seems to be slightly more flexible than that of
Rok.
The elastic modulus of Rok at 1.0 Hz in a range of temperatures was always
higher than the corresponding values obtained for Ice (Table 3). However, these
values were statistically different (P  0.05) at about 70 C and higher. The same
trend was also observed for the viscous modulus of the two composites all over the
temperature range. The tan of Rok was not statistically different than that of Ice
for the region 2570 C, while it was lower for higher temperatures, up to 130 C.
In the literature some theories indicate that the elastic modulus of a material should
be independent of the size of the filler [20]. In a recent publication it is reported
that, there was no significant difference in elastic modulus among composites with
various filler sizes [21]. However, other authors have reported an increase in mod-
ulus with a decrease in particle size [22]. Various theories have been suggested,
including that as the particle size decreases, the increase in surface area and surface
energy provides more efficient interfacial bonds [22]. In this study the different size
distribution of the filler particles of the two composites have no significant effect
on the elastic modulus at 2550 C.
I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808 1807

After storage of the composites for 7 or 30 days in air or water the two drops
observed in the elastic modulus curve of the light-cured samples converged to a
single drop and the region of thermal reaction inside the glass transition region dis-
appeared (Figs 3 and 4). This result shows that during the storage additional curing
reaction took place. This means that by increasing the temperature from room tem-
perature to 37 C, it is clear that we can observe a slight increase in the degree of
conversion. On the other hand, the ingestion of hot food and beverages would only
have limited impact on the degree of conversion since resin composites are not ther-
mal conductors and it is known that an increase of only 5.5 C of the tooth can lead
to irreversible adverse effects on the pulp tissues (pulpitis and secondary necrosis).
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

By the way, an increase of the degree of the degree of conversion is more probably
observed only at the surface of the restoration than in the depth of the material.
The viscous modulus curves of composites stored in air or water for 7 or 30 days
showed an abrupt decrease above about 80100 C, in contrast to the non-stored
samples, which have shown two different decreasing regimes, the first between 60
and 100 C, with a smaller slope, and the second between 100 and 175 C, with
a greater slope. This is due to the additional curing reaction occurring during the
storage of samples.
After 7 or 30 days storage of composites in air or water the two loss tangent
peaks observed for the non-stored samples shifted towards each other until a single
and higher peak appeared at an intermediate temperature (Figs 3 and 4). The Tg
values of both composites stored in water was lower than those of samples stored in
air for the same period of storage, due to the plasticizing effect of water (Table 4).
The Tg values of Rok stored for 7 or 30 days in air or water were not significantly
different from the corresponding values of Ice.

5. Conclusions

The light-cured samples contain unreacted C=C methacrylate bonds. Storage of


them in air or water 37 C caused additional curing reaction. Therefore, in the aque-
ous oral environment at normal mouth temperature of 37 C it is clear that we can
observe a slight increase in the degree of conversion. On the other hand, the inges-
tion of hot food and beverages would only have limited impact on the degree of
conversion since resin composites are not thermal conductors and it is known that
an increase of only 5.5 C of the tooth can lead to irreversible adverse effects on
the pulp tissues (pulpitis and secondary necrosis). By the way, an increase of the
degree of the degree of conversion is more probably observed only at the surface of
the restoration than in the depth of the material.
The elastic modulus and tan of the two composites are not statistically different
(P  0.05) at 2550 C. The elastic modulus of Rok at 37 C and 1.0 Hz (E  = 8.40
(0.61) GPa) and that of Ice (E  = 7.58 (0.50) GPa) was much lower than that of
dentine (E  = 15.54 GPa).
1808 I. D. Sideridou et al. / Journal of Biomaterials Science 20 (2009) 17971808

References
1. J. L. Ferracane, Crit. Rev. Oral Biol. Med. 6, 302 (1995).
2. F. Lutz and R. W. Phillips, J. Prosthet. Dent. 50, 480 (1983).
3. N. Moszner and S. Klapdohr, Int. J. Nanotechnol. 1, 130 (2004).
4. S. C. Bayne, J. Dent. Educ. 69, 571 (2005).
5. K. Saber-Sheikh, R. L. Clarke and M. Braden, Biomaterials 20, 817 (1999).
6. R. V. Mesquita and J. Geis-Gerstorfer, Dent. Mater. 24, 623 (2008).
7. R. V. Mesquita, D. Axmann and J. Geis-Gerstorfer, Dent. Mater. 22, 258 (2006).
8. N. Emami and K.-J. M. Sderholm, Dent. Mater. 21, 977 (2005).
9. J. Sabbagh, J. Vreven and G. Leloup, Dent. Mater. 18, 64 (2002).
Downloaded by [University of Auckland Library] at 15:12 26 October 2014

10. Y. Abe, P. Lambrechts, S. Inoue, M. J. A. Braem, M. Takeuchi, G. Vanherle and B. Van Meerbeek,
Dent. Mater. 17, 520 (2001).
11. P. H. Jacobsen and A. H. Darr, J. Oral Rehabil. 24, 265 (1997).
12. D. Sideridou, M. M. Karabela and E. Ch. Vouvoudi, Dent. Mater. 24, 737 (2008).
13. J. F. Bates, G. D. Stafford and A. Harrison, J. Oral Rehabil. 2, 281 (1975).
14. J. K. Lee, J.-Y. Choi, B.-S. Lim, Y.-K. Lee and R. L. Sakaguchi, J. Biomed. Mater. Res. Part B:
Appl. Biomater. 68B, 216 (2004).
15. R. L. Clarke, Biomaterials 10, 549 (1989).
16. R. Chartoff, P. T. Weissman and A. Sircar, in: Assignment of the Glass Transition, ASTM STP
1249, R. J. Seyler (Ed.), p. 88. ASTM, West Conshohocken, PA (1994).
17. E. A. Turi (Ed.), in: Thermal Characterization of Polymeric Materials, 2nd edn, Vol. 1, p. 529.
Academic Press, New York, NY (1997).
18. J. S. Rees, P. H. Jacobsen and J. Hickman, Clin. Mater. 17, 11 (1994).
19. L. E. Neilsen, J. Appl. Phys. 41, 4626 (1970).
20. Y. Tanimoto, T. Kitagawa, M. Aida and N. Nishiyama, Acta Biomater. 2, 633 (2006).
21. L. E. Neilsen and R. F. Landel (Eds), in: Mechanical Properties of Polymers and Composites, 2nd
edn, p. 377. Marcel Dekker, New York, NY (1994).
22. S. N. Nazhat, R. Joseph, M. Wang, R. Smith, K. E. Tanner and W. Bonfield, J. Mater. Sci. Mater.
Med. 11, 621 (2000).

Anda mungkin juga menyukai