Você está na página 1de 12

Dental Materials (2005) 21, 5667

www.intl.elsevierhealth.com/journals/dema

Conversion-dependent shrinkage stress and strain


in dental resins and composites
Jeffrey W. Stansburya,b,*, Marianela Trujillo-Lemona, Hui Lua,
Xingzhe Dinga, Yan Linb, Junhao Geb
a

Department of Restorative Dentistry, University of Colorado Health Sciences Center, Aurora, CO, USA
Department of Chemical and Biological Engineering, University of Colorado, Boulder, CO, USA

Received 26 October 2004

KEYWORDS
Conversion;
Kinetics;
Photopolymerization;
Polymerization
shrinkage;
Shrinkage strain;
Shrinkage stress

Summary The placement of dental composites is complicated by the contraction


that accompanies polymerization of these materials. The resulting shrinkage stress
that develops during cure of a bonded restoration can induce defects within the
composite, the tooth or at the interface resulting in compromized clinical
performance and/or esthetics. In light of the substantial efforts devoted to
understanding and attempting to control shrinkage stress and strain in dental
composite restoratives, this paper offers a perspective on the conversion dependent
development of shrinkage and stress. The relationships between polymer property
development and the physical evolution of the network structures associated with
dental polymers as well as the interrelated kinetics of the photopolymerization
reaction process are examined here. Some of the methods used to assess conversion
in dental resins and composites are considered. In particular, newly introduced
techniques that allow real time analysis of conversion by near-infrared spectroscopy
to be coupled directly to simultaneous dynamic measurements of either shrinkage
stress or strain are described. The results are compared with reports from the dental
materials literature as well as complementary studies in other related fields of
polymer science. The complex, nonlinear correlation between conversion, shrinkage
and stress are highlighted. A brief review of some of the materials-based approaches
designed to minimize polymerization shrinkage and stress is also provided.
Q 2004 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

Introduction
* Corresponding author. Address: Biomaterials Research Center, University of Colorado Health Sciences Center, P.O. Box
6508/Mail Stop F-436, Aurora, CO 80045-0508, USA. Tel.: C1 303
724 1044; fax: C1 720-859-4110.
E-mail address: jeffrey.stansbury@uchsc.edu
(J.W. Stansbury).

Continued improvements in composite restoratives,


which includes advances in dentin bonding, have
led to the widespread clinical acceptance of these
materials by dental practitioners. Coupled with
this, the increasing demand by patients for esthetic
restoration of tooth structure has maintained

0109-5641/$ - see front matter Q 2004 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.dental.2004.10.006

Shrinkage stress and strain in composites


steady growth in the utilization of these polymerbased materials [1]. Composite materials are well
suited for repair of damaged or decayed tooth
structure because in addition to an esthetic
appearance, they can easily be adapted to a wide
variety of direct placement applications and be
bonded chemically to the tooth. While liquid resins
can be converted to solid, crosslinked polymers
very conveniently by simple exposure to visible light
activation, the polymerization process involving
conventional dimethacrylate monomers used in
dental composites is quite complex and the final
polymers are not without intrinsic problems.
Clinically, there are concerns raised if the photocurable composite materials are under-cured
through limited irradiance levels or exposure
times as well as any potential incompatibility
between a photo-initiator system and the spectral
output of the curing light employed. Composite
brand and shade, cavity preparation geometry and
composite layer thickness also affect how efficiently a given material can be photopolymerized.
If inadequate levels of conversion are achieved in
the polymerization, mechanical properties and
wear performance can be compromized. With
incomplete cure, leachable residual monomer and
initiator become greater biocompatibility issues
and color stability may decline as well. However, if
conversion is maximized to reduce these difficulties, then alternative problems of polymerization
shrinkage and brittle fracture of the composite
become more critical.
There appears to be broad consensus that the
decrease in volume that accompanies the polymerization of dental composites is one of the primary
deficiencies that complicates the use of this very
versatile class of restorative materials and possibly
limits service life expectations. Bulk shrinkage in
vinyl addition polymerizations is an unavoidable
result of the formation of new covalent bonds that
bring monomer units closer together and reduce
their mobility as part of an extended polymeric
structure. Unlike many other polymer applications
where the actual shrinkage strain is of predominant
concern, the main problem associated with shrinkage in dental composite restoratives is the shrinkage stress that is developed during polymerization
[2,3]. When the inherent polymerization shrinkage
is frustrated by sufficient interfacial adhesion
between the developing polymer and a non-freely
compliant substrate, as is the case with chemically
bonded dental restorations, stress is conveyed to
the substrate. If the level of stress developed
exceeds either the adhesive or cohesive strength of
any component of the system, a micro- or macrodefect can result. Thus, in spite of surface

57
treatments that provide improved adhesion of
composites to dentin, as well as multi-step composite layering techniques, reliable adhesion without
marginal gap formation has proven elusive [4,5].
Therefore, a substantial amount of research and
product development effort in dental materials has
been directed toward understanding and potentially reducing the development of shrinkage strain
and stress in composites as a means to further
improve these materials.
The purpose of this paper is to provide a
correlation between the conversion-based development of the polymeric network and the evolution
of the properties of shrinkage stress and strain.
Complementing this examination of the chemical
basis of shrinkage stress and strain, will be coverage
of some new materials approaches designed to
minimize these factors in practical dental composites. This report represents a review of concepts
and research related to shrinkage and stress
development in dental polymers as well as an
overview of recent studies conducted at the
University of Colorado School of Dentistry on
these topics.

Resin composition and the


polymerization process
The focus of this paper is mainly on the resins
composed of comonomer mixtures of typical
dimethacrylates. In most dental resins, a low
viscosity diluent monomer, commonly triethylene
glycol dimethacrylate (TEGDMA), is added to a
bulkier and structurally rigid base monomer, such
as bisphenol A glycidyl methacrylate (Bis-GMA) or
urethane dimethacrylate (UDMA). The higher molecular weights or more precisely, the lower
reactive group concentrations, associated with
these base monomers contribute to reduced polymerization shrinkage compared with the smaller
diluent comonomer [6,7]. The overall shrinkage
that occurs during polymerization is proportional to
the degree of conversion [8]. Thus, it is not
particularly meaningful to report shrinkage values
of different materials, or even values for the same
material polymerized under different curing modes,
in the absence of a coordinated measurement of
conversion, ideally on the same specimen used in
the determination of shrinkage.
A better understanding of the various stages of
polymerization can facilitate an appreciation for
how shrinkage stress and strain progress as a
consequence of the reaction process. This approach
can provide improved insights into new methods

58
and materials approaches that can potentially
provide more efficient control over composite
shrinkage and stress. Extensive studies have been
conducted to elucidate the reaction kinetics of
these network polymers [919]. As the extent of
polymerization advances, it passes through several
interrelated physical and kinetic landmarks, including the gel point, auto-acceleration leading to a
rate maximum, vitrification and residual unsaturation in the final polymer. All these features can be
expected to impact the development of physical
and mechanical properties. It is also important to
recognize that these highly crosslinked polymeric
networks are quite heterogeneous in nature [20].
Common measurements of polymer modulus, such
as that obtained by the widely used flexural
strength test, provide results that are averaged
across the material. Use of dynamic thermomechanical or dielectric analyses can effectively portray
the range of modulus that represents the variety of
micro-environments present in the polymer. As
such, it may be misleading to characterize dental
polymers by a single glass transition temperature
without acknowledging that thermal transitions
occur over a very broad range of temperatures
extending to below room temperature [2123].
Therefore, even in fully cured glassy polymers,
there are regions where significant mobility
persists.

Measurement of conversion
The two most widely used techniques to assess the
extent of polymerization in dental composites have
been the physical determination of surface hardness and the direct chemical analysis of conversion
by mid-infrared (mid-IR) spectroscopy. The surface
hardness value of a composite can be correlated
with conversion but the results will vary based on
storage and testing conditions. A relatively simple
test for adequate conversion in a composite specimen involves the comparison of hardness on the top
and bottom surfaces [24]. A generally acceptable
result is obtained when the bottom surface hardness is at least 8090% of the top surface value,
usually on a 2 mm specimen thickness. This type of
measurement indicates uniform conversion across
the material, presumably at or near the limiting
conversion available for a given material and curing
condition. Alternatively, the direct measurement
of conversion by mid-IR spectroscopy has been an
extremely valuable tool not only for the general
characterization of dental resins and composites,
but also in numerous studies that have advanced

J.W. Stansbury et al.


the fundamental polymer science that underlies
these dental materials. A variety of IR techniques
are routinely used in dental polymer evaluations.
Thin films (ca. 525 mm) of unfilled resins and, in
some cases, filled composites can be analyzed in
transmission mode. For thicker specimens more
representative of clinical applications of composites, the polymeric material can be ground and
pressed into a salt matrix pellet or fractured for
direct study under an IR microscope. Reflectancebased IR techniques allow conversion of thick
specimens to be analyzed nondestructively,
although it should be recognized that with this
approach, only the immediate surface region of the
material is sampled. Other studies have effectively
utilized Raman spectroscopy, which relies on
scattering of the IR signal at the surface of a
material [25,26]. Recently, near-IR spectroscopy
has been applied to the characterization of conversion in dental polymers with the primary
advantage that thick specimens can be analyzed
simply in transmission [27,28]. Previous to this,
near-IR also had been used to evaluate water
uptake in dental materials [29,30].
As opposed to static determinations of final
conversion in a polymer, the dynamic measurement
of conversion during polymerization, either directly
with some form of IR spectroscopy, or indirectly by
calorimetry, is a useful exercise that can aid in
understanding any dynamic property that develops
with polymerization. Photo-differential scanning
calorimetry (photo-DSC) has been extensively used
to examine the reaction kinetics of many dimethacrylate monomers and comonomer mixtures [31]. In
a more limited way, it has also been used to study
composite photopolymerizations. The photo-DSC
technique is limited to microgram quantities of
materials and as such, a purged atmosphere is
generally required to avoid excessive oxygen
inhibition effects. The analysis of conversion
kinetics by DSC does have some limitations. The
temporal resolution of the technique, which is
based on the measurement of heat flux arising from
the exothermic polymerization reaction, may not
be suitable for fast-reacting materials [32]. Also the
sensitivity is not typically adequate to accurately
monitor the subtle continuous increase in conversion that occurs as the limiting conversion is
approached in glassy polymers.
Advantages associated with the real-time
measurement of conversion by IR spectroscopic
techniques include a high temporal resolution that
can follow extremely rapid kinetic processes
through the direct analyses of the changing methacrylate (or other) group concentration. In addition,
since full spectra can be obtained at each data

Shrinkage stress and strain in composites


sampling interval over the course of the polymerization, other coordinated events that have a
separate response in another spectral region can
be monitored. This information-rich data stream
may include features such as changes in hydrogen
bonding interactions or the polymerization reaction
of a second type of monomer, which can be
examined in detail as a function of the primary
reaction kinetics being analyzed. Post-cure that
follows the extinction of the curing light source can
be examined with high sensitivity using IR-based
methods and with nondestructive IR techniques,
the same specimens can be reanalyzed for
additional post-cure after extended storage
intervals.

Development of shrinkage stress and


strain during polymerization
It is apparent that both polymerization shrinkage
and modulus evolution are dependent on conversion. As such, polymerization stress must also
display conversion dependence since stress development is directly proportional to the instantaneous product of the shrinkage and modulus
during polymerization. The stress developed is
also affected by the specimen geometry, configuration of bonded surfaces and substrate compliance
[33,34]. The ability of partially or fully cured
polymeric networks to flow and thereby relieve
some portion of the stress is also a factor that has
received considerable attention [7,3540]. Because
of the fundamental, although not necessarily linear
influence that conversion exerts on the development of all polymer properties, it is evident that
efforts should be made to include conversion data
whenever possible in studies involved with characterization of polymeric dental materials.
In the network-forming photopolymerization of
dimethacrylate monomers, there is a very limited
pre-gel regime. The gel point is defined as the
appearance of an insoluble polymer fraction and it
involves a continuous network structure that can
span the entire macroscopic specimen dimension.
In theory, the gel point for polymerizations of
tetrafunctional monomers is expected at much less
than 1% conversion [41,42]. However, in practice,
cyclization reactions that do not lead to effective
crosslink formation results in gel points of less than
5% conversion. Many dental materials studies have
rightly pointed out that shrinkage strain that occurs
prior to gelation does not contribute to stress since
this involves viscous but unrestricted flow [43].
However, there may be some confusion as to when

59
the gel point is reached and how this feature can
potentially be manipulated based on modified
photo-curing protocol. The gel point is conversion
dependant and this critical extent of polymerization is not expected to be dramatically altered by
slowing the polymerization process through use of
ramped, stepped or pulse photo-curing modes. A
reduction in the concentration of active radical
centers, which would be the effect of limiting initial
irradiance, should result in increased primary chain
lengths and thus, theoretically in decreased gel
point conversion [44] in direct opposition to the
desired goal. Once gelation occurs at very low
conversion, viscoelastic flow is still possible, but
the timescale is increased relative to flow in the
pre-gel regime [4547]. Obviously, shrinkage continues beyond the gel point and in constrained
systems, even at this early stage of conversion with
the weak polymeric network structure plasticized
by a large excess of free monomer, external stress
can begin to be conducted to the bonded
interfaces.
As the reaction continues beyond the gel point,
the rate maximum in photopolymerizations of
dental resins and composites is encountered generally over the conversion range of approximately
1020% depending in part on the resin viscosity
[48,49]. If the photo-curing conditions are modified
to produce an increased reaction rate, the rate
maximum shifts to higher conversion. With unfilled
resins, there can be a significant exothermic
temperature rise that accompanies the rapid
polymerization of clinically relevant specimen
thicknesses. In dental composites, the heat rise
during cure is moderated by the high proportion of
inert filler included [50]. For quite appropriate
clinical reasons, many studies have assessed the
heat rise at the cavity floor or within the pulp
chamber due to the coupled heating effect of the
curing light and the reaction exotherm during
composite polymerization [5153]. Somewhat less
consideration has been devoted to how the polymer
formation itself is affected by the temperature
variations that occur within the polymer [5457].
With the relatively viscous resin systems used in
dental composites, even without considering the
addition of filler, it is typical that photopolymerization reactions result in an immediate onset of autoacceleration [48]. This occurs where initial viscosity
is high enough to retard macro-radical termination
reactions but not so high as to significantly inhibit
free monomer diffusional mobility.
As modulus continues to increase along with
conversion, the glass transition temperature of the
developing polymer reaches the effective cure
temperature [58]. As this vitrification stage is

60
approached, the material transitions from a rubbery to a glassy polymer. Unlike the gel point, which
has a well-defined onset, vitrification is a more
gradual process due to the previously mentioned
heterogeneity of the polymer network at this latter
stage of conversion. The residual unsaturation
present in fully cured dental polymers is a mixture
of pendant reactive groups and free monomer. For
dental resins, the development of gel fraction as a
function of conversion fits the classical calculations
for crosslinked polymers but bulky base monomers
such as Bis-GMA or UDMA exhibit decreased reactivity relative to TEGDMA as conversion progresses
[59]. Residual unreacted TEGDMA, which is more
hydrophilic and more mobile than the base monomers, is known to be the predominant monomer
that rapidly leaches from dental composites stored
in aqueous environments [6062]. However, the
eventual loss of the unreacted base monomer that
constitutes the majority portion of free monomer
left behind in dental composites, may contribute in
part to a gradual decline in polymer mechanical
properties over extended aqueous storage intervals. Therefore, when considering polymer stability, higher conversion appears desirable even
though this also leads to greater shrinkage and
stress development.
There have been several studies that evaluated
the rate of shrinkage stress or strain development
based on the reaction kinetics. In photopolymerization reactions, the conversion rate can be adjusted
easily by changing photo-initiator concentration or
curing light incident irradiance. In an alternate
approach to control reactivity, additional inhibitor
was incorporated in the resin to delay the early
stages of the reaction [63,64]. Regardless of the
means used to manipulate the reaction kinetics, as
expected, higher reaction rates lead to higher rates
of shrinkage and stress development [6469]. There
is concern that rapid development of polymerization shrinkage stress and strain may compromize
the competitive formation of an adequate adhesive
bond between the composite and the tooth [70].
Related to this, not only is depth of cure a factor
that impacts dentin bonding, but also a 2 mm thick
layer of composite clearly does not behave as an
ideal thin film which experiences uniform active
wavelength irradiation exposure and radical production simultaneously throughout the material
[7173]. Therefore, even with adequate conversion
achieved at the cavity floor, it can be assumed that
shrinkage stress develops at different rates across
the thickness of the restoration.
The kinetics of a photopolymerization reaction
are important in considerations of polymerization
shrinkage strain development since several studies

J.W. Stansbury et al.


have demonstrated the potential for conversion
kinetics to exceed the shrinkage kinetics in rapid
polymerizations that form glassy polymers [74,75].
A lag in shrinkage with respect to conversion can be
explained by the molecular level chemical reactions leading to conversion as opposed to the
coordinated network rearrangement necessary for
bulk physical shrinkage. It is reasonable that these
two processes could occur on significantly different
time scales. The reaction kinetics also determines
the extent of the temperature rise associated with
the reaction exotherm, which is added to any
radiant heating effects due to the incident
irradiation. Elevated temperature will serve to at
least partially counter the increase in density
connected to polymer formation. Higher effective
cure temperatures also delay the onset of mobility
restrictions to later stages of conversion; thus
raising the final conversion that can be achieved.
The extent of the thermal excursion is also related
to specimen geometry and heat transport properties of surrounding substrates. Any increase in
specimen temperature during cure acts to effectively reduce both shrinkage and modulus, which in
concert can diminish a portion of the stress
development temporarily until the exotherm
subsides.

Simultaneous measurement of
conversion with shrinkage stress or
strain
For the dynamic studies that track the development
of shrinkage stress or strain during photopolymerization, techniques have recently been introduced
in this lab that allow the simultaneous, real-time
monitoring of conversion on the same specimen.
Near-IR spectroscopy, which covers the vibrational
spectrum between 14,000 and 4000 cmK1, was
found to be uniquely suited to the requirements
imposed by conversion analysis under these conditions. Some of the advantages afforded by near-IR
include tolerance of thick specimen geometries
(0.25 to O5 mm) and the virtual transparency of
glass over this spectral range, which allows both
unfilled resins and highly filled dental composites to
be analyzed in transmission mode. Two other
critical features for the success of this approach
include the high efficiently of near-IR transmission
through fiber optic cables as well as no requirement
of a purged environment. This allows the near-IR
signal to be conveniently routed to a remote
specimen mounted in separate instrumentation,

Shrinkage stress and strain in composites


such as that used for dynamic shrinkage strain or
stress analysis.
There are several experimental devices for
active measurement of polymerization shrinkage.
These include dilatometry, interferometry, strain
gage techniques, optical scanning and direct
linear displacement transducer measurements
[65,7686]. In our lab, one method to monitor
shrinkage uses a modified linometer (constructed
by Academisch Centrum Tandheelkunde Amsterdam) to provide dynamic linear displacement
during the photopolymerization of resins or composites [79]. With the application of a thin layer
grease of the surfaces in contact with the polymer,
the assumption can be made, although not assured,
that isotropic shrinkage of the specimen is
achieved, which then allows the linear displacement data to be converted to volumetric shrinkage
data. In a modification to the linometer set-up
(Fig. 1), a stage was added to provide transverse
alignment for the near-IR fiber optic cables as well
as a separate channel that provides access for a
thermocouple to monitor the temperature changes
within the specimen over the course of the
polymerization. The linometer was selected for
this application since in addition to the accessible
specimen design, it allows shrinkage analysis on a
wide range of material viscosities from highly filled
composite pastes to unfilled diluent monomers. The
separate data collected on shrinkage and internal
temperature as a function of time can be replotted
using conversion as the common index. An example

Figure 1 Schematic of the modified linometer (ACTA)


coupled with fiber optic near-IR spectroscopy for monitoring dynamic volume shrinkage and polymerization
kinetics simultaneously. A specimen is placed between
the fixed glass cover and movable aluminum disc with the
metal stage providing a uniform 1 mm specimen thickness
as well as alignment for the optical fibers.

61

Figure 2 Temperature variation (dT) and volumetric


polymerization shrinkage (VS) as a function of conversion,
for Bis-GMA/TEGDMA systems. All samples contain
0.2 wt% DMPA and were irradiated with 4 mW/cm2
incident irradiance at 320390 nm continuously for
30 min. All the temperature profiles returned to ambient
within 10 min.

of this analysis technique is shown in Fig. 2 where


very different dynamic shrinkage results are evident
for the more reactive Bis-GMA/TEGDMA copolymerizations compared with the homopolymerization of
TEGDMA [87]. The shrinkage of TEGDMA rapidly
increases beyond the gel point (evident at approximately 5% conversion) and becomes nonlinear with
respect to conversion. The slowing in rate of
shrinkage development is apparently due to thermal expansion in response to the exotherm generated by the polymerization reaction. Other studies
conducted at higher reaction rates show more
dramatic temporary reductions in shrinkage that
also directly coincide with the peak temperature
rise. For composite materials, the lower temperature rise and the lower coefficient of thermal
expansion compared with unfilled resins [88],
would tend to reduce this effect. In the vitrified
state near the end of the TEGDMA reaction, the
conversion increase is clearly not matched by a
corresponding increase in shrinkage. This is direct
evidence of a lag between shrinkage and conversion
in glassy polymers.
The profiles of the Bis-GMA/TEGDMA resins in
Fig. 2 show a much lower rate of shrinkage
development over the low conversion regime. This
result again appears to be coordinated with the
exothermic heat rise, which occurs at much lower
conversion in the copolymerizations of these more
reactive resins compared with the TEGDMA homopolymerization [48]. The data also confirms the
expected results that show higher conversion and
markedly higher shrinkage associated with TEGDMA

62
homopolymerization relative to copolymerizations
that include significant proportions of Bis-GMA [7].
Preliminary studies of polymerizations conducted
over a variety of reaction rates, controlled based on
photo-curing conditions, have demonstrated that
higher curing rates can provide polymers with
higher conversion. However, surprisingly, the final
shrinkage reached may be less than that for the
same polymer cured to a somewhat lower conversion at a slower rate. The explanation for this
appears to be based on the extent of the exothermic temperature rise, which allows the network
formation to take place in a more expanded state in
the case of reactions performed at higher rates.
This implies that subtle structural differences may
be introduced into dental polymers as a consequence of the photo-curing process used and
temperature rise experienced.
As with shrinkage, there are several diverse
techniques available for shrinkage stress analysis.
These include the bonded disc method, photoelastic
analysis and variations on beam bending principles
[34,8996]. With the complex behavior of dynamic
shrinkage strain, it is not surprising that shrinkage
stress development, which depends in part on the
evolution of shrinkage, is also complex. The fiber
optic near-IR conversion measurement approach
already described was also adapted to dental resin
and composite stress measurement specimens
mounted in a cantilever beam tensometer (constructed by the American Dental Association Health
Foundation) [97]. With this device, the shrinkage
strain arising from polymer formation induces a
deflection in a calibrated cantilever beam with
beam displacement measured by a linear transducer
(Fig. 3) [98]. The level of compliance in this system
can be conveniently adjusted to match that of a
variety of clinically relevant dental restoration
configurations [99]. This instrument provides the
necessary optical access to the specimen during
testing so conversion monitoring can be readily
combined. Most studies of dynamic stress development in dental resins and composites have presented this data in the form of stress versus time,
which has limited meaning in the absence of any
conversion information. Other dental materials
studies have correlated stress development with
conversion, but in these investigations, the analyses
were conducted separately on different specimens
[40,64]. These more comprehensive studies show a
delay in stress development with respect to the
progress of conversion. In our view, the simultaneous dynamic monitoring of shrinkage stress as
a function of conversion provides an even more
informative and versatile method to characterize
stress development. As shown in Fig. 4, the stress

J.W. Stansbury et al.

Figure 3 The tensometer shrinkage stress measurement device (ADAHF; data acquisition and processing
control box not shown): (a) cantilever beam holder; (b)
upper collet holder; (c) cantilever beam; (d) LVDT; (e)
base stand; (f) curing light guide; (g) upper collet; (h) top
quartz rod; (i) PTFE sleeve; (j) sample; (k) bottom quartz
rod; (l) lower collet. Fiber optic near-IR cables (not
shown) are positioned opposite each other to transmit
through the specimen and containment sleeve.

development initially increases gradually in a nearly


linear fashion with conversion beyond the gel point.
As the vitrification stage is approached and the
increase in polymer modulus exceeds the rate at
which conversion continues to develop, the shrinkage stress begins to rise dramatically. This demonstrates that although stress development
commences immediately after the gel point is
reached at very low conversion, the large majority
of the overall stress that develops, is focused in

Figure 4 Analysis of stress development as a function


of conversion for a Bis-GMA/TEGDMA resin (7:3 wt ratio)
photopolymerization conducted at 28 mW/cm2 showing
the concentrated build-up of stress in the final stages of
conversion.

Shrinkage stress and strain in composites


the final stage of polymerization. These results are
in agreement with separate studies of conversiondependent stress buildup in UV-cured thin film
coatings [38,95,96].
The effect of the reaction exotherm can also be
observed in the profile of the stress development.
For photopolymerization reactions conducted at
higher rates, obtained by increased incident irradiance, preliminary studies demonstrate that not
only was higher final conversion achieved along
with a higher ultimate stress level, but also that
the vitrification stage associated with rapid stress
development was delayed to higher conversion
values. The potential for stress relaxation mechanisms to relieve the overall stress was also
investigated by conducting partial cure studies in
which the irradiation was terminated at various
stages of conversion prior to full cure [100].
As seen in Fig. 5, when a Bis-GMA/TEGDMA
composite was photopolymerized at higher incident irradiance compared with the data shown in
Fig. 4, the higher reaction rate and exotherm
delayed the onset of the rapid rise in stress to
higher conversion. If the photopolymerization
reaction is interrupted at relatively low degrees
of conversion while the developing polymer is still
in the rubbery state, a substantial degree of stress
relaxation can be achieved given sufficient time. It
should be noted, however, that over this region,
the stress levels available for any potential stress

Figure 5 Final shrinkage stress value as a function of


the final double bond conversion of Bis-GMA/TEGDMA
(70/30 by wt; initiator: camphorquinone 0.3 wt%; ethyl 4N,N-dimethylaminobenzoate 0.8 wt%) filled with 30 wt%
silanized barium glass partially cured for 2 s ($), 3 s (;),
6 s (:), or 10 s (C) at 450 mW/cm2. The conversion
values corresponding to the point of light extinction were
approximately 6, 21, 38 and 54%, respectively. Thus,
post-cure was most pronounced in specimens that
received the least irradiation. Continuous shrinkage
stress development is shown as a function of conversion
for the composite cured for a full 60 s interval ( ).

63
relief represents a relatively small fraction of the
overall stress expected to ultimately develop
should polymerization be reinitiated. Also the
relaxation times involved even at the fairly low
conversions are over the range of 530 min, which
would limit the clinical practicality of photo-curing
protocols that delay completion of polymerization.
If the reaction is halted after the vitrification
process begins, no significant stress reduction was
observed even when the stress was monitored for
as much as 8 h following a 60 s photo-cure. The
stress values actually continue to rise in the dark
due to continued post-cure that is common in
highly crosslinked polymeric systems [101]. It
appears that significant stress relaxation can only
occur in the low conversion regime where relatively little stress is actually developed.

Materials approaches to reduced


shrinkage stress and strain
As previously mentioned, considerable research
activity has been directed at the problems associated with polymerization shrinkage in dental
composites. The efforts to minimize the shrinkage
and/or stress that result from the formation of a
dental polymer can be divided into two areas:
changes in monomer structure or chemistry and
changes in fillers or use of additives. Some or all of
the typical dimethacrylate monomers used can be
replaced with alternative methacrylates that
reduce the reactive group density. Related to
this, bulky or oligomeric methacrylate monomer
structures have been developed and evaluated
[102105]. Ordered liquid crystalline monomers
that ideally yield amorphous polymers have been
developed as another approach to produce minimal
shrinkage. Within conventional monomer compositions, increased monomer functionality, which
results in increased reactive group density, provides higher levels of stress development, even at
lower values of final conversion [106]. Related work
has established that cross-linking photopolymerization of methacrylate monomers reaches much
higher stress levels for a given conversion compared
with acrylate monomers of similar structures [107].
By the use of oxirane (epoxy) or other ring-opening
polymerization approaches, the differences in
monomer chemistry and reaction mechanism compared with methacrylate resin systems can yield
significantly lower stress compared with methacrylate resins of similar cross-link density [108,109]. In
our lab, there are a number of aspects involving
variations in monomer structure currently under
investigation. These include bulky derivatives of

64
Bis-GMA as base monomers, dimer acid-derivatives
as alternative diluent monomers that produce
reaction-induced phase separation during photopolymerization, hybrid radical/cationic methacrylate/vinyl ether systems, step-growth thiol-ene
resin compositions that produce true delayed
gelation relative to addition polymers and studies
on expanding monomers related to spiro
orthocarbonates.
Modifications to stress and strain development
can also be made based on the fillers used in a
dental composite. It is well known that maximized
filler loading is one avenue to reduced shrinkage but
surprisingly, there has been little systematic work
conducted on the direct effect of filler volume on
shrinkage stress [91]. These reports indicate stress
increases with filler loading. Our preliminary efforts
on this topic using the simultaneous conversion and
stress measurement technique indicate that the
incremental addition of filler results in progressive
slight reductions in conversion, which provide for
decreases in ultimate stress, even though the
modulus increases with the filler content. Microscopic porosity in the resin or composite prior to
polymerization, results in reduced shrinkage and
stress development due to the formation of
significantly enlarged voids in the polymer
[110,111]. The introduction of non-bonded microor nano-filler has been used to achieve controlled
pore structure and reduced shrinkage and stress in
composites [91,112,113]. Polymerization-induced
phase separation, either through selection of
comonomer composition or by the use of specialized additives, can result in appreciably lower
shrinkage and stress than similar systems that
remain homogeneous [114116]. The addition of
plasticizers or comonomers that contribute toward
lower modulus coatings and bulk polymers also can
significantly reduce the final stress level of the
cured material. There are currently several
approaches in our lab related to the use of modified
fillers or additives to reduce stress in composites.
One method is based on the introduction of new
silane treatments for fillers that produce a polymer
brush surface rather than the traditional shortchain silane-tethered methacrylate surface. These
modified fillers allow higher filler loading, higher
conversion and yet lower stress compared with
conventional dental composites. Another strategy
involves the formation of nanogel polymer structures with high molecular weights. These prepolymerized fillers can be dissolved in typical dental
resins resulting in final polymers that have significantly lower shrinkage and stress than the unmodified resins and composites. In either the modified
filler or nanogel prepolymer approach, mechanical

J.W. Stansbury et al.


properties can be improved with respect to the
unmodified control materials.

Conclusions
The recent results from our lab regarding conversion dependent stress development and relaxation
in dental resins and composites indicate that
modified photo-curing protocols including stepped
or pulsed irradiation programs in their current
forms appear to have limited potential to reduce
the overall stress developed if comparable levels of
conversion are achieved relative to a standard
protocol. It should be noted that other studies have
demonstrated different results in which composite
stress can be reduced based on curing mode
without compromize to conversion [117]. Indeed,
at least some of the many conflicting results
reported in the dental materials literature regarding the efficacy of modified photo-curing modes on
stress development and marginal leakage in composite restoratives could well be explained by
relatively small discrepancies in conversion that
could be expected to arise from the variations in
reaction kinetics involved. The ongoing development of new and improved analytical techniques
that integrate conversion measurement with the
characterization of polymerization shrinkage stress
and strain will allow more informed evaluations of
the effects of modified curing modes. In addition,
the many varied approaches to new restorative
materials specifically designed to provide significant reductions in shrinkage and stress will also
benefit from the availability of improved testing
methods.

Acknowledgements
This investigation was supported in part by NIH/
NIDCR grant DE14227. The corresponding author is
grateful to Ron Sakaguchi and the other organizers
of the Portland Composites Symposium for the
invitation to present there as well as the additional
opportunity to present this work and these opinions
here.

References
[1] Services Rendered Survey of 1999 by the ADA Health Policy
Resources Center; 2002.
[2] Ferracane JL, Mitchem JC. Relationship between composite contraction stress and leakage in Class V cavities. Am
J Dent 2003;16:23943.

Shrinkage stress and strain in composites


[3] Davidson CL, Feilzer AJ. Polymerization shrinkage and
polymerization shrinkage stress in polymer-based restoratives. J Dent 1997;25:43540.
[4] Versluis A, Douglas WH, Cross M, Sakaguchi RL. Does an
incremental filling technique reduce polymerization
shrinkage stresses? J Dent Res 1996;75:8718.
[5] Hilton TJ. Can modern restorative procedures and
materials reliably seal cavities? In vitro investigations.
Part 1 Am J Dent 2002;15:198210.
[6] Patel MP, Braden M, Davy KWM. Polymerization shrinkage
of methacrylate esters. Biomaterials 1987;8:536.
[7] Venhoven BAM, de Gee AJ, Davidson CL. Polymerization
contraction and conversion of light-curing BisGMA-based
methacrylate resins. Biomaterials 1993;14:8715.
[8] Rueggeberg F, Tamareselvy K. Resin cure determination by
polymerization shrinkage. Dent Mater 1995;11:2658.
[9] Dusek K. Polymer networksstructure, formation and
properties. J Bioact Compat Polym 1991;6:24755.
[10] Cook WD. Photopolymerization kinetics of oligo(ethylene
oxide) and oligo(methylene) oxide dimethacrylates.
J Polym Sci: Part A: Polym Chem 1993;31:105367.
[11] Matsumoto A. Mechanistic discussion on deviation from
ideal network formation in radical polymerization of
multivinyl monomers. Makromol Chem: Macromol Symp
1993;76:3342.
[12] Anseth KS, Wang CM, Bowman CN. Kinetic evidence of
reaction-diffusion during the polymerization of multi
(meth)acrylate monomers. Macromolecules 1994;27:
6505.
[13] Anseth KS, Newman SM, Bowman CN. Polymeric dental
composites: properties and reaction behavior of multimethacrylate dental restorations. Biopolymers II 1995;
122:177217.
[14] Anseth KS, Goodner MD, Reill MA, Kannurpatti AR,
Newman SM, Bowman CN. The influence of comonomer
composition on dimethacrylate resin properties for dental
composites. J Dent Res 1996;75:160712.
[15] Morgan DR, Kalachandra S, Shobha HK, Gunduz N,
Stejskal EO. Analysis of a dimethacrylate copolymer (BisGMA and TEGDMA) network by DSC and C-13 solution and
solid-state NMR spectroscopy. Biomaterials 2000;21:
1897903.
[16] Andrzejewska E. Photopolymerization kinetics of multifunctional monomers. Prog Polym Sci 2001;26:60565.
[17] Elliott JE, Bowman CN. Monomer functionality and
polymer network formation. Macromolecules 2001;34:
46429.
[18] Lovell LG, Berchtold KA, Elliott JE, Lu H, Bowman CN.
Understanding the kinetics and network formation of
dimethacrylate dental resins. Polym Adv Technol 2001;
12:33545.
[19] Elliott JE, Nie J, Bowman CN. The effect of primary
cyclization on free radical polymerization kinetics: experimental characterization. Polymer 2003;44:32732.
[20] Dusek K. Are cured thermoset resins inhomogeneous?
Angew Makromol Chem 1996;240:115.
[21] Kannurpatti AR, Anseth JW, Bowman CN. A study of the
evolution of mechanical properties and structural heterogeneity of polymer networks formed by photopolymerizations of multifunctional (meth)acrylates. Polymer 1998;
39:250713.
[22] Lu H, Lovell LG, Bowman CN. Exploiting the heterogeneity
of cross-linked photopolymers to create High Tg polymers
from polymerizations performed at ambient conditions.
Macromolecules 2001;34:80215.
[23] Rey L, Duchet J, Galy J, Sautereau H, Vouagner D,
Carrion L. Structural heterogenieties and mechanical

65

[24]

[25]

[26]

[27]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

properties of vinyl/dimethacrylate networks synthesized


by thermal free radical polymerization. Polymer 2002;43:
437584.
Cohen ME, Leonard DL, Charlton DG, Roberts HW,
Ragain JC. Statistical estimation of resin composite
polymerization sufficiency using microhardness. Dent
Mater 2004;20:15866.
Pianelli C, Devaux J, Bebelman S, Leloup G. The microRaman spectroscopy, a useful tool to determine the
degree of conversion of light-activated composite resins.
J Biomed Mater Res 1999;48:67581.
De Santis A, Baldi M. Photo-polymerisation of composite
resins measured by micro-Raman spectroscopy. Polymer
2004;45:3797804.
Lovell L, Lu H, Elliott J, Stansbury J, Bowman C. The effect
of cure rate on the mechanical properties dental resins.
Dent Mater 2001;17:50411.
Stansbury J, Dickens S. Determination of double bond
conversion in dental resins by near infrared spectroscopy.
Dent Mater 2001;17:719.
Venz S, Dickens B. NIR-spectroscopic investigation of
water sorption characteristics of dental resins. J Biomed
Mater Res 1991;25:123148.
Diaz-Arnold A, Arnold M, Williams V. Measurement of
water sorption by resin composite adhesives with nearinfrared spectroscopy. J Dent Res 1992;71:43842.
Cook WD. Photopolymerization kinetics of dimethacrylates using the camphorquinone amine initiator system.
Polymer 1992;33:6009.
Decker C, Moussa K. A new method for monitoring ultrafast photopolymerizations by real-time infrared (RTIR)
spectroscopy. Makromol Chem: Macromol Chem Phys
1988;189:238194.
Alster D, Venhoven BAM, Feilzer AJ, Davidson CL. Influence
of compliance of the substrate materials on polymerization contraction stress in thin resin composite layers.
Biomaterials 1997;18:33741.
Laughlin GA, Williams JL, Eick JD. The influence of system
compliance and sample geometry on composite polymerization shrinkage stress. J Biomed Mater Res 2002;63:
6718.
Feilzer AJ, De Gee AJ, Davidson CL. Relaxation of
polymerization contraction shear stress by hygroscopic
expansion. J Dent Res 1990;69:369.
Sakaguchi RL, Ferracane JL. Stress transfer from polymerization shrinkage of a chemical-cured composite bonded
to a pre-cast composite substrate. Dent Mater 1998;14:
10611.
Labella R, Lambrechts P, Van Meerbeek B, Vanherle G.
Polymerization shrinkage and elasticity of flowable composites and filled adhesives. Dent Mater 1999;15:12837.
Vaessen DM, Ngantung FA, Palacio MLB, Francis LF,
McCormick AV. Effect of lamp cycling on conversion and
stress development in ultraviolet-cured acrylate coatings.
J Appl Polym Sci 2002;84:278493.
Chen HY, Manhart J, Kunzelmann KH, Hickel R. Polymerization contraction stress in light-cured compomer
restorative materials. Dent Mater 2003;19:597602.
Feilzer AJ, Dauvillier BS. Effect of TEGDMA/BisGMA ratio
on stress development and viscoelastic properties of
experimental two-paste composites. J Dent Res 2003;82:
82834.
Dusek K. Special features of network formation by chain
cross-linking copolymerization. Collect Czech Chem Commun 1993;58:224565.

66
[42] Jancar J, Wang W, Dibenedetto AT. On the heterogeneous
structure of thermally cured bis-GMA/TEGDMA resins.
J Mater Sci: Mater Med 2000;11:67582.
[43] Versluis A, Tantbirojn D, Douglas WH. Do dental composites always shrink toward the light? J Dent Res 1998;77:
143545.
[44] Dusek K. Network formation by chain cross-linking (co)polymerization. In: Haward RN, editor. Developments in
polymerization, vol. 3. London: Applied Science; 1982. p.
143206.
[45] Lange J, Manson J-AE. Build-up of structure and viscoelastic properties in epoxy and acrylate resins cured below
their ultimate glass transition temperature. Polymer 1996;
37:585968.
[46] Lange J. Viscoelastic properties and transitions during
thermal and UV cure of a methacrylate resin. Polym Eng Sci
1999;39:165160.
[47] Lange J, Toll S, Manson JAE. Residual stress build-up in
thermoset films cured below their ultimate glass transition
temperature. Polymer 1997;38:80915.
[48] Dickens SH, Stansbury JW, Choi KM, Floyd CJE. Photopolymerization kinetics of methacrylate dental resins.
Macromolecules 2003;36:604353.
[49] Trujillo M, Newman SM, Stansbury JW. Use of near-IR to
monitor the influence of external heating on dental
composite photopolymerization. Dent Mater 2004;20:
76677.
[50] Mohsen NM, Craig RG, Filisko FE. Effects of curing time and
filler concentration on curing and postcuring of urethane
dimethacrylate composites: a microcalorimetric study.
J Biomed Mater Res 1998;40:22432.
[51] Shortall AC, Harrington E. Temperature rise during
polymerization of light-activated resin composites.
J Oral Rehab 1998;25:90813.
[52] Hannig M, Bott B. In-vitro pulp chamber temperature rise
during composite resin polymerization with various lightcuring sources. Dent Mater 1999;15:27581.
[53] Weerakoon AT, Meyers IA, Symons AL, Walsh LJ. Pulpal
heat changes with newly developed resin photopolymerisation systems. Aust Endod J 2002;28:10811.
[54] Broer DJ, Mol GN, Challa G. Temperature effects on the
kinetics of photoinitiated polymerization of dimethacrylates. Polymer 1991;32:6905.
[55] Cook WD. Thermal aspects of the kinetics of dimethacrylate photopolymerization. Polymer 1992;33:215261.
[56] Lecamp L, Youssef B, Bunel C, Lebaudy P. Photoinitiated
polymerization of a dimethacrylate oligomer: 1. Influence
of photoinitiator concentration, temperature and light
intensity. Polymer 1997;38:608996.
[57] Scherzer T, Decker U. The effect of temperature on the
kinetics of diacrylate photopolymerizations studied by
real-time FTIR spectroscopy. Polymer 2000;41:768190.
[58] Lange J, Altmann N, Kelly CT, Halley PJ. Understanding
vitrification during cure of epoxy resins using dynamic
scanning calorimetry and rheological techniques. Polymer
2000;41:594955.
[59] Stansbury JW, Dickens SH. Network formation and compositional drift during photo-initiated copolymerization of
dimethacrylate monomers. Polymer 2001;42:63639.
[60] Muller H, Olsson S, Soderholm KJ. The effect of comonomer composition, silane heating, and filler type on
aqueous TEGDMA leachability in model resin composites.
Eur J Oral Sci 1997;105:3628.
[61] Spahl W, Budzikiewicz H, Geurtsen W. Determination of
leachable components from four commercial dental
composites by gas and liquid chromatography mass
spectrometry. J Dent 1998;26:13745.

J.W. Stansbury et al.


[62] Munksgaard EC, Peutzfeldt A, Asmussen E. Elution of
TEGDMA and BisGMA from a resin and a resin composite
cured with halogen or plasma light. Eur J Oral Sci 2000;
108:3415.
[63] Cook WD, Lau M, Mehrabi M, Dean K, Zipper M. Control of
gel time and exotherm behaviour during cure of unsaturated polyester resins. Polym Int 2001;50:12934.
[64] Braga RR, Ferracane JL. Contraction stress related to
degree of conversion and reaction kinetics. J Dent Res
2002;81:1148.
[65] Watts DC, Cash AJ. Kinetic measurements of photopolymerization contraction in resins and composites.
Meas Sci Technol 1991;2:78894.
[66] Silikas N, Eliades G, Watts DC. Light intensity effects on
resin-composite degree of conversion and shrinkage strain.
Dent Mater 2000;16:2926.
[67] Hofmann N, Denner W, Hugo B, Klaiber B. The influence of
plasma arc vs. halogen standard or soft-start irradiation on
polymerization shrinkage kinetics of polymer matrix
composites. J Dent 2003;31:38393.
[68] Watts DC, Marouf AS, Al-Hindi AM. Photo-polymerization
shrinkage-stress kinetics in resin-composites: methods
development. Dent Mater 2003;19:111.
[69] Alvarez-Gayosso C, Barcelo-Santana F, Guerrero-Ibarra J,
Saez-Espinola G, Canseco-Martinez MA. Calculation of
contraction rates due to shrinkage in light-cured composites. Dent Mater 2004;20:22835.
[70] Davidson CL, De Gee AJ, Feilzer A. The competition
between the composite-dentin bond strength and the
polymerization contraction stress. J Dent Res 1984;63:
13969.
[71] Kawaguchi M, Fukushima T, Miyazaki K. The relationship
between cure depth and transmission coefficient of visible
light-activated resin composites. J Dent Res 1994;73:
51621.
[72] Lecamp L, Lebaudy P, Youssef B, Bunel C. Influence of UV
radiation wavelength on conversion and temperature
distribution profiles within dimethacrylate thick material
during photopolymerization. Polymer 2001;42:85417.
[73] Lee JH, Prudhomme RK, Aksay IA. Cure depth in
photopolymerization: experiments and theory. J Mater
Res 2001;16:353644.
[74] Kloosterboer JG, Lijten GFCM, Boots HMJ. Network
formation by chain crosslinking photopolymerization and
some applications in electronics. Makromol Chem: Macromol Symp 1989;24:22330.
[75] Anseth KS, Bowman CN, Peppas NA. Polymerization
kinetics and volume relaxation behavior of photopolymerized multifunctional monomers producing highly crosslinked networks. J Polym Sci: Part A: Polym Chem 1994;
32:13947.
[76] Penn RW. A recording dilatometer for measuring polymerization shrinkage. Dent Mater 1986;2:789.
[77] Lai JH, Johnson AE. Measuring polymerization shrinkage of
photo-activated restorative materials by water-filled
dilatometer. Dent Mater 1993;9:13943.
[78] Bandyopadhyay S. A study of the volumetric setting
shrinkage of some dental materials. J Biomed Mater Res
1982;16:13544.
[79] De Gee AJ, Feilzer AJ, Davidson CL. True linear polymerization shrinkage of unfilled resins and composites
determined with a linometer. Dent Mater 1993;9:1114.
[80] Fano V, Ma WY, Ortalli I, Pozela K. Study of dental materials
by laser beam scanning. Biomaterials 1998;19:15415.
[81] Hudson AJ, Martin SC, Hubert M, Spelt JK. Optical
measurements of shrinkage in UV-cured adhesives.
J Electron Packag 2002;124:3524.

Shrinkage stress and strain in composites


[82] Watts DC, Marouf AS. Optimal specimen geometry in
bonded-disk shrinkage-strain measurements on light-cured
biomaterials. Dent Mater 2000;16:44751.
[83] Sakaguchi RL, Sasik CT, Bunczak MA, Douglas WH. Strain
gauge method for measuring polymerization contraction of
composite restoratives. J Dent 1991;19:3126.
[84] Sakaguchi RL, Versluis A, Douglas WH. Analysis of strain
gage method for measurement of post-gel shrinkage in
resin composites. Dent Mater 1997;13:2339.
[85] Dudi O, Grubbs WT. Laser interferometric technique for
measuring polymer cure kinetics. J Appl Polym Sci 1999;
74:213342.
[86] Fogleman EA, Kelly MT, Grubbs WT. Laser interferometric
method for measuring linear polymerization shrinkage in
light cured dental restorations. Dent Mater 2002;18:
32430.
[87] Lin Y, Trujillo M, Stansbury JW. Simultaneous characterizations of polymerization reaction kinetics and volume
shrinkage in photo-cured dimethacrylate systems. Macromolecules, Submitted.
[88] Sideridou I, Achilias DS, Kyrikou E. Thermal expansion
characteristics of light-cured dental resins and resin
composites. Biomaterials 2004;25:308797.
[89] Feilzer AJ, Degee AJ, Davidson CL. Setting stress in
composite resin in relation to configuration of the
restoration. J Dent Res 1987;66:16369.
[90] Alster D, Feilzer AJ, de Gee AJ, Davidson CL. Polymerization contraction stress in thin resin composite layers as a
function of layer thickness. Dent Mater 1997;13:14650.
[91] Condon JR, Ferracane JL. Assessing the effect of composite formulation on polymerization stress. J Am Dent Assoc
2000;131:497503.
[92] Kinomoto Y, Torii M. Photoelastic analysis of polymerization contraction stresses in resin composite restorations.
J Dent 1998;26:16571.
[93] Benabdi M, Roche AA. Mechanical properties of thin and
thick coatings applied to various substrates. 1. An elastic
analysis of residual stresses within coating materials.
J Adhes Sci Technol 1997;11:28199.
[94] Payne JA, McCormick AV, Francis LF. In situ stress
measurement apparatus for liquid applied coatings. Rev
Sci Instrum 1997;68:45648.
[95] Stolov AA, Xie T, Penelle J, Hsu SL, Stidham HD. An analysis
of photopolymerization kinetics and stress development in
multifunctional acrylate coatings. Polym Eng Sci 2001;41:
31428.
[96] Francis LF, McCormick AV, Vaessen DM, Payne JA. Development and measurement of stress in polymer coatings.
J Mater Sci 2002;37:4897911.
[97] Lu H, Stansbury JW, Dickens SH, Eichmiller FC, Bowman
CN. Probing the origins and control of shrinkage stress in
dental resin-composites: II. Novel method of simultaneous
measurement of polymerization shrinkage stress and
conversion. J Biomed Mater Res Part B: Appl Biomater
2004; 71:20613.
[98] Lu H, Stansbury JW, Dickens SH, Eichmiller FC, Bowman
CN. Probing the origins and control of shrinkage stress in
dental resin-composites: I. Shrinkage stress characterization technique. J Mater Sci: Mater Med, in press.
[99] Meredith N, Setchell DJ. In vitro measurement of cuspal
strain and displacement in composite restored teeth.
J Dent 1997;25:3317.

67
[100] Lu H, Stansbury JW, Bowman CN. Towards the elucidation
of shrinkage stress development and relaxation in dental
composites. Dent Mater, in press.
[101] Tarumi H, Imazato S, Ehara A, Kato S, Ebi N, Ebisu S. Postirradiation polymerization of composites containing bisGMA and TEGDMA. Dent Mater 1999;15:23842.
[102] Pezron E, Magny B. Modeling of UV oligomers and
monomers properties: viscosity and shrinkage. Eur Coat
J 1996;6024.
[103] Choi KM, Stansbury JW. New families of photocurable
oligomeric fluoromonomers for use in dental composites.
Chem Mater 1996;8:27047.
[104] Klee JE, Neidhart F, Flammersheim HJ, Mulhaupt R.
Monomers for low shrinking composites. 2Synthesis of
branched methacrylates and their application in dental
composites. Macromol Chem Phys 1999;200:51723.
[105] Gao F, Schricker SR, Tong YH, Culbertson BM. Novel
trimethacrylates: Synthesis, characterization, and
evaluation of new monomers for improved dental
restoratives. J Macromol Sci: Pure Appl Chem 2002;
39:25165.
[106] Payne JA, Francis LF, McCormick AV. The effects of
processing variables on stress development in ultravioletcured coatings. J Appl Polym Sci 1997;66:126777.
[107] Wen M, Scriven LE, McCormick AV. Differential scanning
calorimetry and cantilever deflection studies of polymerization kinetics and stress in ultraviolet curing of
multifunctional (meth)acrylate coatings. Macromolecules
2002;35:11220.
[108] Moszner N, Salz U. New developments of polymeric dental
composites. Prog Polym Sci 2001;26:53576.
[109] Guggenberger R, Weinmann W. Exploring beyond methacrylates. Am J Dent 2000;13:824.
[110] Alster D, Feilzer AJ, Degee AJ, Mol A, Davidson CL. The
dependence of shrinkage stress reduction on porosity
concentration in thin resin layers. J Dent Res 1992;71:
161922.
[111] Eom Y, Boogh L, Michaud V, Sunderland P, Manson JA.
Stress-initiated void formation during cure of a threedimensionally constrained thermoset resin. Polym Eng Sci
2001;41:492503.
[112] Condon JR, Ferracane JL. Reduction of composite contraction stress through non-bonded microfiller particles.
Dent Mater 1998;14:25660.
[113] Condon JR, Ferracane JL. Reduced polymerization stress
through non-bonded nanofiller particles. Biomaterials
2002;23:380715.
[114] Kinkelaar M, Wang B, Lee LJ. Shrinkage behavior of lowprofile unsaturated polyester resins. Polymer 1994;35:
301122.
[115] Huang Y-J, Liang C-M. Volume shrinkage characteristics in
the cure of low-shrink unsaturated polyester resins.
Polymer 1996;37:40112.
[116] Vaessen DM, McCormick AV, Francis LF. Effects of phase
separation on stress development in polymeric coatings.
Polymer 2002;43:226777.
[117] Lim B-S, Ferracane JL, Sakaguchi RL, Condon JR.
Reduction of polymerization contraction stress for dental
composites by two-step light-activation. Dent Mater 2002;
18:43644.

Você também pode gostar