You are on page 1of 8

Australian Dental Journal

The official journal of the Australian Dental Association

Australian Dental Journal 2011; 56:(1 Suppl): 5966


doi: 10.1111/j.1834-7819.2010.01296.x

Resin composite restorative materials


N Ilie,* R Hickel*
*Department of Restorative Dentistry, Dental School of the Ludwig-Maximilians-University, Munich, Germany.

ABSTRACT
This paper surveys the most important developments in resin-based dental composites and focuses on the deficits (e.g.
polymerization shrinkage) and strengths of the materials and their clinical implications. Moreover, differences between
composite categories, such as hybrid, nanohybrid, microfilled, packable, ormocer-based, silorane-based, polyacid-modified
composites (compomers) and flowable composites are highlighted, especially in view of their mechanical behaviour. In
addition to the classical dimethacrylate-based composites, special attention is given to alternative monomers, such as
siloranes, ormocers or high-molecular-weight dimethacrylate monomers (e.g. dimer acid-based dimethacrylates and
tricyclodecane (TCD)-urethane), analysing their advantages, behaviour and abilities. Finally, the paper attempts to establish
the needs and wishes of clinicians for further development of resin-based composites.
Keywords: Composites, mechanical properties, strength, modulus of elasticity.
Abbreviations and acronyms: CS = compressive strength; DSC = differential scanning calorimetry; FM = flexural modulus; FS = flexural
strength; NMR = nuclear magnetic resonance; RBC = resin-based composite; TCD = tricyclodecane; UDMA = urethane dimethacrylate.

INTRODUCTION
Choosing a suitable resin-based composite (RBC) for a
restoration in modern dentistry requires balancing a
large number of requirements. This requires functional
properties, including enhanced longevity of the restoratives by excellent mechanical properties such as high
strength, fracture toughness, surface hardness, optimized modulus of elasticity, low wear, low water
sorption and solubility, low polymerization shrinkage,
low fatigue and degradation, high radiopacity and a
better detection of the material during removal of a
composite restoration. At the same time, biological
properties resembling a good biocompatibility (systemic and local), no postoperative pain or hypersensitivity, preservation of tooth integrity in terms of not
causing fractures or cracks, as well as caries-inhibiting
abilities are required. In addition, aesthetic considerations, such as good colour matching and colour
stability (translucency, shades), optimum polishability,
long-term surface gloss, absence of marginal or surface
staining and a good long-term anatomical form should
also be fulfilled.1
To improve many of the above-mentioned properties,
the size of filler particles incorporated in the resin
matrix of commercial dental composites has continuously decreased over the years, from the traditional to
2011 Australian Dental Association

the nano-composite materials.2 Apart from changes in


filler amount, shape or surface treatment, changes in
monomer structure or chemistry and modification of
the dynamics of the polymerization reaction have been
introduced.
Several clinical negative effects in resin-based composite restorations, such as marginal discrepancies,3,4
marginal staining, white lines around the restoration,
cusp fractures,5 microleakage,6 debonding,5 secondary
caries, postoperative sensitivity or pain are still
observed in restorations with modern RBCs. These
are frequently connected to polymerization shrinkage
stress. Even if these assertions are nowadays broadly
accepted, only limited clinical evidence exists to
support a clear relationship between these negative
outcomes and polymerization shrinkage stress.38
Nevertheless, new materials to reduce internal stress
as a result of low polymerization shrinkage are
predicted to dominate the market in the future, even
before a complete understanding of the clinical effect of
shrinkage is achieved.9
In a review article on clinical challenges with
posterior resin-based restorations, Sarrett10 asserted
that clinical data indicate secondary caries and restoration fracture are the most common clinical problems
in posterior composite restorations. Reviewing data
collected from 24 clinical studies published between
59

N Ilie and R Hickel


1996 and 2002 regarding direct posterior restorations
in service periods up to five years, Brunthaler et al.11
showed that composite fracture was the most frequent
type of failure, whereas secondary caries was the most
common failure beyond five years of clinical service.
The predominant reason for failure of composite
restorations placed in larger cavities was found to be
fracture, also for periods longer than 11 years.12,13 This
trend in increased fractures in fillings placed in loadbearing areas, compared to the decades before, was
certainly caused by the extension of the indication for
composites to larger multi-surface posterior approximal
cavities. However, the compromise necessary to create
a universal composite by decreasing the filler size in
order to improve aesthetics could also have led to
decreased mechanical properties.
New monomers for resin-based composites
The changes in monomer chemistry were at first
directed to improve the already clinical successful
methacrylate-based systems by modifying the Bowen
monomer (Bis-GMA: 2,2-bis[4-(2-hydroxy-3-methacrylyloxypropoxy)phenyl] propane) and the UDMA
(urethane dimethacrylate) developed by Foster and
Walker14 to create monomers with lower viscosity,1517
such as hydroxyl-free Bis-GMA,18 aliphatic urethane
dimethacrylates (UEDMA: 1,6-bis[methacrylyloxy-2ethoxycarbonylamino]-2,4,4-trimethylhexane),19 partially aromatic urethane dimethacrylate20 or highly
branched methacrylates.21 Also, ring-opening monomers such as spiro-orthocarbonates for non- or
minimally-shrinking dental composites as additives to
dimethacrylates,22 and epoxy-base resins like the siloranes,23 as well as a series of high-molecular-weight
monomers like dimer acid-based dimethacrylates,24
tricyclodecane (TCD) urethane and organically-modified ceramics (ormocers)25,26 were introduced to the
market for the same purpose.
Modification of the dynamics of the polymerization
reaction by slowing down the polymerization rate is a
further mechanism to compensate stress in RBCs,
thereby increasing the material flow capacity and being
associated with lower stress build-up and better interfacial integrity.27,28 A recently introduced flowable
resin-based composite material, intended to be used as a
liner in occlusal and posterior approximal restorations
(SureFil SDRTM Flow; SDRTM Posterior Bulk Fill
Flowable respectively; Dentsply, York, PA, USA),
includes a monomer having incorporated a photoactive
group in a urethane-based methacrylate resin, showing
a 6070% reduction in shrinkage stress in the unfilled
resin when compared to conventional methacrylatebased resins.29 The activated resin demonstrated a
relatively slow radical polymerization rate, suggesting
that the photoinitiator incorporated into the resin is
60

affecting the radical polymerization process. This lower


curing stress was also shown to be retained in filled
compositions, particularly in cases of low fillerloading.29 These resins are patented as being based on
Stress Decreasing Resin (SDRTM) technology.30 Incorporated in the commercial material SureFil SDRTM
Flow, it was shown that shrinkage stress after polymerization was lower not only when compared to
conventional flowable RBCs (EsthetX Flow; Dentsply
and Filtek Supreme Plus Flow; 3M-ESPE, St Paul, MN,
USA), but also when compared to nano- and microhybrid RBCs (EsthetX Plus; Dentsply and Filtek Supreme
Plus, 3M-ESPE) or even with a silorane-based RBC
(Filtek Silorane, 3M-ESPE). The material showed the
lowest rate of shrinkage stress, but was also more rigid
(i.e. higher modulus of elasticity) and more plastic (low
elastic deformation, high plastic deformation and high
creep values) as the regular flowable materials, therefore making its effect on interfacial stress build-up
difficult to predict.59
Dimer acid-based monomers (Fig 1a)
The idea behind the synthesis of dimer acid-based
monomers was to reduce volume shrinkage during
polymerization by using high-molecular-weight monomers with decreased initial double-bond concentrations. The monomers are derived from a core structure
based on hydrogenated dimer acid, a derivative of
linoleic acid, which is an essential fatty acid and a
major constituent of many vegetable oils, such as
soybean, peanut and cottonseed oil.24
Compared to conventional dimethacrylate monomers
like bis-GMA or UDMA, the dimer acid dimethacrylate
monomers showed higher molecular weights and lower
initial double bond concentrations with relatively low
viscosities.24 A higher degree of conversion, a lower
polymerization shrinkage and water sorption values in
comparison with conventional monomers were also
measured. In addition, the relatively low cross-link
density of dimethacrylates constructed from dimer acid
was shown to produce polymers with high flexibility
but low modulus of elasticity.24
The commercial nanohybrid RBC based on this
dimer acid derivate (NDurance, Septodont, France)
has included in the organic matrix also BisGMA,
UDMA and dicarbamate, as well as fillers based on
ytterbium fluoride, barium glass and quartz (80 wt%,
65 vol%). The composite showed in comparison to
other modern nanohybrid RBCs e.g. Venus Diamond (Heraeus Kulzer, Hanau, Germany), Grandio
(VOCO, Cuxhaven, Germany), Miris 2 (Colte`ne Whaledent, Altstatten, Switzerland), Simile (Jeneric Pentron, Wallingford, CT, USA) and Premise
(Kerr, Orange, CA, USA) good flexural strength and
modulus of elasticity values, being significantly infe 2011 Australian Dental Association

Composites
(a)

(c)

(b)

(d)

(e)

Fig 1. (a) Structures of the dimethacrylate monomers based on dimer acid as described by Trujillo-Lemon et al.24 (b) TCD-Urethan32 as described in
US 2010 0076115 A1, Utterodt et al., United States Patent Application Publication, 23 March 2010. (c) DuPont Monomer as described in http://
www.kalore.net/index2.php. (d) Silorane monomer, as described by Weinmann et al.23 (e) Structure of dimethacrylate ormocers as described by
Moszner et al.26

rior only to Venus Diamond and Grandio (publication


in preparation).
TCD-urethane based monomers (Fig 1b)
TCD-urethane monomers were incorporated into
recently introduced composite materials (Venus Diamond and Venus Diamond Flow). According to the
manufacturer, the new methacrylic acid derivatives,
containing urethane groups of tricyclodecanes are
prepared by reaction of hydroxyalkyl (meth)acrylic
acid esters with diisocyanates and subsequent reaction
with polyols.31 Similar to bisphenol-A, the structure of
the TCD-urethane backbone was proven to be rigid,
resulting in low shrinkage polymerization behaviour.
As a result of the steric restriction of the mobility, the
urethane derivatives of 1,3-bis(1-isocyanato-1-methylethyl) benzene are very similar in terms of their
properties to bis-GMA, and can be used in dental
composites in its place, as described in the patent EP
0934926.32 TCD-urethane is a low-viscosity monomer
that, according to the manufacturer, would dispense
with the use of diluents responsible for the high
polymerization shrinkage of BisGMA-based composites. It was shown that RBCs based on TCD-urethane
2011 Australian Dental Association

(Venus Diamond in the experimental version NEUN)


have lower shrinkage and polymerization stress compared to others containing conventional dimethacrylates.33
Further, in the new monomer, the high reactivity of
an acrylic acid ester containing urethane groups was
combined with the rigid structure of the TCD skeleton,
thus higher degrees of conversion with an increased
reactivity are achieved (United States Patent Application 20080167399, publication in preparation). Toxicological tests showed a high biocompatibility of the
polymerized composite (United States Patent Application 20080167399).
New monomer technology from DuPont
Reducing shrinkage through the implementation of
monomers with a higher molecular weight was also the
idea behind the synthesis of a new monomer (Fig 1c)
from DuPont (USA) as a matrix for the composite
Kalore (GC). The new methacrylate monomer consists
of a long rigid core with flexible side arms and a lower
number of C=C double bonds. Until recently, there has
been little information regarding the performance of the
new material. Recent conference presentations confirm
61

N Ilie and R Hickel

180
160
140
120
100
80
60
40
20
0

E-micro (GPa)

Flexural strength (MPa)

Modern developments in dental composite research


have focused on the use of ring-opening systems like
oxirane-based resins cured under visible light conditions. Oxirane resins have shown many desirable
properties, such as improved depth of cure, lower
polymerization shrinkage, higher strength, as well as
equivalent hardness and acceptable glass transition
temperature when compared with conventional
bis-GMA-based dental resins.37,38 Residual monomers additives released from commercial methacrylate-based composite materials after polymerization are
found to be toxic.39,40 The in vitro cytotoxicity and
mutagenicity of oxirane resins are also presumed.41 As
a result, research continues in the direction of ringopening monomers, with intense effort and focus on
reducing the effects of cytotoxicity and mutagenicity.
Weinmann et al.23 described the synthesis of a new
monomer system, a silorane (Fig 1d) obtained from
the reaction of oxirane and siloxane molecules. The
novel resin claimed to have combined the two key
advantages of the individual components: low polymerization shrinkage due to the ring-opening oxirane
monomer and increased hydrophobicity due to the
presence of the siloxane species. Further, Schweikl et
al.42 showed that the mutagenic potential of various
siloranes tested in diverse test systems was much lower
than those of related oxiranes, while Eick et al.43 found

20
18
16
14
12
10
8
6
4
2
0

Ormocers
In an attempt to overcome the problems created by the
polymerization shrinkage of conventional composites,
the organically-modified ceramics (ormocers; Fig 1e)
as a new material class were developed.25 Having a very
similar coefficient of thermal expansion to natural
tooth structure, the materials were formulated as a
novel three-dimensional cross-linked inorganic-organic
polymer, synthesized from multifunctional urethaneand thioether(meth)acrylate alkoxysilanes as sol-gel
precursors. Alkoxysilyl groups of the silane permit the
formation of an inorganic Si-O-Si network by hydrolysis and polycondensation reactions. The methacrylate
groups are available for photochemical polymerization.
Flexural modulus (GPa)

Silorane-based monomers

that siloranes were stable and insoluble in biological


fluids simulated using aqueous solutions containing
either epoxide hydrolase, porcine liver esterase or dilute
HCl. All these reported advantageous characteristics
serve to enhance the potential of silorane monomers to
be used successfully in dental composite materials. The
low shrinkage of siloranes and the toxicological safety
of the otherwise critical epoxides in cured dental
composites have been also described in DE 10001228
and EP 1117368.
The macromechanical (flexural strength and modulus
of elasticity from a flexural test) and micromechanical
performance (Vickers hardness and modulus of elasticity measured in a universal hardness test) of the above
mentioned four RBCs are presented in Fig 2. The mean
values of the material categories nanohybrid, microhybrid and flowable composites were taken as references
(publication in preparation).

Vickers hardness (MPa)

a lower shrinkage stress compared to regular RBCs,34,35


but also a higher water sorption and radial expansion
when compared to the low shrinkage silorane-based
composite (Filtek Silorane; 3M-ESPE).36

10
9
8
7
6
5
4
3
2
1
0

140
120
100
80
60
40
20
0

Fig 2. Macromechanical (exural strength and modulus of elasticity in exural test) and micromechanical performance (Vickers hardness and
modulus of elasticity measured in a Universal hardness test E-micro). The mean values of the material category nanohybrid, microhybrid and
owable composites were taken as references (publication in preparation).
62

2011 Australian Dental Association

Composites
These materials proved to have lower wear rates
compared with composites,44,45 and a shrinkage equal
to that of hybrid composites, despite having less filler
content.46 However, due to problems with upscaling of
prototypes and handling properties, conventional methacrylate had to be added to the ormocer matrix of the
first commercial products, diminishing the initial promising advantages. As a material group, with the
limitation of the low amount of tested ormocer
composites, their mechanical properties do not differ
from the hybrid composites. However, newly developed
but still experimental ormocers, synthesized from
amine or amide dimethacrylate trialkoxysilanes and
being dimethacrylate-diluent free, showed a clearly
improved flexural modulus of elasticity compared to
the ormocer-based composite Definite (then Degussa
GmbH, now Evonik GmbH, Essen, Germany),26 promising a revival of this material category.
Fillers
The size of filler particles incorporated in the resin
matrix of commercial dental composites has continuously decreased over the years from the traditional to
the nano-composite materials.2 Filler not only directly
determines the mechanical properties of composite
materials but also allows reduction in the monomer
content and consequently the polymerization shrinkage, optimizing wear, translucency, opalescence, radiopacity, intrinsic surface roughness and thus polishability, as well as enhancing aesthetics and improving
handling properties. The smaller the particles, the better
the polish and gloss, but the reduction of filler size and
subsequent increase in surface area to volume ratio has
limited the achievable filler loading, resulting in
decreased handling and mechanical properties.2
In an attempt to have a clear clinical indication for the
existing commercial composites, classification criteria
were developed, most of them influenced by the filler
system.4750 These criteria are primarily based on the
amount of inorganic filler fraction in volume per cent or
on the mean particle size. In addition, the Youngs
modulus of elasticity and the intrinsic surface roughness
were taken into consideration as valuable classification
parameters49 in view of the crucial role of the modulus of
elasticity in the deformability of a material under
masticatory stresses, particularly in the posterior region,
or considering the fact that for aesthetic demands a low
intrinsic surface roughness is necessary.
A valuable overview in the development and diversity
of composites offers the conventional classification
system introduced by Lutz and Phillips,48 based on
the specific filler-size distribution and amount of
incorporated filler, which divides the composites into
traditional, hybrids containing a mixture of ground
glass and microfill particles, and microfill composites.
2011 Australian Dental Association

The microfills are further divided into subclasses


including a characterization of the type of prepolymerized resin fillers incorporated, i.e. splintered,
agglomerated, or spherical, since not only the filler size
and amount but also the shape consistently influence
the composites properties. Whereas spherical fillers can
be incorporated in a higher amount in a composite than
in irregular fillers of the same size, spherical fillers result
in a higher wear rate.51 The generally accepted view
nowadays is that microfill composites have the most
ideal aesthetic qualities due to their excellent polishability and capacity to retain surface smoothness over
time. However, it is also accepted that, due to their
poor mechanical properties, these materials are contraindicated for stress-bearing restorations, such as incisal
edges and moderate-to-large stress-bearing restorations
in occlusal contact with opposing cusps.2
A more recent classification system49 based on the
filler volume fracture and the filler size, distinguishes
between densified composites, microfine composites,
miscellaneous composites, traditional composites, and
fibre-reinforced composites. Considering Youngs modulus value of dentine, an imaginary filler volume
percentage of 60% was calculated as a correspondent
to a filled composite. Since composites intended for
posterior use should have a Youngs modulus at least
equal to that of dentine, the densified composites were
subdivided into two classes, midway-filled (<60 vol%)
and compact-filled (>60 vol%) with classifications of
ultrafine (<3 lm) and fine (>3 lm) within each category
as a function of the mean particle size of the filler. A
material with a low Youngs modulus, particularly
placed in posterior regions, will result in a higher
deformation under masticatory stresses, potentially
having more catastrophic failures as a consequence.
The single term hybrid is no longer used in this
classification as nearly all dental composites are now
hybrids with particles of two or more size ranges
because they contain some amorphous silica to improve
handling by reducing stickiness.
A much simpler classification system distinguished
between three types of composites classified by the size
of their largest fillers as microfills (mean particle size =
0.010.1 lm), minifills (mean particle size = 0.11.0
lm), and midifills (mean particle size = 1.010.0 lm).47
Another classification based only on the filler system
was proposed by an FDI Science Commission Project in
1998.50
The variety and amount of commercial composite
materials for tooth restoration available today makes it
difficult to compare their efficiency in clinical studies.
Thus, for comparative investigations of material behaviour, it is valuable to evaluate data measured under
identical test conditions. An analysis of 72 commercial
materials belonging to eight different material categories according to their mechanical behaviour (contain63

N Ilie and R Hickel


ing the material groups nano-composites, hybrid,
microfilled, packable, ormocer-based and flowable
composites, compomers and flowable compomers),
showed large varieties between the tested materials
within the same material category. The mechanical
properties analysed were flexural strength (FS), flexural
modulus (FM), diametric tensile strength (DTS) and
compressive strength (CS), and were measured after the
samples had been stored in water for 24 hours at 37 C.
The groups of hybrid, nanohybrid, packable and
ormocer-based composites did not differ significantly
among each other as a material type, reaching the
highest flexural strength values. Nanohybrid composites were characterized by a good flexural strength, the
best diametral tensile strength, but a low flexural
modulus.52 Furthermore, the above mentioned data
showed that the filler volume has the highest influence
on the mechanical properties. Compared to hybrid
composites, the filler weight and volume of the
nanohybrids were similar. Considering the variation
of the FS and FM as a function of the filler volume for
all materials (without taking into account the material
category), a trend to enhance the mechanical properties
until a filler volume of ca. 60% was observed by
contouring the areal density of the measured data with
an Epanechnikov kernel function. This showed that
introducing a volume of filler higher than 60% will
probably also introduce a higher amount of defects,
leading to a decrease in mechanical properties. These
macroscopical data are only showing a general trend in
modern resin-based composites, demonstrating however, that the volume amount of filler also in optimized
resin-based composites is restricted to a maximum
value. Moreover, the influence of the type of material
on the mechanical properties was significant but very
low, showing the strongest influence on the CS.52
Developing nanocomposite materials in dentistry is
not just another step toward miniaturization of fillers
since the physical and chemical properties of nanomaterials often exhibit important differences from the
bulk properties that cannot be fully explained by
current theories. Enlarging the surface area to volume
ratio of the fillers in nano-filled materials tends to result
in increased water uptake and therefore a potential
degradation of the filler/matrix interface. This negatively affects the mechanical properties of nano-filled
materials, when compared to microhybrid composites
(Example: nanocomposite Filtek Supreme vs. microhybrid filtek Z250; 3M-ESPE).53
It was shown that even if nanocomposites perform
better or comparable to microhybrid composites in
regards to flexural strength, their behaviour by ageing
or storage in aggressive conditions is different. Comparing nanohybrid RBCs (Tetric EvoCeram (Ivoclar
Vivadent; Liechtenstein) and Filtek Supreme XT) with
microhybrid RBCs (Tetric, Tetric Ceram, Tetric Ceram
64

HB (Ivoclar), Filtek Silorane (3M ESPE), EsthetX


(Dentsply)), the Weibull analysis of flexural strength
distinguished between three different behaviour patterns: a decrease in strength after storing for four weeks
in water or saliva and an even more pronounced
decrease by storing in alcohol (behaviour characteristic
for the nanohybrid RBCs: Tetric EvoCeram and Filtek
Supreme XT); a decrease in strength only after alcohol
storage (Tetric, Tetric Ceram, Tetric Ceram HB); and
no influence from storage conditions (Filtek Silorane,
EsthetX). The last two materials were different with
respect to the chemical composition of the matrix
(silorane matrix and methacrylate-matrix without
UDMA), and were also the two less filler-loaded
composite materials. The materials measured in this
study and stored for four weeks had additionally been
thermocycled for 5000 cycles at 5 to 55 C.54
Remarkable for Tetric EvoCeram was that it showed
a lower flexural strength than Filtek Supreme, but the
Weibull modulus m was higher, indicating a lower
degree of dispersion of the strength data or, in other
words, showing that Tetric EvoCeram was the more
reliable material under measurement conditions. However, it was also demonstrated that the nanoclusters in
Filtek Supreme provide a distinct fracture mechanism,55
resulting in significant improvements in strength and
reliability when subjected to cyclic pre-loading56 which
could improve fatigue and thus enhance the clinical
longevity of the material.
Aged Filtek Supreme Plus (37 C for five months in
air, artificial saliva, water or a 50:50 by volume mixture
of ethanol and water) showed increased failure during
cyclic fatigue loading and significantly decreased fracture toughness when compared to a hybrid composite
(Renew; Bisco Inc, Schaumberg, IL, USA). The nonaged groups of the two materials showed a similar
performance.57 The increased susceptibility of the
nano-filler RBCs to ageing in this study is also
explained by the increased intersurface area between
ultra-small filler particles and resin matrix being one of
the weakest links in a dental resin composite. By
comparing the two material classes, microhybrid and
nanohybrid dental composites, no significant differences between the filler weight and volume were
found.52 A continuous decrease in particle dimensions
in nanohybrid composites at a constant filler volume
has, as a consequence, larger filler surfaces relative to
the volume size, resulting in a large interface between
filler and organic matrix. The interaction of these
phases at the interfaces also determines special properties in the material. This large specific surface forms a
large boundary surface with the surrounding matrix,
therefore changing not only the morphology of the
composite but also influencing factors, such as the
mobility of the macromolecules at the particle surfaces
and polymerization kinetics. At the same time, the
2011 Australian Dental Association

Composites
distance between reinforcing particles is consistently
decreasing with decreased filler size, having a considerable influence on the mechanical properties of the
organic matrix between the particles.
Besides the reduced interparticle distance in composites with small filler dimensions, an additional phenomenon changes the properties of the materials. The
adhesion of the macromolecules of the matrix to the
filler particles can form an immobile polymer boundary
layer. Generally, this disturbed polymer layer around
the filler particles was shown to have in nanocomposites a thickness varying between some nanometres to
20 nm. However, this is difficult to measure in low
filled materials and the result is strongly dependent on
the measuring procedures (e.g. differential scanning
calorimetry (DSC), nuclear magnetic resonance (NMR)
and neutron roaming). This boundary layer or zone of
influence will have an important effect on the mechanical characteristics of RBCs and might affect clinical
performance as smaller particles, and therefore less
interparticle spacing, will better protect the surrounding polymer matrix.58
Comparing the mechanical properties of more
recently introduced nanohybrid and microhybrid RBCs
(Fig 2) showed that the former had significantly superior properties, encouraging their wide clinical use
(publication in preparation).
REFERENCES
1. Hickel R, Roulet JF, Bayne S, et al. Recommendations for conducting controlled clinical studies of dental restorative materials.
Science Committee Project 2 98FDI World Dental Federation
study design (Part I) and criteria for evaluation (Part II) of direct
and indirect restorations including onlays and partial crowns. J
Adhes Dent 2007;9:121147.
2. Ferracane JL. Current trends in dental composites. Crit Rev Oral
Biol Med 1995;6:302318.
3. Irie M, Suzuki K, Watts DC. Marginal gap formation of lightactivated restorative materials: effects of immediate setting
shrinkage and bond strength. Dent Mater 2002;18:203210.
4. Sakaguchi RL, Peters MC, Nelson SR, Douglas WH, Poort HW.
Effects of polymerization contraction in composite restorations. J
Dent 1992;20:178182.
5. Ferracane JL. Buonocore Lecture. Placing dental compositesa
stressful experience. Oper Dent 2008;33:247257.
6. Ferracane JL, Mitchem JC. Relationship between composite
contraction stress and leakage in Class V cavities. Am J Dent
2003;16:239243.
7. Loguercio AD, Reis A, Schroeder M, Balducci I, Versluis A,
Ballester RY. Polymerization shrinkage: effects of boundary
conditions and filling technique of resin composite restorations. J
Dent 2004;32:459470.

11. Brunthaler A, Konig F, Lucas T, Sperr W, Schedle A. Longevity of


direct resin composite restorations in posterior teeth. Clin Oral
Investig 2003;7:6370.
12. van Dijken JW. Direct resin composite inlays onlays: an 11-year
follow-up. J Dent 2000;28:299306.
13. Van Nieuwenhuysen JP, DHoore W, Carvalho J, Qvist V. Longterm evaluation of extensive restorations in permanent teeth. J
Dent 2003;31:395405.
14. Foster J, Walker R. US Patent No. 3825518, 1974.
15. Peutzfeldt A. Resin composites in dentistry: the monomer systems. Eur J Oral Sci 1997;105:97116.
16. Moszner N. New developments of polymer dental composites.
Prog Polym Sci 2001;26:535576.
17. Moszner N. Recent developments of new components for dental
adhesives and composites. Macromol Mater Eng 2007;292:245
271.
18. Patent EfPPG. New acrylates and methacrylates. UK Patent No.
1263541, 1972.
19. Patent ICIL. Dental compositions. England Patent No. 1465897,
1977.
20. Moszner N, Fischer UK, Angermann J, Rheinberger V. A partially
aromatic urethane dimethacrylate as a new substitute for BisGMA in restorative composites. Dent Mater 2008;24:694699.
21. Klee JE, Neidhart F, Flammersheim H-J, Mulhaupt R. Monomers
for low shrinking composites. 2a: Synthesis of branched methacrylates and their application in dental composites. Macromol
Chem Phys 1999;200:517523.
22. Eick JD, Robinson SJ, Byerley TJ, Chappelow CC. Adhesives and
nonshrinking dental resins of the future. Quintessence Int
1993;24:632640.
23. Weinmann W, Thalacker C, Guggenberger R. Siloranes in dental
composites. Dent Mater 2005;21:6874.
24. Trujillo-Lemon M, Ge J, Lu H, Tanaka J, Stansbury J. Dimethacrylate derivates of dimer acid. J Polymer Science
2006;44:39213929.
25. Wolter H, Storch W, Ott H. New inorganic organic copolymers
(ORMOCERS) for dental applications. Mat Res Soc Symp Proc
1994;143149.
26. Moszner N, Gianasmidis A, Klapdohr S, Fischer UK, Rheinberger
V. Sol-gel materials 2. Light-curing dental composites based on
ormocers of cross-linking alkoxysilane methacrylates and further
nano-components. Dent Mater 2008;24:851856.
27. Feilzer AJ, De Gee AJ, Davidson CL. Quantitative determination
of stress reduction by flow in composite restorations. Dent Mater
1990;6:167171.
28. Feilzer AJ, de Gee AJ, Davidson CL. Setting stresses in composites
for two different curing modes. Dent Mater 1993;9:25.
29. Jin X, Bertrand S, Hammesfahr PD. New radically polymerizable
resins with remarkably low curing stress. J Dent Res
2009;88(Spec Iss A):1651.
30. Jin Xea. US Patent Application No. US2008 0076853.
31. Reiners J, Podszun W, Winkel J. (Meth)acrylic acid derivatives,
containing urethane groups, of tricyclo[5.2.1.02.6]decanes.
European Patent EP0254185, 1990. Assignee: Bayer AG.
32. Utterodt Aea. Dental composites with tricyclo[5.2.02.6]decane
derivatives. European Patent EP1935393, 2008. Assignee: Heraeus Kulzer GmbH.

8. van Dijken JW. Durability of resin composite restorations in high


C-factor cavities: a 12-year follow-up. J Dent 2010;38:469474.

33. Kurokawa R, Finger WJ, Hoffmann M, et al. Interactions of selfetch adhesives with resin composites. J Dent 2007;35:923929.

9. Ferracane JL. Developing a more complete understanding of


stresses produced in dental composites during polymerization.
Dent Mater 2005;21:3642.

34. Maseki T, Nitta T, Yamase M, et al. Characteristics in polymerization shrinkage of latest low-shrinkage resin composite restoratives. J Dent Res 2010;89: Spec Iss Letter A. Abstract No. 457.

10. Sarrett DC. Clinical challenges and the relevance of materials


testing for posterior composite restorations. Dent Mater
2005;21:920.

35. Platt J, Macpherson M, Rhodes B. Polymerization shrinkage and


contraction stress of an experimental composite. J Dent Res
2010;89: Spec Iss Letter A. Abstract No. 1024.

2011 Australian Dental Association

65

N Ilie and R Hickel


36. Wei Y, Silikas N, Watts DC. Water sorption and hygroscopic
expansion in new dental resin composites. J Dent Res 2010;89:
Spec Iss Letter B. Abstract No. 2307.

51. Venhoven BA, de Gee AJ, Werner A, Davidson CL. Influence of


filler parameters on the mechanical coherence of dental restorative resin composites. Biomaterials 1996;17:735740.

37. Tilbrook DA, Clarke RL, Howle NE, Braden M. Photocurable


epoxy-polyol matrices for use in dental composites I. Biomaterials
2000;21:17431753.

52. Ilie N, Hickel R. Investigations on mechanical behaviour of


dental composites. Clin Oral Investig 2009;13:427438.

38. Eick JD, Kostoryz EL, Rozzi SM, et al. In vitro biocompatibility
of oxirane polyol dental composites with promising physical
properties. Dent Mater 2002;18:413421.
39. Reichl FX, Walther UI, Durner J, et al. Cytotoxicity of dental
composite components and mercury compounds in lung cells.
Dent Mater 2001;17:95101.
40. Bettencourt AF, Neves CB, de Almeida MS, et al. Biodegradation
of acrylic based resins: a review. Dent Mater 2010;26:e171180.
41. Schweikl H, Schmalz G, Weinmann W. Mutagenic activity of
structurally related oxiranes and siloranes in Salmonella typhimurium. Mutat Res 2002;521:1927.
42. Schweikl H, Schmalz G, Weinmann W. The induction of gene
mutations and micronuclei by oxiranes and siloranes in mammalian cells in vitro. J Dent Res 2004;83:1721.
43. Eick JD, Smith RE, Pinzino CS, Kostoryz EL. Stability of
silorane dental monomers in aqueous systems. J Dent 2006;
34:405410.
44. Tagtekin DA, Yanikoglu FC, Bozkurt FO, Kologlu B, Sur H.
Selected characteristics of an Ormocer and a conventional hybrid
resin composite. Dent Mater 2004;20:487497.
45. Tagtekin D, Tan C, Chung S. Wear behavior of new composite
restoratives. Oper Dent 2004;29:269274.
46. Cattani-Lorente M, Bouillaguet S, Godin CH, Meyer JM. Polymerization shrinkage of Ormocer based dental restorative composites. Eur Cell Mater 2001;1:2526.
47. Bayne SC, Heymann HO, Swift EJ Jr. Update on dental composite restorations. J Am Dent Assoc 1994;125:687701.
48. Lutz F, Phillips RW. A classification and evaluation of composite
resin systems. J Prosthet Dent 1983;50:480488.
49. Willems G, Lambrechts P, Braem M, Celis JP, Vanherle G. A
classification of dental composites according to their morphological and mechanical characteristics. Dent Mater 1992;8:310
319.

53. Curtis AR, Shortall AC, Marquis PM, Palin WM. Water uptake
and strength characteristics of a nanofilled resin-based composite.
J Dent 2008;36:186193.
54. Ilie N, Hickel R. Macro-, micro- and nano-mechanical investigations on silorane and methacrylate-based composites. Dent
Mater 2009;25:810819.
55. Curtis AR, Palin WM, Fleming GJ, Shortall AC, Marquis PM.
The mechanical properties of nanofilled resin-based composites: characterizing discrete filler particles and agglomerates
using a micromanipulation technique. Dent Mater 2009;25:
180187.
56. Curtis AR, Palin WM, Fleming GJ, Shortall AC, Marquis PM.
The mechanical properties of nanofilled resin-based composites:
the impact of dry and wet cyclic pre-loading on bi-axial flexure
strength. Dent Mater 2009;25:188197.
57. Lin L, Drummond JL. Cyclic loading of notched dental composite
specimens. Dent Mater 2010;26:207214.
58. Kahler B, Kotousov A, Swain MV. On the design of dental resinbased composites: a micromechanical approach. Acta Biomater
2008;4:165172.
59. Investigations on a methacrylate-based flowable composite based
on the SDRTM technology. Dent Mater 2010 Dec 29 [Epub ahead
of print].

Address for correspondence:


Dr Nicoleta Ilie
Department of Restorative Dentistry
Dental School of the Ludwig-Maximilians-University
Goethestr. 70
80336 Munich
Germany
Email: nicoleta.ilie@dent.med.uni-muenchen.de

50. Hickel R, Dasch W, Janda R, Tyas M, Anusavice K. New direct


restorative materials. FDI Commission Project. Int Dent J
1998;48:316.

66

2011 Australian Dental Association

You might also like