Está en la página 1de 12

CRITICAL REVIEWS IN ORAL BIOLOGY & MEDICINE

M.-H. Chen
Graduate Institute of Clinical Dentistry, School of Dentistry,
National Taiwan University, Taipei 100, Taiwan, R.O.C.,
and Department of Dentistry, National Taiwan University
Hospital; minhueychen@ntu.edu.tw

Update on Dental
Nanocomposites

J Dent Res 89(6):549-560, 2010

Abstract
Dental resin-composites are comprised of a photopolymerizable organic resin matrix and mixed
with silane-treated reinforcing inorganic fillers. In
the development of the composites, the three main
components can be modified: the inorganic fillers,
the organic resin matrix, and the silane coupling
agents. The aim of this article is to review recent
studies of the development of dental nanocomposites and their clinical applications. In nanocomposites, nanofillers are added and distributed in a
dispersed form or as clusters. For increasing the
mineral content of the tooth, calcium and phosphate ion-releasing composites and fluoridereleasing nanocomposites were developed by the
addition of DCPA-whiskers or TTCP-whiskers or
by the use of calcium fluoride or kaolinite. For
enhancing mechanical properties, nanocomposites
reinforced with nanofibers or nanoparticles were
investigated. For reducing polymerization shrinkage, investigators modified the resin matrix by
using methacrylate and epoxy functionalized
nanocomposites based on silsesquioxane cores or
epoxy-resin-based nanocomposites. The effects of
silanization were also studied. Clinical consideration of light-curing modes and mechanical properties of nanocomposites, especially strength
durability after immersion, was also addressed.

KEY WORDS: dental nanocomposites, nanofillers, resin-matrix, silanization.

Dental Composite Resins

odern dental composite restorative materials began with the discovery, in


the early 1960s, of Bowens Bis-GMA (2,2-bis[4-(2-hydroxy-3-methacryloxypropoxy)phenyl]-propane) with inorganic particle formulations. Dental
composites are composed of synthetic polymers, inorganic fillers, initiators,
and activators that promote light-activated polymerization of the organic
matrix to form cross-linked polymer networks, and silane coupling agents
which bond the reinforcing fillers to the polymer matrix.

Matrix
Most commercial dental composites use Bis-GMA monomer as their organic
matrix. Other base monomers used in current commercial composites include
triethyleneglycol dimethacrylate (TEGDMA), urethane dimethacrylate
(UDMA), ethoxylated bisphenol-A-dimethacrylate (Bis-EMA), decanediol
dimethacrylate (D3MA), bis(methacryloyloxymethyl) tricyclodecane, and urethane tetramethacrylate (UTMA). The most commonly used organic matrix,
Bis-GMA, has a very high viscosity due to the hydrogen bonding interactions
that occur between the hydroxyl groups and the monomer molecules. Therefore,
Bis-GMA must be diluted with more fluid monomers to provide the proper
viscosity for use in dental composites (Ferracane, 1995). TEGDMA, which is
less viscous and has excellent copolymerization characteristics, is frequently
used as the diluent monomer for UDMA and Bis-GMA-based composites to
provide a fluid resin that can be maximally filled with inorganic particles.
TEGDMA increases vinyl double-bond conversion (Reed et al., 1997). Since
UDMA and Bis-EMA have higher molecular weights and fewer double bonds
per unit of weight, they generally have less shrinkage than TEGDMA.
Therefore, TEGDMA has been replaced by UDMA and Bis-EMA in several
products to reduce shrinkage, aging, and the negative effects of environmental
factors such as moisture, acid, and temperature changes (Yap et al., 2000).

Fillers
Modern composite systems contain fillers such as quartz, colloidal silica, and
silica glass containing barium, strontium, and zirconium. These fillers
increase strength and modulus of elasticity and reduce polymerization shrinkage, the coefficient of thermal expansion, and water absorption.

Silane Coupling Agents

DOI: 10.1177/0022034510363765
Received November 17, 2008; Last revision October 28,
2009; Accepted November 9, 2009

The formation of a strong covalent bond between inorganic fillers and the
organic matrix is essential for obtaining good mechanical properties in dental
composites. Bonding of these two phases is achieved by coating the fillers
with a silane coupling agent that has functional groups to link the filler and
the matrix chemically. A typical coupling agent is 3-methacryloxypropyltrimethoxysilane (MPTS). One end of the molecule can be bonded to the

549
Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

550

Chen

J Dent Res 89(6) 2010

shrinkage is mainly due to the resin matrix, the increase in filler


level results in a lower amount of resin in nanocomposites and
will also significantly reduce polymerization shrinkage and dramatically improve the physical properties of nanocomposites.

Nanofillers in Commercial Nanocomposites

Figure 1. Nanomers and nanoclusters in Filtek Supreme nanocomposite. Scale bar = 0.6 m.

hydroxyl groups of silica particles, and the other end is capable


of copolymerizing into the polymer matrix.

Visible-light Initiator
Light-activated composite resins undergo polymerization by
irradiation via a blue-light-curing unit in the wavelength range
of 410-500 nm. Light in this region is most effectively absorbed
by an -diketone photoinitiator. Camphoquinone (CQ) is a commonly used visible-light initiator, and ethyl-4-(N,N-dimethylamino)benzoate (4EDMAB) is a commonly used co-initiator.
CQ creates an excited state that reacts with an aminereducing agent such as N,N-dimethylaminoethyl methacrylate
(DMAEMA) or ethyl p-dimethylaminobenzoate (DMAB) to
produce free radicals that initiate the cross-linking polymerization. The absorption spectrum of CQ lies in the 450- to 500-nmwavelength range, with peak absorption at 470 nm (Lee et al.,
1993).

Development of Nanocomposites
Nanotechnology, also known as nanoscience or molecular engineering, is defined as the creation of functional materials and
structures with characteristic dimensions in the range of 0.1-100
nm. When inorganic phases in an organic/inorganic composite
become nanosized, they are called nanocomposites.

Modification of Nanofillers
Nanofillers can be prepared by various techniques, such as
flame pyrolysis, flame spray pyrolysis, and sol-gel processes.
Because extremely small filler particles have dimensions below
the wavelength of visible light (0.4-0.8 m), they are unable to
scatter or absorb visible light. Thus, nanofillers are usually
invisible and offer the advantage of optical property improvement (Mitra et al., 2003). Additionally, nanofillers are capable
of increasing the overall filler level due to their small particle
sizes. More fillers can be accommodated if smaller particles are
used for particle packing. Theoretically, with the use of nanofillers, filler levels could be as high as 90-95% by weight. However,
the increase in nanofillers also increases the surface area of the
filler particles, which limits the total amount of filler particles
because of the wettability of the fillers. Since polymerization

There are several products of nanocomposites on the market.


Three nanocomposites with different compositions and nanosized fillers have been chosen as typical examples.
The first is Filtek Supreme (3M ESPE, St. Paul, MN, USA),
which contains nanometric particles (nanomers) and nanoclusters (NCs). Nanomers are monodispersed, non-agglomerated,
and non-aggregated silica particles of 20 and 75 nm in diameter.
Nanocluster fillers are loosely bound agglomerates of nanosized
particles (Fig. 1). Two types of NCs were synthesized. The first
type consisted of zirconia-silica particles synthesized from a
colloidal solution of silica and zirconyl salt. The primary particle size of the NC filler ranged from 2 to 20 nm, and the average
size of the spheroidal agglomerated particles was less than 0.6
m. The second type of NC filler was synthesized from 75-nm
primary particles of silica, and the average size of the agglomerated particles was 0.6 m. The silica particles were treated with
3-methacryloxypropyltrimethoxysilane (MPTS), as a coupling
agent, which contains a silica ester functional group on one end
for bonding to the inorganic surface and a methacrylate group
on the other end to make the filler compatible with the resin
before curing to prevent any agglomeration or aggregation
(Mitra et al., 2003). The filler content of Supreme is about
58-60% by volume and 78.5% by weight.
A second commercial product, Premise (Kerr/Sybron, Orange,
CA, USA), is a nanohybrid composed of 3 different types of filler
components: nonagglomerated discrete silica nanoparticles,
prepolymerized fillers (PPF), and barium glass fillers (Fig. 2).
The non-agglomerated discrete silica nanoparticles are spheroidal
and 20 nm in size. The prepolymerized fillers (PPF) are about
30-50 m in size, and the barium glass filler has an average particle size of 0.4 m. The technique for incorporating barium glass
fillers into the resin matrix in the nanocomposites uses the same
technology as the microhybrid resin Point 4 (Kerr/Sybron,
Orange, CA, USA). Compositionally, Point 4 is based on BisGMA resin matrix filled to 57.2% by volume (76% by weight)
with barium aluminoborosilicate glass and fumed silicon dioxide
filler particles. The filler particles have an average size of 0.4 m
(hence the name "Point 4"), and 90% of the particles are smaller
than 0.8 m. The manufacturer claims that by incorporating a
"polymerizable dispersant" it has been able to: (1) increase the
percentage filler level and (2) use filler particles with an average
size smaller than that used in traditional hybrid resin composites.
This provides the material with excellent esthetics and strength.
Thus, the nanocomposite Premise uses a similar technique for
incorporating barium glass fillers into the resin matrix, with a
trimodel approach to provide an optimal combination of 3 different filler components: silica nanoparticles, PPF, and barium
glass fillers. This combination of 3 fillers allows for increased
filler loading of 69% by volume and 84% by weight. The discrete
unassociated nanoparticles that are well-dispersed in the matrix
on a nanoscale level allow for increased filler loading and reduced

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

J Dent Res 89(6) 2010

Update on Dental Nanocomposites 551

Figure 2. Three different types of filler components: nonagglomerated


discrete silica nanoparticles, prepolymerized fillers (PPF), and barium glass fillers in Premise nanocomposite. Scale bar = 0.4 m.

viscosity of nanocomposites, and thus result in increased hardness, abrasion resistance, fracture resistance, and polishability and
in reduced polymerization shrinkage (reported to be 1.4% to 1.6%
by volume) and shrinkage stress (Bauer et al., 2003). As the interparticle dimension decreases, the load-bearing stress on the resin
is reduced, inhibiting crack formation and propagation. The spheroidal shape of the nanoparticles provides smooth and rounded
edges, distributing stress more uniformly throughout the composite resin (Terry, 2004a).
The third commercial nanocomposite, Ceram-X (Dentsply
DeTrey, Konstanz, Germany), is an ormocer-based, nanoceramic composite (Schirrmeister et al., 2006). Ceram-X contains glass fillers (1.1-1.5 m), but differs from conventional
hybrid composites in two important features: Methacrylatemodified silicon-dioxide-containing nanofiller (10 nm) substitutes for the microfiller that is typically used in hybrid
composites (agglomerates of silicon dioxide particles).
According to the manufacturers data, filler concentration is
57% by volume and 76% by weight. Furthermore, most of the
conventional resin matrix is replaced by a matrix full of highly
dispersed methacrylate-modified polysiloxane particles (2-3
nm). According to the manufacturers information, these nanoceramic particles are inorganicorganic hybrid particles. Both
nano-ceramic particles and nanofillers have methacrylate groups
available for polymerization (Fig. 3).
A summary of these 3 commercial nanocomposites is shown
in Table 1.

Reinforced Fillers
Nanofibers
For reinforcement of dental composites, electrospun nylon 6
nanocomposite nanofibers containing highly aligned fibriller
silicate single crystals were added in Bis-GMA/TEGDMA (Tian
et al., 2007). The hypothesis was that the uniform distribution of
nanoscaled and highly aligned fibrillar silicate single crystals
into nylon 6 nanofibers would improve the mechanical properties of the resulting nanocomposite nanofibers, and thereby
reinforce dental composites. Investigators first soaked the

Figure 3. Nanofillers in Ceram-X nanocomposite.

electrospun nylon 6/fibrillar silicate nanocomposite nanofiber


(about 100-400 nm and average 250 nm in diameter) felt with
CQ/4EDMAB-activated Bis-GMA/TEGDMA monomers. After
the soaked felt was photo-cured, the resulting composite was
milled into a powder with an average particle size of approximately 20 m. Subsequently, the powder was mixed with
CQ/4EDMAB-activated Bis-GMA/TEGDMA monomers to
prepare dental pastes containing nanofibers of various mass
fractions (1%, 2%, 4%, and 8%). Finally, the pastes were photocured, and the fabricated dental composites were characterized
and evaluated. The results indicate that small mass fractions (1%
and 2%) of nanofiber impregnation improved the mechanical
properties substantially, while larger mass fractions (4% and
8%) of nanofiber impregnation resulted in less desirable
mechanical properties, such as lower flexural strength, lower
elastic modulus, and lower fracture toughness of composites.
The results also suggested that to achieve better reinforcement,
the electrospun nanofibers may need to be collected as a highly
aligned yarn instead of randomly distributed felt.
Short E-glass Fibers
Nanohybrid composites with short E-glass fibers in combination
with composites were investigated (Garoushi et al., 2008). The
investigators hypothesized that E-glass fibers can induce the
light transmission of composite resin. Dimethacrylate (BisGMA) 66.7%, TEGDMA 32.6%, CQ and DMAEMA 0.7%
resin consisting of 50 wt% nanofibers (SiO2, 20 nm in size) and
E-glass fibers (3 mm in length) with Bis-GMA-PMMA resin
matrix and silane-treated radiopacity fibers of BaAlSiO2 (3 2
m in size) were incorporated within the resin matrix. In those
studies, experimental fiber composites were prepared by mixing
22.5 wt% of E-glass fibers and 22.5 wt% of resin matrix and the
gradually addition of 55 wt% of BaAlSi O2 radiopacity fibers. A
high-speed mixing machine was used at 3500 rpm for 5 min.
The dimethacrylate-based resin matrix containing PMMA
formed a semi-interpenetrating polymer network (semi-IPN)
matrix for the fiber composite. The results showed that E-glass
fiber-reinforced composite resin with semi-IPN matrix achieved
an acceptable depth of cure and microhardness, although lower
than those of commercial composite resins.

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

552

Chen

J Dent Res 89(6) 2010

Table 1. Comparison of Commercial Nanocomposites


Brand Name
Filtek Supreme

Type

Filler Composition, vol% (wt%)

Nanofilled

58-60 vol% (78.5 wt%)combination of aggregated zirconia-silica/silica


cluster filler with primary particle size of 2-20 nm and 75 nm, and nonagglomerated/non-aggregated 20-nm and 75-nm silica fillers
69 vol% (84 wt%)polymerized resin fillers (PPRF) (30-50 m), barium
glass (0.4 m), silica nanoparticles (20 nm)
57% vol% (76 wt%)contains glass fillers (1.11.5 m) with
methacrylate-modified silicon-dioxide-containing nanofiller (10 nm)

Premise

Nanohybrid

Ceram-X

Nanohybrid

TiO2 Nanoparticles

Manufacturer
3M ESPE, St. Paul, MN, USA

Kerr/Sybron, Orange, CA, USA


Dentsply DeTrey, Konstanz,
Germany

TTCP-whiskers

To improve the mechanical properties of composites, investigators modified the surfaces of TiO2 nanoparticles (< 20 nm) with
organosilane allytriethoxysilane (ATES), and the nanoparticles
were blended with resin monomers. The particles were then
manually added into Z100 dental-resin-based composites (3M,
ESPE) at different weight ratios, depending on the sample
group. The mixture was then thoroughly blended (Xia et al.,
2008). Two groups of specimens were used as controls for comparison, one with the same amount of unmodified TiO2 nanoparticles, and one without any TiO2 nanoparticles. It was found that
surface modification by the organosilane ATES influenced the
dispersion and linkage of TiO2 nanoparticles within a resin
matrix. Adding the modified TiO2 nanoparticles improved the
microhardness and flexure strength of the composites.

Caries-prevention Fillers

Other calcium and phosphate ion releasing composites with fine


TTCP particles and nanosilica-fused whiskers have been developed (Xu et al., 2009). TTCP enabled the composite to release
Ca and PO4 ions, while the whiskers provided the needed loadbearing ability. It was found that the TTCP-whisker composite
has strengths about two-fold those of the TTCP composite without whiskers. The TTCP-whisker composite increased Ca and
PO4 release by about six-fold when the pH was reduced from
neutral to 4. After immersion, the TTCP-whisker composite
matched the strength of non-releasing hybrid composite (TPH)
at all 3 pHs, 7.4, 6, and 4 (Xu et al., 2009). It was suggested that
this composite may have the potential to provide the necessary
combination of load-bearing and caries-inhibiting capability
(Xu et al., 2009).
Calcium Fluoride

Calcium and Phosphate Ion-releasing Fillers


To increase mineral content to control dental caries, calcium and
phosphate ion-releasing fillers have been developed, such as
nanoparticles of dicalcium phosphate anhydrous (DCPA) (Xu
et al., 2006, 2007a) and tetracalcium phosphate [TTCP:
Ca4(PO4)2O]-whiskers (Xu et al., 2009).
DCPA-whiskers
In the study of DCPA nanoparticles, 2 types of fillers were used:
DCPA particles and nano-silica fused whiskers (Xu et al., 2007a).
It was found that decreasing the DCPA particle size decreased the
composites strength, while whisker reinforcement more than
doubled the composites strength and significantly increased the
elastic modulus. The investigators also found that silanization of
the DCPA particles increased the composites strength, but
decreased the Ca and PO4 release. The use of unsilanized nano
DCPA together with whisker reinforcement appeared to be the
best method to produce a composite with high strength and Ca
and PO4 release. Previous studies (Skrtic et al., 1996a, 1996b,
2000; Dickens et al., 2003) showed that when Ca and PO4 ions
were released from the composite restoration, they re-precipitated
to form hydroxyapatite outside the composite and inside the tooth
lesion, significantly increasing the mineral content of the lesion.
Increasing the DCPA particle surface area significantly increased
the Ca and PO4 release, and composites with the nano DCPA
exhibited the highest release (Xu et al., 2007b).

Fluoride release from restorative materials is considered to


inhibit tooth demineralization and caries development and also
to strengthen the neighboring enamel or dentin (Eichmiller and
Marjenhoff, 1998; Forsten, 1998; Burke et al., 2006; Wang
et al., 2007). Xu et al. (2008) added calcium fluoride (CaF2)
nanoparticles (30%) and reinforcing whisker fillers (35%) to
nanocomposites and found that the fluoride release was better
than that of traditional and resin-modified glass-ionomer materials. The strength and elastic modulus of the nanocomposite also
matched those of commercial stress-bearing, non-releasing
composites. Nanocomposites containing CaF2 and DCPA, which
can release F, Ca, and PO4 ions for precipitation of fluoroapatite
and inhibition of caries, were also formulated with good
mechanical properties (Xu et al., 2008).
Polymer-kaolinite
Wang et al. (2007) developed 3 types of fluoride-releasing
polymer-kaolinite nanocomposites, including C(K-diamine),
C(K-acrylamide), and C(K-acetate). In their study, kaoline
[Al2Si2O5(OH)4] was used for the intercalation reaction, because
it is a layered aluminosilicate of the 1:1 type that is formed by 2
different types of interlayer surfaces. Aluminum atoms coordinate octahedrically with oxygen and hydroxyl groups on one
lamellar face, and silicon atoms coordinate tetrahedrically with
oxygen atoms on the other lamellar face (Thompson and
Cuff, 1985). Adjacent layers are linked to one another by
hydrogen bonds (Al-O-H-O-Si). Because of these structural

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

J Dent Res 89(6) 2010

Update on Dental Nanocomposites 553

characteristics and its high surface area for adsorption of fluoride, kaolinite is an excellent carrier compared with the conventional SiO2 filler used in composite resins. Therefore, these
polymer-kaolinite nanocomposites have the potential to provide
sustained release of fluoride due to strong adsorption of kaolinite to fluoride during the fabrication process (Wang et al., 2007).
According to their studies, the fluoride release and recharge
properties of C(K-acrylamide) were superior to those of Fuji IX
and Z-100. These findings suggest that C(K-acrylamide) might
be a useful dental restorative material with caries-preventive
properties (Wang et al., 2007).

Modification of Resin Matrix


All methacrylate resins shrink to a greater or lesser degree
according to the number of polymerizable units which they contain. This shrinkage is approximately equal to 22.5 x 10-6 m3mol-1
of moieties polymerized (Patel et al., 1987) and 2-14% volumetric shrinkage on average (Emami et al., 2003; Soh et al., 2006,
2007a). This shrinkage is naturally directly linked to the degree
of conversion of the methacrylate resin and has been used by
some workers to determine the actual degree of conversion
(Rueggeberg and Tamareselvy, 1995). There are two main strategies to reduce polymerization shrinkage: One is to reduce the
reactive sites per volume unit, the other is to use different types
of resin.
The density of reactive sites per volume unit can be reduced
principally in two ways. The first method is to increase the
molecular weight per reactive group. The second method is to
increase the filler load. However, there are some limitations in
these two methods. The use of high-molecular-weight monomers is limited by their viscosity, increased stickiness, and
undesirable general rheology, which compromise the handling
characteristics of the resulting composites. The filler load is
limited, because a given amount of resin can incorporate only a
limited amount of filler particles without adversely affecting the
wetting of the increased filler surface (Weinmann et al., 2005).
Recently, investigators made several attempts to reduce
shrinkage by changing the nature of the resin. One approach was
the use of liquid crystalline monomers as a resin; this was found
to shrink less, due to the transition of its nematic phase to an
isotropic amorphous state when photocured (Rawls et al., 1997).
There were also some monomers developed with the goal of
reducing polymerization shrinkage and its associated stresses,
such as oxetanes (Nuyken et al., 1996), oxybismethacrylates
(Stansbury, 1992), and highly branched methacrylates (Klee
et al., 1999). Other classes of monomers which have been investigated include ring-opening molecules such as spiro-ortho
esters (Miyazaki et al., 1994) and vinyl cyclopropane derivatives which can be copolymerized with methacrylate-based
resins (Moszner et al., 1999). Smith et al. (2004) developed the
cationic ring-opening spiro-ortho carbonates in combination
with epoxy monomers. In their study, they investigated the photocationic polymerization of an expanding monomer, 1,5,7,11tetraoxaspiro[5.5]undecane (TOSU), and an aromatic dioxirane,
bisphenol A diglycidyl ether (BADGE). Both homopolymerizations and binary polymerizations were conducted.

Epoxy-Polyol Matrix
In general, resins shrink less by ring-opening polymerization
(ROP), because of the increase in excluded free-volume associated with the ring-opening process (Tilbrook et al., 2000).
Epoxy-polyols were shown to have significant advantages over
dimethacrylates, including lower polymer shrinkage, no oxygen
inhibition layer, higher strength, and equivalent hardness, as
well as acceptable glass transition temperatures (Tilbrook et al.,
2000). However, the mass increase of the epoxy-polyol materials at 37oC is almost double that of the conventional dimethacrylate matrices, and cracking was observed in some samples
during hydration. The properties of epoxy-polyol matrices can
be controlled by varying their formulation. The key parameter in
modifying their properties is the molar ratio of epoxy groups to
polyol groups (R), which, for a sample mixture of one epoxide
resin and one polyol, is calculated according to the equation
R=

Wepoxy
Wpolyol

RMMpolyol
RMMepoxy

fepoxy
fpolyol

where Wepoxy and Wpolyol, RMMepoxy and RMMpolyol, and fepoxy and
fpolyol are the weights (in grams), the relative molecular masses,
and the functionalities of the epoxy resin and polyol, respectively. By varying R, it is possible to vary the properties and
cure characteristics of the resin system. It is generally accepted
that decreasing R increases the cure speed. This increase in
reactivity is usually accompanied by an increase in flexibility
and toughness (Tilbrook et al., 2000). It was concluded that R
should be between the limits of 4 and 8 to ensure that the matrix
has a balance of acceptable properties (Tilbrook et al., 2000).

Epoxy Functionalized Cyclic SiloxaneSiloraneTM


SiloraneTM (Fig. 4) is an epoxy functionalized cyclic siloxane
whose name is derived from the combination of its chemical
building blocks siloxanes and oxiranes. The SiloraneTM material
is a cationic ring-opening monomer system, with the target profile of a low-shrinkage, high-reactive, biocompatible composite
(Schweikl et al., 2002, 2004) developed by 3M ESPE (Seefeld,
Germany). The network of Silorane is generated by the cationic
ring-opening polymerization of the cycloaliphatic oxirane
moieties. The cationic cure starts with the initiation process
of an acidic cation which opens the oxirane ring and generates
a new acidic center, a carbocation. After the addition to an
oxirane monomer, the epoxy ring is opened to form a chain or,
in the case of two- or multifunctional monomers, a network.
Nanocomposites containing Silorane and fillers exhibit low
shrinkage and comparable mechanical properties (Weinmann
et al., 2005).

Epoxy Resin ERL4221


In our previous study (Chen et al., 2006), we developed a
low-shrinkage, high-strength nanocomposite by using a 4epoxycyclohexylmethyl-(3,4-epoxy) cyclohexane carboxylate
(ERL4221) (Fig. 5) matrix with 55% of 70- to 100-nm nanosilica fillers through ring-opening polymerization. In our

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

554

Chen

J Dent Res 89(6) 2010

Figure 6. TEM of epoxy-resin-based nanocomposite with GPS for silanization demonstrated no aggregation of silica nanoparticles. Scale
bar = 25 nm.

Figure 4. Silorane monomer.

O
O

Figure 5. Chemical structure of ERL4221.

design, we used -glycidoxypropyl trimethoxysilane (GPS) to


modify the surfaces of the silica nanoparticles. The nanocomposite was shown to exhibit low polymerization shrinkage
strain, which was only a quarter of that of currently used
methacrylate-based composites. It also exhibited a low thermal
expansion coefficient comparable with that of the methacrylate-based composites. The strong interfacial interactions
between the resin and fillers at nanoscales were demonstrated
by the observed high strength and high thermal stability of the
nanocomposite. A transmission electron microscope (TEM)
study of the nanocomposite showed no aggregation of nanoparticles (Fig. 6).

Silsesquioxane (SSQ)
Soh et al. (2007a) found that the hardness and modulus of
nanocomposites with different wt% silsesquioxane (SSQ)
ratios decreased when increased amounts of SSQ monomers
were added, indicating that the incorporation of SSQ monomers into the control generally helped to reduce both rigidity
and polymerization shrinkage. The results demonstrated that,
in the correct formulation, SSQ materials have great potential
to be used as low-shrinkage composites. In other experiments,
nanocomposites based on multifunctional SSQ with reduced

shrinkage have also been reported (Soh et al., 2007b). For


multifunctional SSQ-based nanocomposites, SSQ with various
amounts of methacrylates and/or epoxide groups were prepared via Pt-catalyzed hydrosilylation of 8 equivalents of
di(propylene glycol) allyl ether methacrylate and/or propargyl
methacrylate, and 4-vinyl-cyclohexene epoxide combinations
with (HMe2SiOSiO1.5)8. The results also demonstrated that
these materials have significantly lower shrinkage during
curing than traditional monomers used for dental composite
applications.

Bioactive Poly(methyl methacrylate)/SiO2-CaO


Nanocomposites with Dimethyldiethoxysilane
(DMDES)
Lee and Rhee (2009) developed a bioactive poly(methyl methacrylate)/SiO2-CaO nanocomposite using either dimethyldiethoxysilane (DMDES) or tetraethoxysilane (TEOS), which
could produce 2 and 4 siloxane linkages, respectively, after a
sol-gel reaction. In their study, methylmethacrylate was copolymerized with 3-(trimethoxysilyl) propyl methacrylate and
then co-condensed with either DMDES (specimen D) or TEOS
(specimen T), respectively, with calcium nitrate tetrahydrate
under acidic conditions. The results demonstrated that the
fracture toughness of specimen D was higher than that of
specimen T and did not lose apatite-forming ability. They suggested that these results were due to the decrease of siloxane
linkage numbers and the introduction of alkyl groups into the
silica structure in specimen D. The covalently bonded siloxane
linkages may produce hard and brittle fracture behavior in the
nanocomposite, while the alkyl groups help to make the silica
a linear chain structure. This nanocomposite can be applied
to the filler materials for bone cement and dental composite
resin because of its good bioactivity and improved mechanical
properties.

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

J Dent Res 89(6) 2010

Update on Dental Nanocomposites 555

Modification of Silanes
Silanization with MPTS and OTMS
Nanoparticles often form microscopic aggregations during
processing. To eliminate this, the silica fillers are often coated
with MPTS (Fig. 7a) to deter particle aggregation and promote
interfacial adhesion by allowing the particle surface to copolymerize with the matrix polymer through covalent and H-bonding
(Wilson et al., 2005). By contrast, n-octyltrimethoxysilane
(OTMS) (Fig. 7b) is a non-reactive aliphatic silane which does
not react with the resin matrix, but interacts through weak von
der Waals forces. Dual-silanization of silica particles with
blends of reactive MPTS and non-reactive OTMS offers several
potential advantages compared with silanization with MPTS
only. These advantages include improved uncured pastehandling characteristics, higher double-bond conversion during
photopolymerization (Wilson et al., 2005), improved durability
in the aqueous oral environment, and lower polymerization
stress through non-bonded nanofiller particles (Condon and
Ferracane, 2002). Wilson et al. (2007) also found that the covalent bonding and H-bonding of MPTS-rich nanoparticles with
the matrix are necessary for preparing well-dispersed nanocomposites and interphases containing equal masses of MPTS and
OTMS, to yield composites with optimal properties. The effects
of the interface are more prominent in nanocomposites than in
conventional composites containing microsized fillers, because
nanocomposites and their clusters have a much larger surface
area per unit mass, and the film derived from organosilane can
influence dispersion and improve bonding between inorganic
nanoparticles and the resin matrix (Mohsen and Craig, 1995).

Silanization with UDMS, MPTS, and OTMS


Karabela and Sideridou (2008) compared the absorption characteristics of a water or ethanol/water solution (75 vol%/25 vol%)
by dental nanocomposites consisting of a Bis-GMA/TEGDMA
(50/50, wt/wt) matrix and silica nanoparticles (Aerosil
OX50), silanized with various silanes. The silanes used were:
3-[(1,3(2)-dimethacryloyloxypropyl)-2(3)-oxycarbonylamido]
propyltriethoxysilane (UDMS), which is a urethane dimethacrylate silane (Fig. 7c); 3-methacryloxypropyltrimethoxysilane
(MPTS); octyltrimethoxysilane (OTMS); a blend of UDMS/
OTMS (50/50, wt/wt); and a blend of MPTS/OTMS (50/50, wt/
wt). In their studies, the silane structure used for the silanization
of nanosilica was found to affect the absorption behavior of an
water or ethanol/water solution (37oC) by the composites. The
composite containing the UDMS with the hydrophilic urethane
group showed the highest amount of absorbed water. The composite with the OTMS, which does not contain a methacrylate
moiety and cannot react with the dimethacrylate monomers,
showed the highest solubility in both water and ethanol/water. In
all composites, the amount of absorbed ethanol/water solution
was much higher than that of water. The OTMS-composite
absorbed the highest amount of ethanol/water, and the MPTScomposite absorbed the lowest amount. Therefore, it was concluded that the silane structure used for the silanization of
nanosilica has an effect on the solvent absorption and solubility
of composites (Karabela and Sideridou, 2008).

Figure 7. Chemical structure of silanes.

Silanization with GPS


As described in the matrix section of epoxy resin ERL4221,
GPS (Fig. 7d) has been used in epoxy-resin-based nanocomposites (Chen et al., 2006). With GPS as a coupling agent, the

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

556

Chen

nanosilica particles were well-dispersed in the epoxy resin without aggregation (Fig. 6). It was also found that a high concentration of SiO2 nanoparticles without the coupling agent GPS has a
higher tendency to aggregate into a macrosize cluster that scatters light and reduces curing efficiency.

Silanization with ATES


Organosilane allytriethoxysilane (ATES) (Fig. 7e) was used in
the TiO2 nanoparticle-reinforced nanocomposite developed by
Xia et al. (2008), who found that surface modification by the
organosilane ATES increases the dispersion and linkage of TiO2
nanoparticles within the resin matrix.

Silanization for Improving Fracture Toughness


Chan et al. (2007) proposed that the near-tip fracture processes
in nanocomposites involve several sequences of fracture events,
including (1) particle bridging, (2) debonding at the poles of
the particle/matrix interface, and (3) crack deflection around
the particles. The fracture processes of crack deflection and
interface cracking were modeled for investigation of their role
in the fracture toughness of the nanocomposites with various
levels of particle fillers. Nanosized particles appear to enhance
fracture toughness of nanocomposites in two ways: (1) The
large surface-to-volume ratio improves interface bonding and,
consequently, interface toughness; and (2) the high strength of
the nanosized particles helps to prevent particle fracture during
interface cracking. With nanosized particles, the interface toughness can be increased to higher levels without the risk of causing
particle fracture, and thus allows a higher fracture toughness
value to be attained in nanocomposites. By using analytical and
finite-element methods to model the observed sequences of
fracture events, these investigators found that silanization and
nanoparticle loading improved the fracture toughness of dental
nanocomposites through a combination of enhanced interface
toughness by silanization, crack deflection, and crack bridging
(Chan et al., 2007).

J Dent Res 89(6) 2010

Supreme). There was no significant difference in the average


size of the opposing enamel wear facet generated by the different composites. Microfill composite (Heliomolar RO) resulted
in a significantly rougher surface within the wear track than
either nanohybrid composite (Premise) or microhybrid composite (Point 4), but was not significantly different from nanofilled
composite (Filtek Supreme) (Yesil et al., 2008).

Consideration of Light-curing Modes


Nanocomposites may present higher degradation in the oral
environment than hybrid composites (da Silva et al., 2008). A
soft-start light-activation mode may increase the solubility of
resin composites. However, Ilie et al. (2005) suggested that the
soft-start polymerization concept is still valid, even with highpower light-emitting diode (LED) curing units. In their study, it
was found that a soft-cure polymerization resulted in reduced
shrinkage stress while simultaneously keeping the degree of
cure and mechanical properties constant. Atai and Motevasselian
(2009) reported that the Filtek Supreme (R) nanocomposite
showed less increase in temperature and a lower degree of conversion in comparison with the hybrid composite. It was also
found that the LED curing unit induced considerable total and
irradiation temperature increase without any improvement in the
degree of conversion. Ramp-curing mode (soft-start mode)
showed lower temperature rise and delayed gel point and was
found to be more effective than the quartz-tungsten-halogen
(QTH) standard mode and LED units. The study showed that the
LED curing units have no advantage over conventional QTH
units in terms of temperature rise and degree of polymerization
conversion. Gritsch et al. (2008) compared the role of light
parameters on nanohybrid composite curing. In their study, 2
nanohybrid resins were cured by 2 LED devices and by 1 QTH
device with different combinations of energy density and power
density. The results indicated that, above a certain energy density threshold, the power density may not significantly influence
the polymerization kinetics.

Clinical Considerations

Consideration of Mechanical Properties

Nanocomposites have been widely applied in dental clinics.


Terry (2004) presented the direct applications of nanocomposites with positive response. He also concluded that, if a patients
condition is clearly and thoroughly evaluated pre-operatively,
the nanocomposites can provide an esthetic and natural appearance (Terry and Leinfelder, 2008). Yesil et al. (2008) compared
the wear resistance of 2 commercial nanocomposites, Filtek
Supreme (3M ESPE, St. Paul, MN, USA) and Premise (Kerr/
Sybron, Orange, CA, USA), with those of the more traditional
microhybrid composite, Point 4 (Kerr/Sybron, Orange, CA,
USA), and a microfill composite, Heliomolar RO (Ivoclar
Vivadent, Amherst, NY, USA). They found that the compositeresin type did not significantly affect the amount of measured
attrition, but did significantly affect abrasive wear. The conventional microfill composite resin (Heliomolar RO) exhibited
significantly less abrasive wear than the nanohybrid material
(Premise), but was not significantly different from the conventional microhybrid (Point 4) or nanofilled composite (Filtek

Mitra et al. (2003) reported that compressive and diametral tensile strengths and the fracture resistance of nanocomposites
were equivalent to or higher than those of the other commercial
composites that they tested. The nanocomposites also showed
better polish retention than the hybrids and microhybrids tested
after extended brushing periods. Watanabe et al. (2008) demonstrated that the fracture toughness values of hybrid and nanoparticle resin composites are significantly higher than those of
micro-filled resin composites. However, Yesil et al. (2008)
reported that nanocomposites did not significantly improve
wear resistance or the amount of opposing cusp wear when
compared with the traditional materials tested. Curtis et al.
(2008) found that water uptake and mechanical properties of
composites were influenced by the size and morphology of the
reinforcing particulate phase.
Curtis et al. (2009) investigated 7 commercially available
resin-based composites. After being pre-loaded, the specimens
were stored either in a lightproof container or a water-bath for

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

J Dent Res 89(6) 2010

Update on Dental Nanocomposites 557

Table 2. Development of Nanocomposites by Modification of Fillers for


Reinforcement and Caries Prevention

Table 3. Development of Nanocomposites by Modification of Resin


Matrix to Reduce Polymerization Shrinkage

Modification

Modification

Reinforced
fillers

Reinforced
fillers

Reinforced
fillers

Caries
prevention
fillers

Caries
prevention
fillers
Caries
prevention
fillers

Contents

Special Concerns

Electrospun nylon 6
Reinforcement (Tian
nanocomposite nanofibers
et al., 2007)
(100-400 nm) containing
highly aligned silicate
single crystals
Nanohybrid composites with Reinforced (Garoushi
50 wt% nanofibers (SiO2,
et al., 2008)
20 nm in size) and E-glass
fibers (3 nm in length) with
BaAlSiO2 in semi-IPN*
matrix
TiO2 nanocomposites (< 20 Increased micronm) modified with ATES
hardness and flexure
strength (Xia et al.,
2008
Nano-DCPA whiskers
Ca and PO4 releasing
nanofillers for caries
prevention (Xu et al.,
2006, 2007a,b)
TTCP-whiskers
Ca and PO4 releasing
nanofillers for caries
prevention (Xu et al.,
2009)
30% CaF2 nanoparticles with Fluoride released for
reinforcing whisker fillers
caries prevention
(35%)
(Xu et al., 2008)
Polymer-kaolinite
Fluoride released for
nanocomposite
caries prevention
(Wang et al., 2007)

* semi-IPN: semi-interpenetrating polymer network.


ATES: organosilane allytriethoxysilane.

24 hrs prior to being tested. They found that the nanocluster


system provided a distinct reinforcing mechanism compared
with the microhybrid, microfill, or nanohybrid resin-based composites, resulting in significant improvements to the strength
and reliability, regardless of environmental storage or testing
conditions. They suggested that the agglomerated nanoparticles
produced an interconnected network where the interstices were
infiltrated with the silane coupling agent, producing an interpenetrating phase composite (IPC) structure. In a wet testing
environment, hydrolysis and polymerization within the nanocluster silane phase could modify stress transfer both to and within
the cluster particles, producing an enhanced capacity to tolerate
local stresses and cluster deformation (Curtis et al., 2009).
However, such improvements in performance may be compromised over time by hydrolytic degradation of the silane (Curtis
et al., 2008). There are several commercial nanocomposites that
have good strengths after 1 day of immersion in water. However,
their strength can decrease by more than 50% after just a couple
of months immersion (Curtis et al., 2008). Therefore, the
strength durability is also an important issue, especially in
ion-releasing composites. Xu et al. (2006) found that most

Epoxy-polyol matrix

Contents
Epoxy-polyols

Epoxy functionalized SiloraneTM


cyclic siloxane

Bioactive poly(methyl Poly(methyl


methacrylate)/
methacrylate)/
SiO2-CaO
SiO2-CaO
nanocomposite
nanocomposite
using DMDES*
and TEOS
Epoxy resin ERL
Epoxy resin ERL
4221
4221 with 50%
70-100 nm
nanosilica
modified with
GPS
Silsesquioxane
Silsesquioxane
(SSQ)
(SSQ)-based
nanocomposites

Special Concerns
Ring-opening
polymerization for
reducing shrinkage
(Tilbrook et al., 2000)
Cationic ring opening
with low shrinkage
(3M ESPE, Seefeld,
Germany)
Bioactive and improved
mechanical properties
(Lee and Rhee, 2009)

Ring-opening
polymerization with
low shrinkage and
high strength (Chen
et al., 2006)
Ring-opening with low
shrinkage (Soh et al.,
2007a,b)

* DMDES: dimethyldiethoxysilane.
TEOS: tetraethoxysilane.
ERL 4221: 3,4-epoxycyclohexylmethyl-(3,4-epoxy)cyclohexane carboxylate.

composites did not show a significant decrease in strength after


1 days immersion. Only the chemically cured nano-DCPAwhisker composites with HEMA at DCPA:Whisker = 1:2 had a
significant strength loss. After 56-day immersion, the chemically cured DCPA-whisker composites (with and without
HEMA) showed significant strength loss (approximately
20-30%). In another study, Xu et al. (2009) demonstrated that,
after immersion in solution at pH of 7.4, 6, and 4, for 28 days,
the TTCP-whisker composite, similar to a commercial hybrid
composite, did not show reduced strength, in contrast to a significant strength loss for a releasing control material.
Ilie and Hickel (2009) analyzed the mechanical behavior of the
silorane-based composite in comparison with that of 6 homologous clinically successful methacrylate-based composites. They
found that the silorane-based composite was comparable with
clinically successful methacrylate-based composite materials,
encouraging the clinical use of the new composite material.
According to the kinetic model for the shrinkage-strain rates
of dental resin composites developed by Atai and Watts (2006),
there is a linear correlation between the shrinkage-strain (and
shrinkage-strain rate) and filler-volume fraction. The filler fraction did not affect the degree-of-conversion of the composites.
It was suggested that only a relatively high filler-surface area, as
may be obtained with nanofillers, will affect the networkforming kinetics of the resin matrix.

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

558

Chen

Table 4. Development of Nanocomposites by Modification of Nanoparticle


Surfaces with Different Silanes
Modification
MPTS*

Effects

Particles copolymerized
with matrix polymer
through covalent and
H-bonding
OTMS
Non-reactive aliphatic
silane, not reacting
with resin matrix, but
interacting through
weak van der Waals
forces
Dual silanization
Improved uncure
with MPTS and
beyond handling,
OTMS
higher double-bond
conversion during
polymerization
Equal masses of
Optimal properties,
MPTS and OTMS
prominent interface
in nanocomposites
UDMS
With hydrophilic
urethane group,
showed higher water
absorption
OTMS
Does not contain a
methacrylate moiety
and cannot react
with the
dimethacrylate
monomers; showed
higher solubility in
water and ethanol/
water
MPTS
Lower ethanol/water
absorption (Karabela
and Sideridou,
2008)
GPS
Used in the epoxy resin
ERL-4221- based
nanocomposite
ATES
Used in TiO2-reinforced
nanocomposites

Special Concerns
Deter particle
aggregation and
promote interfacial
adhesion
Non-reactive with
resin, interacting
through weak van
der Waals forces

Improved durability in
aqueous oral
environment

Optimal properties
(Wilson et al.,
2005, 2007)

Nanosilica particles
well-dispersed (Chen
et al., 2006)
Dispersion and
linkage of TiO2
nanoparticles (Xia
et al., 2008)

* MPTS: 3-methacryloxypropyltrimethoxysilane.
OTMS: n-octyltrimethoxysilane.
UDMS: 3-[(1,3(2)-dimethacryloyloxypropyl)-2(3)-oxycarbonylamido]propyltriethoxysilane.
GPS: -glycidoxypropyl trimethoxysilane.
ATES: organosilane allytriethoxysilane.

Ozel et al. (2008) investigated the cervical microleakage and


internal voids of nanocomposites, comparing them with a hybrid
composite in Class II restorations with the margins located coronal and apical to the cement-enamel junction (CEJ). The results
indicated that the location of the gingival margin affects the

J Dent Res 89(6) 2010

microleakage of nanocomposites, but has no significant effect


on the internal voids.

Conclusion
Nanocomposites allow for increased filler loading and a reduced
amount of resin matrix, thereby reducing polymerization shrinkage while providing esthetics and strength. There are many
commercial nanocomposites, and 3 typical examples
Filtek Supreme (3M ESPE, St. Paul, MN, USA), Premise
(Kerr/Sybron, Orange, CA, USA), and Ceram-X (Dentsply
DeTrey, Konstanz, Germany)are demonstrated (Table 1).
Nanocomposites can be strengthened by the addition of reinforced fillers with nanofibers (Tian et al., 2007), short E-glass
fibers (Garoushi et al., 2008), and TiO2 nanoparticles (Xia et al.,
2008) (Table 2). Ion-releasing nanocomposites can also be used
to increase the mineral content of dental caries lesions by the use
of nano-DCPA whiskers (Xu et al., 2007a) or TTCP-whiskers
(Xu et al., 2009) for releasing Ca and PO4 ions, by the use of
calcium fluoride (CaF2) nanoparticles for fluoride release (Xu
et al., 2008), or by the use of both CaF2 and DCPA for F, Ca, and
PO4 release (Xu et al., 2008). Polymer-kaolinite nanocomposites can also release fluoride, and C(K-acrylamide) might be
a useful material for caries prevention (Wang et al., 2007)
(Table 2). Nanocomposites with a ring-opening resin matrix can
reduce polymerization shrinkage by the use of epoxy-polyols
(Tilbrook et al., 2000), or by use of the commercial product
SiloraneTM (3M ESPE, Seefeld, Germany), which is generated
by the cationic ring-opening polymerization of the cycloaliphatic oxirane moieties (Weinmann et al., 2005), or by the use
of epoxy resin ERL 4221 (Chen et al., 2006) or SSQ monomers
(Soh et al., 2007a,b) (Table 3). Several approaches have been
used to deter nanoparticles from forming microscopic aggregations during processing. The same silica fillers are coated with
MPTS to deter particle aggregation and promote interfacial
adhesion through covalent and H-bonding (Wilson et al., 2005).
Silica fillers are modified by OTMS, a non-reactive aliphatic
silane, to interact through weak van der Waals forces (Wilson
et al., 2007). The silane structure used for the silanization of
nanosilica has an effect on solvent absorption and the solubility
of composites (Karabela and Sideridou, 2008). The composite
containing UDMS showed the highest amount of absorbed
water, the composite with OTMS showed the highest solubility
in both water and ethanol/water, and the one with MPTS
absorbed the least ethanol/water (Karabela and Sideridou,
2008). GPS can be used as a coupling agent in nanocomposites
with epoxy resin matrix (Chen et al., 2006). ATES increase the
dispersion and linkage of TiO2 nanoparticles within the resin
matrix (Xia et al., 2008). The chemical structure of different
silanes is shown in Table 4. There are several commercial nanocomposites that have good strengths after 1 day of immersion in
water (Curtis et al., 2008, 2009). However, their strength can
decrease by more than 50% after just a couple of months
immersion (Curtis et al., 2008). Therefore, strength durability is
an important issue, especially in ion-releasing composites (Xu
et al., 2006, 2009). Clearly, the development of nanocomposites
has led to significant improvements in dental materials and their

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

J Dent Res 89(6) 2010

Update on Dental Nanocomposites 559

clinical applications. However, there is still room for improvement in the properties of nanocomposites, so it is worthwhile to
continue further research.

ACKNOWLEDGMENTS
The author acknowledges grant support from National
Taiwan University and the National Science Council of Taiwan,
R.O.C., and English help from Professor Tim Casey at
the National Taiwan University for the preparation of this
paper.

References
Atai M, Motevasselian F (2009). Temperature rise and degree of photopolymerization conversion of nanocomposites and conventional dental
composites. Clin Oral Invest 13:309-316.
Atai M, Watts DC (2006). A new kinetic model for the photopolymerization
shrinkage-strain of dental composites and resin-monomers. Dent Mater
22:785-791.
Bauer F, Sauerland V, Ernst H, Glsel HJ, Naumov S, Mehnert R (2003).
Preparation of scratch and abrasion resistant polymeric nanocomposites
by monomer grafting onto nanoparticles. Macromol Chem Phys
204:375-383.
Burke FM, Ray NJ, McConnell RJ (2006). Fluoride-containing restorative
materials. Int Dent J 56:33-43.
Chan KS, Lee YD, Nicolella DP, Furman BR, Wellinghoff S, Rawls HR
(2007). Improving fracture toughness of dental nanocomposites by
interface engineering and micromechanics. Eng Fract Mech 74:18571871.
Chen MH, Chen CR, Hsu SH, Sun SP, Su WF (2006). Low shrinkage light
curable nanocomposite for dental restorative material. Dent Mater
22:138-145.
Condon JR, Ferracane JL (2002). Reduced polymerization stress through
non-bonded nanofiller particles. Biomaterials 23:3807-3815.
Curtis AR, Shortall AC, Marquis PM, Palin WM (2008). Water uptake and
strength characteristics of a nanofilled resin-based composite. J Dent
36:186-193.
Curtis AR, Palin WM, Fleming GJ, Shortall AC, Marquis PM (2009). The
mechanical properties of nanofilled resin-based composites: the impact
of dry and wet cyclic pre-loading on bi-axial flexure strength. Dent
Mater 25:188-197.
da Silva EM, Almeida GS, Poskus LT, Guimares JG (2008). Relationship
between the degree of conversion, solubility and salivary sorption of a
hybrid and a nanofilled resin composite. J Appl Oral Sci 16:161-166.
Dickens SH, Flaim GM, Takagi S (2003). Mechanical properties and biochemical activity of remineralizing resin-based Ca-PO4 cements. Dent
Mater 19:558-566.
Eichmiller FC, Marjenhoff WA (1998). Fluoride-releasing dental restorative
materials. Oper Dent 23:218-228.
Emami N, Sderholm KJ, Berglund LA (2003). Effect of light power density
variation on bulk curing properties of dental composites. J Dent
31:189-196.
Ferracane JL (1995). Current trends in dental composites. Crit Rev Oral Biol
Med 6:302-318.
Forsten L (1998). Fluoride release and uptake by glass-ionomers and related
materials and its clinical effect. Biomaterials 19:503-508.
Garoushi S, Vallittu PK, Lassila LV (2008). Depth of cure and surface
microhardness of experimental short fiber-reinforced composite. Acta
Odontol Scand 66:38-42.
Gritsch K, Souvannasot S, Schembri C, Farge P, Grosgogeat B (2008).
Influence of light energy and power density on the microhardness of
two nanohybrid composites. Eur J Oral Sci 116:77-82.
Ilie N, Hickel R (2009). Macro-, micro- and nano-mechanical investigations
on silorane and methacrylate-based composites. Dent Mater 25:
810-819.

Ilie N, Kunzelmann KH, Visvanathan A, Hickel R (2005). Curing behavior


of a nanocomposite as a function of polymerization procedure. Dent
Mater J 24:469-477.
Karabela MM, Sideridou ID (2008). Effect of the structure of silane coupling agent on sorption characteristics of solvents by dental resinnanocomposites. Dent Mater 24:1631-1639.
Klee JE, Neidhart F, Flammersheim HJ (1999). Monomers for low shrinking
composites. 2. Synthesis of branched methacrylates and their application in dental composites. Macromol Chem Phys 200:517-523.
Lee KH, Rhee SH (2009). The mechanical properties and bioactivity of
poly(methylmethacrylate)/SiO2-CaO nanocomposite. Biomaterials
30:3444-3449.
Lee SY, Chiu CH, Boghosian A, Greener EH (1993). Radiometric and spectroradiometric comparison of power outputs of five visible light-curing
units. J Dent 21:373-377.
Mitra SB, Wu D, Holmes BN (2003). An application of nanotechnology in
advanced dental materials. J Am Dent Assoc 134:1382-1390.
Miyazaki K, Takata T, Endo T, Inanaga A (1994). Thermal and photopolymerization of methacrylates containing a spiro ortho ester moiety
and the properties of polymethacrylates. Dent Mater J 13:9-18.
Mohsen NM, Craig RG (1995). Effect of silanization of fillers on their dispersability by monomer systems. J Oral Rehabil 22:183-189.
Moszner N, Volkel T, Fischer U, Rheinberger V (1999). Polymerization of
cyclic monomers, 8. Synthesis and radical polymerization of hybrid
2-vinylcyclopropanes. Macromol Rapid Commun 20:33-35.
Nuyken O, Bohner R, Erdmann C (1996). Oxetane photopolymerizationa
system with low volume shrinkage. Macromol Symp 107:125-138.
Ozel E, Korkmaz Y, Attar N (2008). Influence of location of the gingival
margin on the microleakage and internal voids. J Contemp Dent Pract.
9:65-72.
Patel MP, Braden M, Davy KW (1987). Polymerization shrinkage of methacrylate esters. Biomaterials 8:53-56.
Rawls HR, Wellinghoff VT, Norling BK, Leamon SH, Swynnerton NF,
Wellinghoff ST (1997). Low shrinkage resins from liquid crystal diacrylate monomers. Polym Prepr 38:167-168.
Reed BB, Choi K, Dickens SH, Stansbury JW (1997). Effect of resin composition of kinetics of dimethacrylate photopolymerization. Polym
Prepr 38:108-109.
Rueggeberg F, Tamareselvy K (1995). Resin cure determination by
polymerization shrinkage. Dent Mater 11:265-268.
Schirrmeister JF, Huber K, Hellwig E, Hahn P (2006). Two-year evaluation
of a new nano-ceramic restorative material. Clin Oral Invest 10:
181-186.
Schweikl H, Schmalz G, Weinmann W (2002). Mutagenic activity of structurally related oxiranes and siloranes in Salmonella typhimurium. Mutat
Res 521:19-27.
Schweikl H, Schmalz G, Weinmann W (2004). The induction of gene mutations and micronuclei by oxiranes and siloranes in mammalian cells
in vitro. J Dent Res 83:17-21.
Skrtic D, Antonucci JM, Eanes ED (1996a). Improved properties of amorphous calcium phosphate fillers in remineralizing resin composites.
Dent Mater 12:295-301.
Skrtic D, Hailer AW, Takagi S, Antonucci JM, Eanes ED (1996b).
Quantitative assessment of the efficacy of amorphous calcium phosphate/methacrylate composites in remineralizing caries-like lesions
artificially produced in bovine enamel. J Dent Res 75:1679-1686.
Skrtic D, Antonucci JM, Eanes ED, Eichmiller FC, Schumacher GE (2000).
Physiological evaluation of bioactive polymeric composites based on
hybrid amorphous calcium phosphates. J Biomed Mater Res Part B
53:381-391.
Smith RE, Pinzino CS, Chappelow CC, Holder AJ, Kostoryz EL, Guthrie
JR, et al. (2004). Photopolymerization of an expanding monomer with
an aromatic dioxirane. J Appl Polym Sci 92:62-71.
Soh MS, Sellinger A, Yap AU (2006). Dental nanocomposites. Current
Nanosci 2:373-381.
Soh MS, Yap AU, Sellinger A (2007a). Physicomechanical evaluation of
low-shrinkage dental nanocomposites based on silsesquioxane cores.
Eur J Oral Sci 115:230-238.

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

560

Chen

Soh MS, Yap AU, Sellinger A (2007b). Methacrylate and epoxy functionalized nanocomposites based on silsesquioxane cores for use in dental
applications. Eur Polymer J 43:315-327.
Stansbury JW (1992). Cyclopolymerizable monomers for use in dental resin
composites. J Dent Res 71:434-437.
Terry DA (2004). Direct applications of a nanocomposite resin system: Part
1The evolution of contemporary composite materials. Pract Proced
Aesthet Dent 16:417-422.
Terry DA, Leinfelder KF (2008). A simplified aesthetic concept: historical
review and current clinical application. Dent Today 27:66, 68, 70.
Thompson JG, Cuff C (1985). Crystal-structure of kaolinite-dimethylsulfoxide intercalate. Clays Clay Miner 33:490-500.
Tian M, Gao Y, Liu Y, Liao Y, Xu R, Hedin NE, et al. (2007). Bis-GMA/
TEGDMA dental composites reinforced with electrospun nylon 6 nanocomposite nanofibers containing highly aligned fibrillar silicate single
crystals. Polymer 48:2720-2728.
Tilbrook DA, Clarke RL, Howle NE, Braden M (2000). Photocurable
epoxy-polyol matrices for use in dental composites I. Biomaterials
21:1743-1753.
Wang YL, Lee BS, Chang KC, Chiu HC, Lin FH, Lin CP (2007).
Characterization, fluoride release and recharge properties of polymerkaolinite nanocomposite resins. Comp Sci Tech 67:3409-3416.
Watanabe H, Khera SC, Vargas MA, Qian F (2008). Fracture toughness
comparison of six resin composites. Dent Mater 24:418-425.
Weinmann W, Thalacker C, Guggenberger R (2005). Siloranes in dental
composites. Dent Mater 21:68-74.
Wilson KS, Zhang K, Antonucci JM (2005). Systematic variation of interfacial phase reactivity in dental nanocomposites. Biomaterials 26:50955103.

J Dent Res 89(6) 2010

Wilson KS, Allen AJ, Washburn NR, Antonucci JM (2007). Interphase


effects in dental nanocomposites investigated by small-angle neutron
scattering. J Biomed Mater Res A 81:113-123.
Xia Y, Zhang F, Xie H, Gu N (2008). Nanoparticle-reinforced resin-based
dental composites. J Dent 36:450-455.
Xu HH, Sun L, Weir MD, Antonucci JM, Yakagi S, Chow LC, et al. (2006).
Nano DCPA-whisker composites with high strength and Ca and PO4
release. J Dent Res 85:722-727.
Xu HH, Weir MD, Sun L, Takagi S, Chow LC (2007a). Effects of calcium
phosphate nanoparticles on Ca-PO4 composite. J Dent Res 86:
378-383.
Xu HH, Weir MD, Sun L (2007b). Nanocomposites with Ca and PO4
release: effects of reinforcement, dicalcium phosphate particle size and
silanization. Dent Mater 23:1482-1491.
Xu HH, Moreau JL, Sun L, Chow LC (2008). Strength and fluoride release
characteristics of a calcium fluoride based dental nanocomposite.
Biomaterials 29:4261-4267.
Xu HH, Weir MD, Sun L (2009). Calcium and phosphate ion releasing
composite: effect of pH on release and mechanical properties. Dent
Mater 25:535-542.
Yap AU, Low JS, Ong LF (2000). Effect of food-simulating liquids on surface characteristics of composite and polyacid-modified composite
restoratives. Oper Dent 25:170-176.
Yesil ZD, Alapati S, Johnston W, Seghi RR (2008). Evaluation of the wear
resistance of new nanocomposite resin restorative materials. J Prosthet
Dent 99:435-443.

Downloaded from jdr.sagepub.com at UNIVERSIDAD DE CHILE on May 10, 2015 For personal use only. No other uses without permission.
2010 International & American Associations for Dental Research

También podría gustarte