progress in dimethacrylate-based dental composite technology and curing efficiency

Upload: durga7

Post on 02-Jun-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    1/18

    d ent a l ma te r ia l s 2 9 ( 2 0 1 3 ) 139156

    Available online at www.sciencedirect.com

    journal homepage: www.int l .e lsevierheal th.com/ journals/dema

    Review

    Progress in dimethacrylate-based dental composite

    technology and curing efficiency

    Julian G. Leprince a,b,d,, William M. Palinc, Mohammed A. Hadis c,Jacques Devaux b,d, Gaetane Leloupa,b,d

    a School of Dentistry and Stomatology, Universit catholique de Louvain, Brussels, Belgium

    b Institute of Condensed Matter and Nanoscience, Bio- and Soft- Matter, Universit catholique de Louvain, Louvain-la-Neuve, Belgiumc Biomaterials Unit, University of Birmingham, College of Medical and Dental Sciences, School of Dentistry, St Chads Queensway,

    Birmingham B4 6NN, UKd CRIBIO (Center for Research and Engineering on Biomaterials), Brussels, Belgium

    a r t i c l e i n f o

    Article history:

    Received 27 September 2012

    Received in revised form

    3 November 2012

    Accepted 4 November 2012

    Keywords:

    Composite materials

    Resins

    Dimethacrylates

    Polymers

    Photoinitiators

    Operative dentistry

    Photopolymerization efficiency

    Curing lights

    Irradiation modes

    Curing time

    Irradiance

    a b s t r a c t

    Objectives. This work aims to review the key factors affecting the polymerization efficiency

    of light-activated resin-based composites. The different properties and methods used to

    evaluate polymerization efficiency will also be critically appraised with focus on the devel-

    opments in dental photopolymer technology and how recent advances have attempted to

    improve the shortcomings of contemporary resin composites.

    Methods. Apart from the classical literature on the subject, the review focused in particularon papers published since 2009. The literature research was performed in Scopus with the

    terms dental resin OR dimethacrylate. The list was screened and all papers relevant to the

    objectives of this work were included.

    Results. Thoughnew monomer technologieshavebeen developed andsome of them already

    introduced to the dental market, dimethacrylate-based composites still currently represent

    the vast majority of commercially available materials for direct restoration. The photopoly-

    merizationof resin-based composites has been the subject of numerous publications, which

    have highlighted the major impact of the setting process on material properties and quality

    ofthe final restoration. Many factors affect the polymerization efficiency, be they intrinsic;

    photoinitiator type and concentration, viscosity (co-monomer composition and ratio, filler

    content) and optical properties, or extrinsic; light type and spectrum, irradiation parame-

    ters (radiant energy, time and irradiance), curing modes, temperature and light guide tip

    positioning.

    Corresponding author at: Universit catholique de Louvain, Ecole de mdecine dentaire et de stomatologie, Centre de recherche CRIBIO,Avenue Hippocrate, 10/5721, B-1200 Brussels, Belgium. Tel.: +32 10 47 30 88; fax: +32 10 45 15 93.

    E-mail address:[email protected] (J.G. Leprince).0109-5641/$ see front matter 2012 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

    http://dx.doi.org/10.1016/j.dental.2012.11.005

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://www.sciencedirect.com/science/journal/01095641http://www.intl.elsevierhealth.com/journals/demamailto:[email protected]://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005mailto:[email protected]://www.intl.elsevierhealth.com/journals/demahttp://www.sciencedirect.com/science/journal/01095641http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    2/18

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    3/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 141

    drawbacks of resin composites and improve material per-

    formances. A primary disadvantage related to their use is

    polymerization shrinkage and the associated stress transmit-

    ted to the adhesive bond and the remaining tooth structure.

    Their clinical consequences include crack formation in dentin

    and enamel, post-operative sensitivity, marginal discoloration

    and secondary caries [6]. Further drawbacks of resin-based

    composites include inferior mechanical properties comparedto tooth structures [7], in line with clinical data reporting bulk

    fracture as the main reason for retreatment [3], as well as sec-

    ondary caries andtooth fractures [35]. Thebiocompatibility of

    dimethacrylate resin and photoinitiatorchemistry extensively

    described by [8] are further concerns, as it may affect the del-

    icate balance between healing and chronic inflammation in

    pulp tissues [9].

    Advances in material formulation may improve the

    shortcomings of resin composites and in this regard, recent

    reviews [1012] have summarized significant developments,

    including improved filler morphology, progress with existing

    dimethacrylate chemistry and novel monomer technologies.

    It appears from those works that, although some non-dimethacrylate monomers materials based on ring-opening

    epoxy chemistry (FiltekTM Silorane, 3M ESPE Dental Products,

    Seefeld, Germany) have already been introduced to the dental

    market, contemporary dimethacrylate-based composites

    still represent the vast majority of commercially available

    materials for direct restoration.

    As stated by [10], the development and implementation

    of resin composites rely on a comprehensive understanding

    of each component . . .. However, the successful placement of

    composite restorations depends also on the method of light

    curing as well as theeffect of thecomponentson theefficiency

    of the photopolymerization process. Therefore, the objective

    of the present work is to review the different factors affectingthe photopolymerization efficiency of dimethacrylate-based

    dental resins (Fig. 1), be they intrinsic; co-monomer com-

    position and ratio, filler content, photoinitiator type and

    concentration, or extrinsic; light spectrum, irradiation pro-

    tocols, temperature and light guide tip positioning, with

    particular focus on papers published since 2009. It is impor-

    tant for researchers, and also particularly for clinicians, to

    first comprehend, and second, to optimize the photopoly-

    merization process of resin-based composites in order to

    improve material properties. To better realize the impact

    of these different factors, a fundamental description of the

    photopolymerization reaction will be presented with a critical

    appraisal of the different properties and methods used toevaluate the polymerization efficiency. Recent advances that

    improve the photopolymerization process will be reviewed

    as well.

    2. Dimethacrylate photopolymerizationreaction characteristics

    The photopolymerization process of dimethacrylate-based

    dental resins is a reaction triggered by free radicals, which

    are generated by irradiation of a light-sensitive initiator and

    open the double bond of methacrylate groups (C C), gen-

    erating a chain reaction. It is classically described in three

    steps: initiation, propagation and termination, schematically

    represented in Fig. 2. Conventionally, the extent of polymer-

    ization is quantified by comparing the amount of remaining

    double bonds in the polymer structure to the initial amount.

    This ratio, expressed in %, is termed the degree of conversion

    (DC). Generally, DC values vary for a wide range of resin com-

    posite types, ca. 3577% [8]. Dimethacrylate monomers have

    the potential to bind to four other molecules or structures,

    whereas mono-methacrylates can only bind to the two adja-cent molecules within a linear polymer chain. As a result,

    and in contrast to forming unconnected long linear chains

    associated with thermoplastic polymers, a highly crosslinked

    network is formed, with covalent bonds between polymer

    chains. The crosslinked structure has major consequences

    on the material properties as well as on the kinetics of the

    polymerization reaction. Following photopolymerization, the

    degree of conversion and crosslinking density increaserapidly.

    An infinite network forms, resulting in a rapid increase of

    the system viscosity and in the first change of state, from a

    viscous liquid to an elastic gel, called gelation [10]. At the gel

    point, the mobility restriction mostly affects radicals, located

    on large molecules (growing polymer chains), whereas smallmonomer molecules can still diffuse easily. Consequently,

    bimolecular termination (Fig. 2) decreases dramatically while

    new growth centers are still created by initiation. Hence,

    the free radical concentration increases, which results in a

    rapid increase of the rate of polymerization (Rp, fraction of

    double bond converted per second, representing the speed

    of the reaction) called autoacceleration [13]. As the reaction

    proceeds, the viscosity becomes so high that it limits dif-

    fusion even for monomer molecules, resulting in significant

    decrease of Rp, or autodeceleration. This corresponds to the

    second change of state, from rubbery to glass, or vitrifica-

    tion [14], which can be visualized by plottingRp against DC

    (Fig. 3). Vitrification prevents any further extensive reaction,and explains why DC cannot reach 100%, even with optimum

    irradiation conditions. Though DC is known to slightly evolve

    up to 24h after irradiation, probably due to free volume relax-

    ation [15], the major part of the polymerization occurs during

    irradiation.

    Another major consequence of vitrification is the entrap-

    ment of free radicals in the polymer network, which is

    sometimes referredto as monomolecular termination in polymer-

    related publications [16,17]. However, such explanation may

    not be wholly appropriate, as trapped free radicals do not

    necessarily terminate since they still exist within the

    network for several weeks after vitrification [18,19]. If the

    mobility of the system increases (swelling by solvents, tem-perature increase, etc.) or through reaction-diffusion [13,19],

    radicals remain capable of reacting with remaining dou-

    ble bonds and continuing polymerization. Hence, a more

    accurate terminology may be defined as radical immobi-

    lization or radical entrapment. Besides free radicals, other

    active species remain trapped in the polymer network due

    to vitrification, i.e. pendant double bonds, free monomers

    and photoinitiators. The remaining active species can influ-

    ence the biological and/or mechanical properties of the

    material. Importantly, the characteristics of the photopoly-

    merization process are significantly modified by the radical

    concentration, the double bond concentration, the degree of

    crosslinking or the system viscosity [13], which depend on

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    4/18

    142 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    Fig. 1 Schematic representation of the different properties used to evaluate photopolymerization efficiency, and of various

    extrinsic and intrinsic factors by which it is affected. Gray arrows indicate the influence of one factor on another or on the

    curing efficiency. The black arrow symbolizes the fact that the curing efficiency is not only governed by extrinsic

    parameters, but by intrinsic parameters as well, since differences in inherent material properties have a major influence on

    the way extrinsic factors affect the success of photopolymerization.

    numerous intrinsic and/or extrinsic factors described fur-

    ther.

    3. Evaluation ofphotopolymerizationefficiency

    3.1. Degree ofconversion

    DC is a very important material feature as it is significantly

    correlated to several other important material characteristics,

    such as mechanical properties [20,21], volumetric shrinkage

    [22], wear resistance [23] and monomer elution [24]. There-

    fore, DC is frequently measured to evaluate polymerization

    efficiency and the most common method is by spectro-

    scopic techniques that infer the quantity of remaining double

    bonds, either mid-infrared Fourier transform (FT) [25] or

    Raman spectroscopy [26]. FT mid-IR is based on the reflec-

    tion of infrared radiation and has been used for many years

    to measure DC by comparing the height of the 1640cm1

    peak corresponding to CH CH2 (vinyl) stretching vibration

    before and after curing[25]. Another peak corresponding to

    aromatic rings at 1608cm1 is used as a reference as it does

    not vary with polymerization. Raman spectroscopy is based

    on the dispersion of the light by the polymer, using similar

    measurement peaks as mid-IR spectroscopy [26]. Microscope

    attachments for mid-IR techniques and the focused beam

    used for micro-Raman spectroscopy enable the measurement

    of DC at specific points by mapping the surface of a sample

    under high magnification, which is useful when considering

    polymerization in depth. A disadvantage of mid-IR techniquesremains high absorption in this wavelength range, which

    may decrease signal-to-noise ratio and increase variability of

    results. While it is generally assumed that vinyl groups on

    silane are neglectible compared to the amount in the resin, it

    has also been reported that vinyl C C absorption of the silane

    layer at the fillermatrix interface may complicate polymer

    conversion calculations [27]. More recently, near-infrared FT

    spectroscopy (FT NIR), which is based on transmission, was

    shown to be an efficient and a more reliable method to mea-

    sure DC in real time, providing accurate measurement ofRp.

    FT NIR is based on transmission, and monitors the decrease

    in the vinyl peak centered at 6164cm1 [28] and less absorp-

    tion allows analysis of thick specimens. DC has also been

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    5/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 143

    Fig. 2 Schematic representation of the 3 steps of photopolymerization reaction CQ: camphorquinone, A: tertiary amine.

    indicates a radical species. (1), (2), (3) and (n) represent the theoretical steps of linear monomer addition. n and m: large

    amount of monomer units. A trapped radical is illustrated in the termination section and segmental mobility is represented

    by a twisting gray arrow. A double gray arrow indicates radicals about to react together (bimolecular termination) during

    initiation, a free radical is generated through the activation of a photosensitive molecule or photoinitiator by a photon of

    specific wavelength. The newly formed free radical quickly reacts with a nearby monomer, opening its C C bond. Then

    follows a chain reaction, the propagation, by which the polymer grows by repetition of this C C opening by the radical and

    addition of a large number of monomer units one after the other. The polymer can either grow linearly by reaction with

    monomers (reactions 1, 2, 3, n), leaving unreacted or pendant double bonds, or grow tri-dimensionally by reacting with

    another polymer chain, creating cross-links. The addition of supplementary units continues until free radicals react with

    each other to form a stable covalent link through bimolecular termination reaction.

    indirectly evaluated by microhardness measurements (Vick-

    ers, Knoop) as a good linear correlation has generally been

    observed between DC and microhardness values [27,29,30].

    However, some disagree with this relationship as a general

    rule [31] because other factors than DC affect the micro-

    hardness measurement, notably the degree of crosslinking.

    In any case, microhardness measurement does not provide

    any quantitative information on the actual change in reactive

    groups.

    3.2. Degree ofcrosslinking

    Besides DC, the degree of crosslinking is another important

    determinant of mechanical properties [32] as well as struc-

    tural stability, since the lesscrosslinked the material,the more

    it swells and degrades in solvents [3335]. To our knowledge,

    there is currently no method to directly measure the degree

    of crosslinking. Indirectly, as increased crosslinking reduces

    molecular mobility, it can be highlighted by an increase of

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    6/18

    144 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    Fig. 3 Rp as a function of DC during a 70s real time

    measurement by FT-NIRS, for 70/30 mass%

    bis-GMA/TEGDMA resins, either unfilled (a) or filled (b) with

    75mass% fillers. Gray and black curves correspond

    respectively to MAPO-based and CQ based resins.

    Irradiation modes are indicated on the chart (time andirradiance). By plotting DC againstRp, the point in

    conversion where Rp approaches its maximum reveals the

    onset of severe diffusion limitation for propagation and

    approximately corresponds to vitrification. It appears

    clearly that the change of state is influenced by fillers,

    irradiation protocol and photoinitiator. It can also be

    observed that fillers reduce the maximum DC, and that

    MAPO-based resins reach significantly higher DC, even at

    such short irradiation time as 3 s (at 3000mW/cm2). CQ:

    camphorquinone, MAPO: monoacylphosphine oxide, Rp:

    Rate of polymerization, DC: degree of conversion.

    glass transition temperature (Tg), which can be measured

    either by dynamic mechanical analysis or differential scan-

    ning calorimetry. Tgrepresents the temperature at which the

    polymer vitrifies during the polymerization process (e.g. about

    47.5 C for a 50/50 mass% bis-GMA-TEGDMA mixture [36]), and

    above whicha polymer wouldreturn from a glassyto a rubbery

    state. Although Tg provides an indication of the crosslink-

    ing density, it also depends on DC and other factors such as

    monomer viscosity. It is therefore very difficult to distinguish

    between the effects of the different factors that contribute

    to crosslink density. Another indirect evaluation method pro-

    posed forcrosslinking density is theevaluation of thematerial

    softening in ethanol, i.e. the comparison of the surface hard-

    ness before and after ethanol storage [33,37].

    3.3. Mechanical properties

    From an engineering perspective, the flexural strength of the

    set material should be maximized, and its elastic modu-

    lus should remain similar to surrounding tissues in order toavoid inadequate stress-transfer on loading. Elastic modulus

    and flexural strength are generally measured by three-point

    flexural bending tests on 25 mm2 mm2mm samples,

    as specified by ISO 4049. However, its reliability in assess-

    ment of photopolymerization efficiency is questionable given

    that specimen preparation requires an overlapping curing

    regime to accommodate manufacture of the rectangular-

    shaped geometry [38]. To allow for consistent irradiation

    especially where the effect of polymerization characteris-

    tics on static mechanical properties are studied, single-shot

    curing protocols are employed as alternative tests with spec-

    imen geometries ensuring full sample coverage by the light

    tip, which include bi-axial flexure (discs) [39,40] and minithree-point flexure (bars) [41]. Although more clinically rel-

    evant geometries and uniform curing regimes are realized,

    the former test does not provide appropriate data for elastic

    modulus calculation and the latter may exhibit loading stress

    patterns that are not comparable with those of the standard

    ISO specification, which may result in non-comparable data.

    Another alternative would be the use of nanoscale measure-

    ment methods to evaluate mechanical properties locally in

    samples cured with single shot curing protocols. Such meth-

    ods have already been proved useful in the characterization of

    dental resin composites, notably nano-dynamic mechanical

    test [42,43], atomic force microscopy [30] or nanoindenta-

    tion [44]. Finally, the use of a rheometer seems promising forthe measurement of the initial viscoelastic modulus change

    of composites during curing, especially with respect to the

    buildup of shrinkage stress [45].

    3.4. Shrinkage and shrinkage stress

    Since the inception of dental resin-basedcomposites there has

    beena five-decade foray of research investigating the effects of

    filler content, morphology and resin chemistry on shrinkage

    [1,10]. As volumetric shrinkage has significant correlation to

    DC [22], both properties should always be measured in parallel

    when considering strategies to reduce shrinkage or shrink-

    age stress. This is rarely the case, but necessary to make surethat any reduction in shrinkage and/or stress is not due to an

    inferior DC [46]

    Although an important consideration, shrinkage prop-

    erties per se are not wholly responsible for interfacial gap

    formation and the clinical implication of lower shrinkage

    values for a particular resin composite may not necessar-

    ily be beneficial [6,47]. The generation of shrinkage stress

    throughout and following cure is not an intrinsic property of

    the material but a complex multi-factorial process affected by

    shrinkage, elastic modulus, polymerization rate, vitrification,

    and restrictions imposed to the composite, hence cavity

    configuration factor and compliance of bonded surfaces

    [6,46,48,49], whether regarding teeth in vivo or the test system

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    7/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 145

    for in vitro stress measurements. Altering compliance of the

    testing system will generate different stress values for the

    same material/conditions, which may account for the lack of

    comparable stress data in previous studies [50]. Interestingly,

    when comparing polymerization stress with four different

    mechanical testing systems, a good agreement was observed

    for experimental composites, but very few similarities for

    commercial composites [51], since not only their filler contentdiffers, but also potentially several other components as will

    be addressed further.

    3.5. Depth ofcure

    In order to ensure optimal material properties throughout the

    depth of a restoration, it is generally advocated to build them

    up using layering techniques, which can have the additional

    advantage of reducing cuspal deflection due to shrinkage

    stress [52,53]. The maximum thickness of each composite

    layer is often referred to as depth of cure. It is generally

    acknowledged that curing depth depends on the material aswell as the irradiation conditions (light type, light spectrum,

    irradiance, time, etc.). Depth of cure is usually assessed by

    scraping tests (ISO 4049) and penetrometer techniques [54]

    or by comparing the DC [55] or microhardness [56,57] at var-

    ious depths to the surface values. However, recent data has

    highlighted the insufficiency or inappropriateness of these

    methods, which may overestimate depth of cure. They are

    indeed unable to highlight the change of state of the resin

    matrix (glassy to rubbery) occurring at a certain depth, which

    correspondsto a major difference in resin modulus [30]. This is

    in line with previous work, showing that despite similar con-

    version in depth, increased resin softening can be observed

    after ethanol storage due to lower crosslinking [35]. Thesekey factors of cure depth measurements should be consid-

    ered, especially whenevaluating the newso-calledbulk cure

    composites [57].

    3.6. Trappedfree radicals

    The study of trapped free radicals is particularly relevant

    regarding dimethacrylate-based resins as their polymeriza-

    tion leads to highly crosslinked networks, where radical

    entrapment becomes the dominant termination pathway [17].

    Owing to thisfeature, freeradicals can be detecteddays, weeks

    or even months after initial photoirradiation, depending on

    the storage conditions, by means of electron paramagneticresonance (EPR) spectroscopy [18]. The height of specific

    EPR peaks can be used to determine the relative concentra-

    tion of trapped radicals [19]. A strong relationship between

    autoacceleration and radical entrapment was established [58],

    which highlights the importance of considering the amount of

    trappedradicals whenstudying the impact of photoirradiation

    protocols. The measurement of this parameter was used to

    better understand the impact of clinically relevant irradiation

    parameters [19]. EPR imaging was also recently demonstrated

    to be a reliable tool to study photoactive dimethacrylate-based

    dental resins [59], and provided complementary informa-

    tion to other techniques in the evaluation of depth of cure

    [30].

    3.7. Biocompatibility

    Unlike the other properties described above, biocompatibil-

    ity is not a direct way to assess the photopolymerization

    efficiency. However, it is an important prerequisite of all

    bio-materials, and should be considered, since a light-cured

    resin composite can affect the surrounding biological tis-

    sues in at least two ways: through the elution of bioactivespecies and by the temperature rise during the polymer-

    ization. First, regarding elution, an efficient polymerization

    process should prevent or at least limit leaching of poten-

    tially harmful components from the restoration material,

    which in case of dimethacrylate-based composites, concerns

    three main species: unconverted dimethacrylate monomers

    or oligomers, free radicals and photoinitiator molecules. The

    release of freemonomersfrom resin-based materials, which is

    inversely correlated to DC [24], can induce undesirable biolog-

    ical responses, such as tooth pulp damage, mucosal irritation,

    contact dermatitis and allergic reactions [60]. Low molecu-

    lar weight monomers such as TEGDMA and HEMA were also

    recently shown to be responsible for disturbances in the dif-ferentiation potential of dental stem cells[61]. Unpolymerized

    monomers also seem to promote bacterial colonization at

    the composite surface [62]. The release of hydroxyl radicals

    from dental resins has also been recently reported, possi-

    bly due to the reaction between oxygen and the unconverted

    methacrylate groups [63]. Reactive oxygen species, either

    released directly or created in situ by unconverted monomers,

    are responsible for oxidative stress and potentially cell struc-

    ture damage [64]. Finally, as photoinitiator molecules (CQ

    and amines) leach from the material [65], there exists some

    concern related to the biological consequences, such as the

    tertiary amine co-initiators, which may exhibit significant

    cytotoxicity [64]. Hence, it is necessary to modify materialsand optimize their photopolymerization conditions in order

    to reduce the amount of leachable species. Similarly, it seems

    reasonable to limit the temperature rise in the pulp cham-

    ber as much as possible during photopolymerization, in order

    to avoid damage to pulp tissues. Pulpal temperature rise is

    related to both the exotherm resulting from the composite

    polymerization andthe irradiation by thecuring light.The lat-

    ter is responsible for the majority of the increase in heat, and

    depends on time and on irradiance, which can be particularly

    high in modern light-curing units [66,67]. Although manufac-

    turers suggest that high irradiance and reduced curing time

    will not result in excessive temperature rise or damage to the

    pulp, the effect of thermal shock of such curing protocols ondental soft-tissue is not fully elucidated. Hence, any improve-

    ment in the material and/or photopolymerization process

    should be validated in terms of pulp thermal safety [67].

    4. Factors affecting photopolymerizationefficiency

    4.1. Intrinsicfactors

    4.1.1. Photoinitiator systems

    In dental composites, the most commonly used photoinitiator

    system is a combination of camphorquinone (CQ) and an

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    8/18

    146 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    electron donor, or co-initiator, which are generally differ-

    ent types of tertiary amines [68,69]. A key factor affecting

    photopolymerization efficiency of dimethacrylate resins is

    photoinitiator concentration. An increase in DC and hardness

    was observed with increased photoinitiator concentration

    [70,71], probably due to the increase of the maximum rate

    of polymerization [71]. Consequently, a shrinkage stress of

    higher magnitude may have been expected, although notobserved, whereby shrinkage stress was significantly influ-

    enced by conversion, but not by the rate of polymerization

    [71]. When CQ/amine concentrations are increased beyond a

    certain optimum value, a reduction of DC and hardness was

    observed, probably due to an excessive absorption of light

    in the superficial regions, resulting in less light transmission

    to the deeper layers, hence in suboptimal polymerization

    [70]. The type of co-initiator [72] as well as their ratio to the

    photoinitiator will influence the resulting polymerization

    quality: any increase above a certain optimal co-initiator level

    will not translate into an increase in polymerization efficiency

    [73] but it may affect clinically important parameters such

    as material yellowing[74] or biocompatibility, as amines mayexhibit significant cytotoxicity [64]. To date, no specific value

    exists for the optimal ratio of photoinitiator to co-initiator

    as large differences exist within the variants of resin-based

    composite components (i.e. monomer types and ratios, fillers

    types, morphology and ratio, pigment content, etc.). Hence,

    the concentration of both CQ and amine and their ratio need

    to be optimized for different properties. The use of ternary

    photoinitiator systems such as the addition of iodonium salts

    to CQ/amine, results in a substantial increase in polymeriza-

    tion rate, and superior DC, degree of crosslinking, mechanical

    properties and color stability [7578].

    Numerous photoinitiators have been considered as

    alternatives curing systems in dental composites, either toimprove esthetic quality by reducing CQ and its yellowing

    effect or because of significantly increased molar absorptiv-

    ity and thus, improved polymerization efficiency. Examples

    of alternative photoinitiators include phenylpropanedione

    (PPD), mono- or bis-acylphosphine oxides (MAPO and BAPO,

    respectively), benzoyl germanium or else benzil [7985].

    Some of these photoinitiators have already been identified

    in commercially available products [67,86,87]. The alterna-

    tive photoinitiators reviewed here all exhibit higher molar

    absorptivity (also called molar absorption coefficient or molar

    extinction coefficient , L mol1 cm1) compared with CQ

    (Table 1), which means that the probability for CQ to absorb

    light at the peak of its absorption range is much lowerthan for the other photoinitiators. Some also have a sig-

    nificantly higher quantum yield conversion (Table 1), which

    correspondsto the quantum yield of theinitiation step (Fig. 2).

    In other words, it is the ratio of the number of converted

    photoinitiators, i.e. those that have lost their absorption prop-

    erties, to the number of absorbed photons [88]. The value

    of 0.07 measured for CQ combined with dimethylaminoethyl

    methacrylate, expresses that the absorption of 14 photons

    is necessary for the conversion of one CQ molecule [88]. By

    contrast, much higher initiation quantum yield values were

    reported for MAPO (Table 1). Furthermore, each converted CQ

    molecule only generates one free radical that will actually ini-

    tiate polymerization. On the contrary, other photoinitiators

    are able to generate several active radicals per molecule, e.g.

    two for MAPO, and four for BAPO [80,89]. This explains the

    lower polymerization quantum yield of CQ (Table 1), which

    is the amount of monomer (or vinyl double bonds) polyme-

    rized per absorbed photon [80]. All these features justify the

    superiority of other systems, notably phosphine oxides, over

    CQ systems in terms of polymerization efficiency [84], but this

    also depends on the light source and irradiation parameters,since conventional LED curing units are not compatible with

    photoinitiators that exhibit peak absorptions at shorter wave-

    lengths (Section 4.2.1). However, the use of MAPO instead of

    CQ leads to higher DC, in shorter irradiation times, even as

    short as 3 s, and with a conventional halogen curinglamp [67].

    This higher efficiency of MAPO has not been confirmed for

    other properties yet, but preliminary results show a signifi-

    cant increase in elastic modulus, similar or improved flexural

    strength, andquite surprisinglygiven the reactionspeed,com-

    parable levels of shrinkage stress (unpublished data). The

    higher DC might also lead to reduced elution, which must also

    be confirmed. The use of Type I photoinitiators such as phos-

    phine oxides has other significant advantages, in that theydo not require the use of amines, which avoids optimizing

    photoinitiator and co-initiator ratio, but may alsoimprovebio-

    compatibility [64] and color stability [90], although there exists

    little evidence regarding the biocompatibility of the alterna-

    tive photoinitiators. Removal of tertiary amine co-initiators

    from light- or dual-cured resin-based materials within acidic

    and aqueous conditions would also prevent premature reac-

    tions or hindrance of effective polymerization in one-step

    self-adhesive systems, or contact of dual-cured materialswith

    the oxygen inhibitedlayerof simplifiedself-etching adhesives.

    Despite many advantages, phosphine oxides have been

    reported to lead to lower depth of cure [83,91,92]. However, in

    these studies, the light emission spectra were more adaptedto CQ than to MAPO (Section 4.2.1). Combinations of CQ and

    MAPO were also investigated, resulting in higher DC and poly-

    merization quantum yield compared to CQ alone, but lower

    than the values obtained with MAPO alone [91,93]. The latter

    is probably due to an energy transfer from a more efficient

    initiator (MAPO) to a less efficient one (CQ) [93]. Recently, an

    innovative approachhas beenproposed to overcome the depth

    of cure limitation of classical photoinitiator systems, by the

    use of phosphors. Thelatter produce blue light insidethe com-

    posite itself under the stimulation by near-infrared light [94].

    However, the up-conversion of NIR to blue light suffers signif-

    icant intensity loss so either unacceptable high laser power is

    required, or conversion is limited.In summary, it is clear that some photoinitiator systems

    have intrinsic advantages over the classical CQ system. How-

    ever, this should be considered in relation to extrinsic factors

    such as light spectrum, irradiance and irradiation time. This

    is particularly well exemplified by observations concerning

    commercial resin composites containing alternative photoini-

    tiators. Whilst recent work reports that single diode blue

    LED lights achieve similar degrees of polymerization than

    broadband (multiple diode) LED (Section 4.2.1) and halogen

    lights [95], the samples were light-cured for 40s, which is

    far longer than the manufacturers recommendation for the

    same light and material combination, i.e. usually 10s. A fur-

    ther study investigating a similar material/light-curing unit

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    9/18

    Table 1 Characteristics of the photoinitiators used in resin-based dental composites (CQ: camphorquinone, MAPO: monoacylphosoxide, PPD: phenylpropanedione, B: benzil, BTG: benzoyltrimethylgermane, DBTG: dibenzoyldiethylgermane).

    Photoinitiator UVvis absorption Quantumyields(mol/einstein)

    Quatum yield efficiency(photons:convertedphotoinitiator molecule)

    Numinitiradi

    Absorptionrange/max(nm)

    Molar extinctioncoefficient at max(Lmol1 cm1)

    CQ 400550/470

    28

    6.61103 0.07 (14:1) 140

    46

    MAPO 300430/381 520

    31.64103 0.35 (3:1)a 2660

    BAPO 300440/370 300 9.62103 0.10 (9:1)a 4

    PPD 300480/393 37 9.46103 0.10 (9.1)a 2

    B 300460/385 50 b b 2

    BTG

  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    10/18

    148 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    combination reportedlarge differences in compositehardness

    at shorter exposure time (10 and20 s) between single and mul-

    tiple diode LED lights, especially in depth. Besides, increased

    pulpal temperature rise went along with spectrum mismatch

    and excessive irradiation [67]. Besides extrinsic factors, other

    key material properties such as viscosity may significantly

    affect curing efficiency of a given photoinitiator compared to

    another.

    4.1.2. Viscosity, monomers andfillers

    There is now compelling evidence that initial resin viscosity is

    an important parameter in the reaction kinetics and final DC

    of dimethacrylate polymers, as it affects the mobility of each

    monomer, hence their reactivity [14]. Two main factors can

    affect local viscosity of the monomer system: first, monomer

    composition, and second, filler content.

    Variations of monomer molecular structure (di- or poly-

    methacrylates, molecular weight, molecule stiffness, etc.) and

    proportions can significantly affect polymerization efficiency.

    As regards the most common bis-GMA/TEGDMA mixture, it

    appears that the high reactivity is mainly related due tobis-GMA, which therefore largely controls the polymeriza-

    tion mechanisms and kinetics [14,96]. For pure bis-GMA, the

    maximum polymerization rate occurs at less than 5% con-

    version due to the very high viscosity, and the final DC is

    limited to about 30%. By contrast, for pure TEGDMA, which

    is far less viscous, the maximum rate is observed around

    22% conversion, with a final DC of over 60%, while the differ-

    ent co-monomer mixtures show intermediate values between

    these two extremes [96]. Based on these observations, it can

    be expected that any new dimethacrylate monomer affect-

    ing the viscosity differently can also potentially influence

    the polymerization efficiency. For example, ethoxylated ver-

    sions of bis-GMA(bis-EMA)and hyperbranchedpolymers(highmolecular weight tree-like molecules), which are used to

    reduce polymerization shrinkage, have a relatively low viscos-

    ity despite their high molecular weight [9799].

    Fillers have also a major impact on polymerization effi-

    ciency. The impact of fillers within a resin is a decrease in

    the maximum DC [27,83,100] (Fig. 3). Even at constant filler

    volume (56.7%) for a given resin formulation and differences

    in filler size and geometry resulted in significant differences

    in DC, from 48 to 61% [101]. Since DC was measured at the

    upper composite surface differences in light scattering as a

    consequence of filler morphology was unlikely [101]. More-

    over, changes in DC may be affected by differences in local

    monomer mobility, modulated by variations of the fillerresincontact area, whichcan lead to local changesin viscosityof the

    resin surrounding the fillers [102], and hence favor early vitri-

    fication. This is in line with the report of trapped free radical

    concentrations nearly 3 times higher in the organic matrix of

    a filled composite than in a corresponding unfilled resin [18].

    Another important aspect of fillers is that the lower the

    filler content, the lower the shrinkage and shrinkage stress,

    which has been demonstrated in experimental composites

    [51]. However, this is not verified in commercial composites,

    probably due to the broader range of shrinkage and elastic

    modulus values [51], as not only is filler content different,

    but also filler type, size and distribution; photoinitiator sys-

    tem, monomer type and co-monomer ratio may differ as well.

    However, recent research has suggested that the resin matrix

    has significantly more influence on polymerization shrink-

    age stress than filler content, and underlined the importance

    of reducing stress by modifying the resin chemistry with-

    out sacrificing filler content and DC [44]. There has been

    interest in reducing resin shrinkage and stress generation by

    using higher molecular weight resins. For example, hydro-

    genated dimer acids admixed withbis-GMAand UDMAexhibithigher conversion without significantly increasing shrinkage

    compared with conventional dimethacrylate resins [47,103].

    Also, the incorporation of reactive organic nanogel prepoly-

    mers into conventional dimethacrylate matrices have been

    reported to delay vitrification and modulus development by

    altering polymerization reaction rates to reduce stress without

    the expected concomitant decrease in polymer conversion or

    modulus associated with conventional methacrylate systems

    [104]. The use of chain-transfer agents such as hybrid thiol-

    ene/methacrylate systems can delay the mobility restriction

    through the creation of shorter polymer chains, which leads to

    an increase of theDC at whichvitrification takes place without

    concomitant increase in shrinkage stress [105,106]. Photo-inducedbond rearrangement using allyl sulfide functionalities

    [107] or else trithiocarbonate [108] are other promising strate-

    gies to reduce shrinkage stress. Finally, although this review

    focuses on dimethacrylates, it is worth mentioning that

    non-dimethacrylate chemistries (ring-opening mechanisms)

    exhibiting reduced shrinkage have also been developed [109].

    However, it is not clear whether they result in lower stress

    and better marginal adaptation, as contradictory results were

    found: low shrinkage stress was observed for these materi-

    als in an experimental setting with low C-factor (0.2) [110],

    and a high stress with higher C-factor (3) [111]; similarly, pos-

    itive [112] and negative [113] impact of these materials were

    reported regarding marginal adaptation.

    4.1.3. Opticalproperties

    The optical properties of a resin composite and their pho-

    topolymerization reaction are interdependent and material

    constituents affect the way light is transmitted through it,

    and therebyinfluence the polymerization quality, especially in

    depth. Several factors can limit light transmission through the

    composite. First, light reflection occurs at the surface [54,114].

    Second, light can be absorbed, either by pigments [31], which

    explains the lower depth of cure observed for darker and

    more opaque shades [115], or by photoinitiators [81,88]. In that

    regard, the higher the molar absorptivity, the lower the depth

    of cure [83]. Filler particles can also hinder light transmissionby scattering, which is dependent on the particle size and on

    the incident wavelength of the curing light, with increased

    scattering by filler particles dimensions approaching half of

    that wavelength [116118]. For example, increase in silica par-

    ticle size was shown to result in reduced microhardness and

    rate of polymerization in depth [119]. Fillers also hamper light

    transmission by refraction at the resinfiller interface, due

    to a mismatch between their refractive index [120,121]. The

    latter describes how light propagates through a medium; as

    light passes through the resinfiller interface, it is accompa-

    nied by a change in velocity and a directional change. This

    change depends on the medium, and is therefore affected by

    the monomer composition. For instance, resins formulated

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    11/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 149

    with a high percentage of diluent monomer showed the great-

    est change in refractive index (and polymerization shrinkage)

    [122]. In the same way, a lack of filler silane coating in com-

    posites cured under constrained conditions can generate gaps

    at the resinfiller interface, resulting in more light attenuation

    [123].

    Optical characteristics will also change along with the

    polymerization of the composite [122]. Upon irradiation,the photoinitiator will decompose/photobleach, and thereby

    increase light transmission, as observed for CQ-based sys-

    tems [124]. For the latter, the photobleaching rate is much

    slower than the polymerization rate andonly 20%of theinitial

    amount of CQ is consumed before the polymerization reac-

    tion is over [125]. Hence, photoinitiator concentrations should

    be optimized to avoid composite yellowing and cure depth

    reduction due to excess photoinitiator molecules. MAPO pho-

    tobleaching was reported to be somewhat counterbalanced

    by the production of colored oxidative by-products during

    the polymerization, possibly associated with high polymer-

    ization rate and reaction temperature [124]. These colored

    by-products can reduce light transmission in depth and affectesthetic quality, though MAPO-based composites still seem

    to have overall lower color values than CQ-based ones [84].

    Polymerization shrinkage occurring during light cure influ-

    ences light transmission as well, as it reduces the optical

    path length with the outcome of increased light transmis-

    sionthroughthe sample accordingto BeerLambertslaw [126].

    Finally, as the resin polymerizes, its refractive index increases

    due to the rapid buildup of crosslink density and viscosity;

    if it approaches that of the filler, interfacial filler/resin scat-

    tering is minimal and light transmission is increased [120].

    Hence, optimizing refractive index matching increases light

    transmission.

    4.2. Extrinsicfactors

    Within the literature, there exists sufficient evidence to sug-

    gest that the impact of light-curing on the polymerization

    process cannot be fully understood independently, but should

    be considered in relation with inherent material properties,

    which can have a major influence on the efficiency and suc-

    cess of photopolymerization.

    4.2.1. Light curing units and emission spectrum

    It can reasonably be stated that, based on technical develop-

    ments, the latest generations of LED curing lights are capable

    of competing with other curing light technologies, and evento outperform them in some aspects (cordless, lightweight,

    high irradiance, long service life, etc.) [127]. However, the

    importance of the light delivery through the curing tip on

    polymerization should not be overlooked. A rapid decrease in

    irradiance as distance increases has been highlighted for so-

    called, turbo light guides [87], and well-collimated straight

    light guides should therefore be given priority. There are also

    concerns regarding heterogeneity of cure over the surface,

    eitherdue to the use of an arrayof severaldiodes(withor with-

    out light guide) or to the way light is transmitted through the

    tip [87,128,129]. Theuniformity of thelight intensity should be

    improved, for example by using additional optical elements

    (mixing tube and diffusing screen), which despite causing a

    Fig. 4 Comparison of commonly used photoinitiators and

    commonly used halogen and LED lights and 3rd generation

    (or Polywave). The right axis represents the absolute

    irradiance of the light sources, i.e. XL: XL2500 (halogen;

    410<

  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    12/18

    150 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    regarding polymerization, but also to avoid excessive pulpal

    temperature rise, since any photon that is absorbed either

    by photoinitiators, pigmentsor reactionby-products, anddoes

    not initiate polymerization is transformed into heat [67,124].

    Finally, while inferior depthof cure was observed forMAPO- vs

    CQ-based composites (at similar filler content and equimolar

    photoinitiator concentration) [83], the effective irradiance of

    MAPO absorption range was much lower than for CQ, whichmight account for the lower depth of cure.

    Once spectral correspondence between light and mate-

    rial is optimized, a further question is then which irradiation

    parameters should be used: the level of irradiance, and irradi-

    ation time, or total energy.

    4.2.2. Radiant exposure, irradiance and irradiation time

    Radiant exposure (J/cm2) sometimes incorrectly, termed

    energy density [132] is the total amount of energy deliv-

    ered to a resin composite surface during the entire irradiation

    procedure. It is the product of light irradiance (mW/cm2) and

    irradiation time (s). A certain number of works support the

    fact that radiant exposure is the main determining factorof the material properties [133137]. Based on these obser-

    vations, the principle known as exposure reciprocity law

    states that comparable material properties can be achieved

    as long as radiant exposure is kept constant, no matter how

    it is obtained, by different combinations of irradiances and

    times. This encourages a tendency among dentists and man-

    ufacturers to use or suggest a high-power illumination to

    reduce curing time. However, the acceptance of this law, as

    a general rule, is contradicted by several other works. First,

    the increase in mechanical properties saturates above a cer-

    tain radiant exposure, and the threshold at which saturation

    occurs is property- and material-specific [138]. Second, it was

    suggested that despite DC, flexural strength, flexural modulusand depth of cureincreasing at higher radiant exposure, differ-

    ent combinations of time and irradiance can leadto significant

    differences in the material properties (Fig. 5) within a certain

    radiant exposure [41]. This is explained by the fact that there

    is probably a loss in radical growth centers by bimolecular ter-

    mination during the early stages of polymerization wherein

    system mobility is high. This loss (or early termination) is

    greater for high irradiance protocols, since at low conversion,

    termination is proportional to the squared concentration of

    free radicals [19,139].

    As highlighted earlier, polymerization efficiency is inti-

    mately linked to intrinsic factors (Fig. 1). Therefore, it makes

    sense that they also affect the applicability of the law, as itwas recently demonstrated [83]. While large differences in

    DC (4472%) were observed between different irradiance/time

    combinations (constant radiant exposure) in a 50/50wt%

    bis-GMA/TEGDMA unfilled resin, the differences were small

    (5158%) when the same resins were filled. Similar observa-

    tions have been reportedin unfilled resinsif theresin viscosity

    is increased, where reciprocity holds quite well for60/40 wt%

    bis-GMA/TEGDMA ratios, but large differences in DC were

    reported for less viscous resins [140]. In this regard, though

    low molecular weight monomers such as TEGDMA are gener-

    ally assumed to reach higher DC [96], it is generally based on

    polymerization at low irradiance (0.42.9 mW/cm2) and long

    irradiation time (8min). However, this is not representative

    of the clinical conditions in which these materials are used,

    and it is not verified in all irradiation conditions. For instance,

    DC of a 70/30 wt% bis-GMA/TEGDMA mixture is higher than

    with 40/60 wt% bis-GMA/TEGDMA (70 vs 55%, respectively)

    when cured at 50mW/cm2 during 5s [140]. Regarding pho-

    toinitiators, it was previously reported that no significant

    difference in DC was observed between CQ-based and PPD-

    based [141] or MAPO-based resins [142]. However, these worksare generally based on a single light irradiance, and changes

    of irradiation time/irradiance combinations can affect the

    results greatly, as illustrated in Fig. 5. While exposure reci-

    procity law for DC is generally upheld for MAPO-based resins,

    even at low viscosity, it is not the case at all for CQ- and

    PPD-based resins. Importantly, even in the cases when reci-

    procity law is upheld for one property (DC in general), it is

    not necessarily the case for others key material properties

    [41,83,138].

    In summary, even if exposure reciprocity law holds for

    some materials for some properties, the important point is

    that is does not for others, and it can therefore not be con-

    sidered as a general rule. This is especially important givenmanufacturers reluctance to divulge (in some cases due to

    proprietary components) accuratecomposition of commercial

    resin composites. As highlighted by comparing the reactivity

    of paste composites and their parent flowable formulations,

    unknown differences in either co-monomer composition,

    filler content and/or photoinitiators can result in substantially

    reduced DC for high irradiance curing parameters [143]. Con-

    sequently, when evaluating new monomer technologies, such

    as the hyperbranched polymers, nanogels, thiol-enes men-

    tioned earlier, it is difficult to strictly conclude that these

    systems lead to improved properties, since in most develop-

    mental studies, they are usually polymerized with a single

    irradiation protocol, sometimes not clinically relevant (e.g.50mW/cm2 during 5 min [104]). Hence, it would be appro-

    priate to test any composite innovation for a large range of

    irradiances and irradiation times.

    4.2.3. Irradiation modes

    In order to reduce stress transmission at the bonded inter-

    face, besides modifications of the material, soft-start curing

    protocols (ramp, step or pulse-delay modes) were pro-

    posed to delay the onset of polymer gelation, and thereby

    contribute to shrinkage stress reduction. Certain authors

    explained this effect by a lower conversion and/or infe-

    rior crosslinking [36,37,144146]. However, others observed

    that soft-start regimen had the potential to reduce shrink-age stress, while keeping DC and/or mechanical properties

    constant [42,147149]. Despite an important number of pub-

    lications on this subject, there is still no definitive answer as

    whether or not soft-start modes are beneficial. Again, this is

    probably due to the differences in composition of the various

    resin-composites used in the different studies, which proba-

    bly affect the efficiency of the soft-start curing modes and the

    resulting properties.

    4.2.4. Temperature

    The temperature during polymerization can significantly

    affect polymerization efficiency, as a rise from room tempera-

    ture (22 C) to mouth temperature (35 C) results in increased

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    13/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 151

    Fig. 5 Comparison of DC (measured in real-time by FT-NIR) for 50/50 bis-GMA/TEGDMA unfilled resins containing

    commonly used photoinitiators, CQ, MAPO and PPD, the latter with (+) and without () amine co-initiator. They were cured

    at similar radiant exposure (18J/cm2) but different times and irradiances, as indicated on the right side on the figure. As it

    can be observed, MAPO-based resins display the highest DC, are less dependent upon time/irradiance combination.

    CQ-based resins reached higher DC than PPD(+), both varying importantly depending on time/irradiance combination.

    PPD() was almost inefficient to initiate polymerization, which means that PPD requires an amine as co-initiator, in

    accordance with [85].

    DC (610%), hardness and rate of polymerization [150], due

    to improved monomer mobility, allowing more of the reac-

    tion to occur prior to vitrification [151]. This involves first that

    temperature stability is important when studying resin-based

    composites, as in any experiment, in order to avoid any arti-

    fact due to temperature. Second, it highlights the interest of

    preheating composites, not only for practical purposes (better

    handling), but also to improve conversion and reduce curingtime [152,153].

    An excessive temperature rise might affect pulp vitality as

    well as create damage to gingival tissues [154]. Hence, man-

    ufacturers and clinicians should be aware that temperature

    rise is energy-related, and therefore, the curing parameters

    should be optimized to ensure improved material properties,

    while avoiding pulp overheating[67,131], which is more haz-

    ardous with thinner dentinal tissue coverage above the pulp

    [155].

    4.2.5. Light guide tippositioning

    The adequate positioning of the light guide tip can signifi-cantly affect the energy received by the RBC, and thereby the

    quality of its polymerization. First, the irradiance decreases

    with the increase of the distance between tip and restora-

    tion [156], resulting in reduced DC [157]. This distance is

    partly governed by the cavity depth (roughly between 2 and

    7mm), but practitioners should also make sure that the tip

    is placed as close as possible to the material. Second, the

    light guide tip should be placed perpendicular to the mate-

    rial surface to optimize light transmission in depth, and its

    position should be stabilized during the irradiation proce-

    dure, since a lack of control of the tip positioning during the

    procedure can result in a decrease of the energy delivered

    [158].

    5. Conclusion

    Photopolymerization conditions have a major impact on the

    final mechanical, physical and biological properties of resin

    composite materials. Hence, it is clear that dentists play

    a significant part in the quality of their restoration by the

    appropriate choice of curing conditions. However, this reviewhighlights that the proper selection of extrinsic curing factors

    is intimately connected with inherent material characteris-

    tics, which are often unknown or, at best, vaguely identified

    by the clinicians. In addition, even if the exact product com-

    position was disclosed, all factors discussed above affect a

    given material property either positively or negatively, making

    it impossible to predict the resulting polymerization quality.

    In light of this, it is clear that there is a critical need for better

    information frommanufacturers on their productsin order for

    dentists to be able to knowingly adapt and optimize the use

    of resin-based composites in their daily practice. Hence, this

    review calls for a significant change in thefuture regarding the

    way instructions for use are provided. For each new mate-rial appearing on the market, some kind of identity card

    should be required, providing essential information such as

    the absorption spectrum, and the impact of various irradi-

    ance/time combination on the principal material properties,

    e.g. in the form of contour maps [138]. Such situation where

    dentists make a decision based on very little information does

    not correspond to the currently advocated trend toward a

    practice of our medical discipline driven by science and evi-

    dence.

    Acknowledgment

    JL is a Postdoctoral Researcher at F.R.S.-FNRS.

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    14/18

    152 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    r e f e r ence s

    [1] Ferracane JL. Resin composite state of the art. DentalMaterials 2011;27:2938.

    [2] Van Meerbeek B, Yoshihara K, Yoshida Y, Mine A, De MunckJ, Van Landuyt KL. State of the art of self-etch adhesives.

    Dental Materials 2011;27:1728.[3] Van Nieuwenhuysen JP, DHoore W, Carvalho J, Qvist V.Long-term evaluation of extensive restorations inpermanent teeth. Journal of Dentistry 2003;31:395405.

    [4] Opdam NJ, Bronkhorst EM, Roeters JM, Loomans BA. Aretrospective clinical study on longevity of posteriorcomposite and amalgam restorations. Dental Materials2007;23:28.

    [5] Opdam NJ, Bronkhorst EM, Loomans BA, Huysmans MC.12-Year survival of composite vs. amalgam restorations.

    Journal of Dental Research 2010;89:10637.[6] Tantbirojn D, Pfeifer CS, Braga RR, Versluis A. Do low-shrink

    composites reduce polymerization shrinkage effects?Journal of Dental Research 2011;90:596601.

    [7] Leprince J, Palin WM, Mullier T, Devaux J, Vreven J, Leloup

    G. Investigating filler morphology and mechanicalproperties of new low-shrinkage resin composite types.Journal of Oral Rehabilitation 2010;37:36476.

    [8] Schmalz G. Resin-based composites. In: Schmalz G,Arenholt-Bindslev D, editors. Biocompatibility of DentalMaterials. Berlin/Heidelberg/Germany: Springer-Verlag;2009. p. 99137.

    [9] Leprince JG, Zeitlin BD, Tolar M, Peters OA. Interactionsbetween immune system and mesenchymal stem cells indental pulp and periapical tissues. InternationalEndodontic Journal 2012;45:689701.

    [10] Cramer NB, Stansbury JW, Bowman CN. Recent advancesand developments in composite dental restorativematerials. Journal of Dental Research 2011;90:40216.

    [11] Ilie N, Hickel R. Resin composite restorative materials.

    Australian Dental Journal 2011;56(Suppl. 1):5966.[12] Chen MH. Update on dental nanocomposites. Journal of

    Dental Research 2010;89:54960.[13] Anseth KS, Newman SM, Bowman CN. Polymeric dental

    composites: properties and reaction behavior ofmultimethacrylate dental restorations. Advances inPolymer Science 1995;122:177217.

    [14] Lovell LG, Stansbury JW, Syrpes DC, Bowman CN. Effects ofcomposition and reactivity on the reaction kinetics ofdimethacrylate/dimethacrylate copolymerizations.Macromolecules 1999;32:391321.

    [15] Truffier-Boutry D, Demoustier-Champagne S, Devaux J,Biebuyck JJ, Mestdagh M, Larbanois P, et al. Aphysico-chemical explanation of the post-polymerizationshrinkage in dental resins. Dental Materials 2006;22:40512.

    [16] Decker C, Elzaouk B. Photopolymerization ofmultifunctional monomers. 7. Evaluation of thepropagation rate-constant and the terminationrate-constant. European Polymer Journal 1995;31:115563.

    [17] Andrzejewska E. Photopolymerization kinetics ofmultifunctional monomers. Progress in Polymer Science2001;26:60565.

    [18] Leprince J, Lamblin G, Truffier-Boutry D,Demoustier-Champagne S, Devaux J, Mestdagh M, et al.Kinetic study of free radicals trapped in dental resinsstored in different environments. Acta Biomaterialia2009;5:251824.

    [19] Leprince JG, Lamblin G, Devaux J, Dewaele M, Mestdagh M,Palin WM, et al. Irradiation modes impact on radicalentrapment in photoactive resins. Journal of Dental

    Research 2010;89:14948.

    [20] Ferracane JL, Greener EH. The effect of resin formulation onthe degree of conversion and mechanical properties ofdental restorative resins. Journal of Biomedical MaterialsResearch 1986;20:12131.

    [21] Li J, Li H, Fok ASL, Watts DC. Multiple correlations ofmaterial parameters of light-cured dental composites.Dental Materials 2009;25:82936.

    [22] Dewaele M, Truffier-Boutry D, Devaux J, Leloup G. Volume

    contraction in photocured dental resins: theshrinkageconversion relationship revisited. DentalMaterials 2006;22:35965.

    [23] Ferracane JL, Mitchem JC, Condon JR, Todd R. Wear andmarginal breakdown of composites with various degrees ofcure. Journal of Dental Research 1997;76:150816.

    [24] Ferracane JL. Elution of leachable components fromcomposites. Journal of Oral Rehabilitation 1994;21:44152.

    [25] Ferracane JL, Greener EH. Fourier transform infraredanalysis of degree of polymerization in unfilled resins methods comparison. Journal of Dental Research1984;63:10935.

    [26] Pianelli C, Devaux J, Bebelman S, Leloup G. Themicro-Raman spectroscopy, a useful tool to determine thedegree of conversion of light-activated composite resins.

    Journal of Biomedical Materials Research 1999;48:67581.[27] Halvorson RH, Erickson RL, Davidson CL. The effect of filler

    and silane content on conversion of resin-basedcomposite. Dental Materials 2003;19:32733.

    [28] Stansbury JW, Dickens SH. Determination of double bondconversion in dental resins by near infrared spectroscopy.Dental Materials 2001;17:719.

    [29] Ferracane JL. Correlation between hardness and degree ofconversion during the setting reaction of unfilled dentalrestorative resins. Dental Materials 1985;1:114.

    [30] Leprince JG, Leveque P, Nysten B, Gallez B, Devaux J, LeloupG. New insight into the depth of cure ofdimethacrylate-based dental composites. Dental Materials2012;28:51220.

    [31] Musanje L, Darvell BW. Curing-light attenuation infilledresin restorative materials. Dental Materials2006;22:80417.

    [32] Tamareselvy K, Rueggeberg FA. Dynamic mechanicalanalysis of two crosslinked copolymer systems. DentalMaterials 1994;10:2907.

    [33] Asmussen E, Peutzfeldt A. Influence of selectedcomponents on crosslink density in polymer structures.European Journal of Oral Sciences 2001;109:2825.

    [34] Benetti AR, Asmussen E, Munksgaard EC, Dewaele M,Peutzfeldt A, Leloup G, et al. Softening and elution ofmonomers in ethanol. Dental Materials 2009;25:100713.

    [35] Asmussen E, Peutzfeldt A. Polymer structure of alight-cured resin composite in relation to distance from thesurface. European Journal of Oral Sciences 2003;111:2779.

    [36] Dewaele M, Asmussen E, Peutzfeldt A, Munksgaard EC,Benetti AR, Finn G, et al. Influence of curing protocol onselected properties of light-curing polymers: Degree ofconversion, volume contraction, elastic modulus, and glasstransition temperature. Dental Materials 2009;25:157684.

    [37] Asmussen E, Peutzfeldt A. Two-step curing: influence onconversion and softening of a dental polymer. DentalMaterials 2003;19:46670.

    [38] Palin WM, Fleming GJ, Marquis PM. The reliability ofstandardized flexure strength testing procedures for alight-activated resin-based composite. Dental Materials2005;21:9119.

    [39] Rueggeberg FA, Cole MA, Looney SW, Vickers A, Swift EJ.Comparison of manufacturer-recommended exposuredurations with those determined using biaxial flexure

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    15/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 153

    strength and scraped composite thickness among a varietyof light-curing units: Masters of esthetic dentistry. JournalofEsthetic and Restorative Dentistry 2009;21:4361.

    [40] Curtis AR, Palin WM, Fleming GJ, Shortall AC, Marquis PM.The mechanical properties of nanofilled resin-basedcomposites: the impact of dry and wet cyclic pre-loadingon bi-axial flexure strength. Dental Materials2009;25:18897.

    [41] Peutzfeldt A, Asmussen E. Resin composite properties andenergy density of light cure. Journal of Dental Research2005;84:65962.

    [42] Ilie N, Jelen E, Hickel R. Is the soft-start polymerisationconcept still relevant for modern curing units? Clinical OralInvestigations 2011;15:219.

    [43] Ilie N, Hickel R, Watts DC. Spatial and cure-timedistribution of dynamic-mechanical properties of adimethacrylate nano-composite. Dental Materials2009;25:4118.

    [44] Goncalves F, Azevedo CLN, Ferracane JL, Braga RR.Bis-GMA/TEGDMA ratio and filler content effects onshrinkage stress. Dental Materials 2011;27:5206.

    [45] Kim MH, Min SH, Ferracane J, Lee IB. Initial dynamicviscoelasticity change of composites during light curing.

    Dental Materials 2010;26:46370.[46] Braga RR, Ballester RY, Ferracane JL. Factors involved in the

    development of polymerization shrinkage stress inresin-composites: a systematic review. Dental Materials2005;21:96270.

    [47] Boaro LCC, Gonalves F, Guimaraes TC, Ferracane JL,Versluis A, Braga RR. Polymerization stress, shrinkage andelastic modulus of current low-shrinkage restorativecomposites. Dental Materials 2010;26:114450.

    [48] Arakawa K. Shrinkage forces due to polymerization oflight-cured dental composite resin in cavities. PolymerTesting 2010;29:10526.

    [49] Koplin C, Jaeger R, Hahn P. A material model for internalstress of dental composites caused by the curing process.Dental Materials 2009;25:3318.

    [50] Meira JB, Braga RR, Ballester RY, Tanaka CB, Versluis A.Understanding contradictory data in contraction stresstests. Journal of Dental Research 2011;90:36570.

    [51] Goncalves F, Boaro LC, Ferracane JL, Braga RR. Acomparative evaluation of polymerization stress dataobtained with four different mechanical testing systems.Dental Materials 2012;28:6806.

    [52] Park J, Chang J, Ferracane J, Lee IB. How should compositebe layered to reduce shrinkage stress: incremental or bulkfilling? Dental Materials 2008;24:15015.

    [53] Kwon Y, Ferracane J, Lee IB. Effect of layering methods,composite type, and flowable liner on the polymerizationshrinkage stress of light cured composites. DentalMaterials 2012;28:8019.

    [54] Shortall AC, Wilson HJ, Harrington E. Depth of cure ofradiation-activated composite restoratives influence ofshade and opacity. Journal of Oral Rehabilitation1995;22:33742.

    [55] Beun S, Glorieux T, Devaux J, Vreven J, Leloup G.Characterization of nanofilled compared to universal andmicrofilled composites. Dental Materials 2007;23:519.

    [56] Czasch P, Ilie N. In vitro comparison of mechanicalproperties and degree of cure of bulk fill composites.Clinical Oral Investigations 2012,http://dx.doi.org/10.1007/s00784-012-0702-8[Epub ahead ofprint].

    [57] Flury S, Hayoz S, Peutzfeldt A, Husler J, Lussi A. Depth ofcure of resin composites: is the ISO 4049 method suitablefor bulk fill materials? Dental Materials 2012;28:5218.

    [58] Zhu S, Tian Y, Hamielec AE, Eaton DR. Radicalconcentrations in free-radical copolymerization ofMMA/EGDMA. Polymer 1990;31:1549.

    [59] Leveque P, Leprince JG, Bebelman S, Devaux J, Leloup G,Gallez B. Spectral spatial electron paramagnetic resonanceimaging as a tool to study photoactivedimethacrylate-based dental resins. Journal of MagneticResonance 2012;220:4553.

    [60] Van Landuyt KL, Nawrot T, Geebelen B, De Munck J,Snauwaert J, Yoshihara K, et al. How much do resin-baseddental materials release? A meta-analytical approach.Dental Materials 2011;27:72347.

    [61] Geurtsen W, Bakopoulou A, Leyhausen G, Volk J, TsiftsoglouA, Garefis P, et al. Effects of HEMA and TEDGMA on thein vitro odontogenic differentiation potential of humanpulp stem/progenitor cells derived from deciduous teeth.Dental Materials 2011;27:60817.

    [62] Brambilla E, Gagliani M, Ionescu A, Fadini L, Garcia-GodoyF. The influence of light-curing time on the bacterialcolonization of resin composite surfaces. Dental Materials2009;25:106772.

    [63] Lamblin G, Leprince J, Devaux J, Mestdagh M, Gallez B,Leloup G. Hydroxyl radical release from dental resins:

    electron paramagnetic resonance evidence. ActaBiomaterialia 2010;6:31938.

    [64] Bakopoulou A, Papadopoulos T, Garefis P. Moleculartoxicology of substances released from resin-based dentalrestorative materials. International Journal of MolecularSciences 2009;10:386199.

    [65] Durner J, Spahl W, Zaspel J, Schweikl H, Hickel R, Reichl FX.Eluted substances from unpolymerized and polymerizeddental restorative materials and their Nernst partitioncoefficient. Dental Materials 2010;26:919.

    [66] Guiraldo RD, Consani S, Sinhoreti MAC, Correr-Sobrinho L,Schneider LFJ. Thermal variations in the pulp chamberassociated with composite insertion techniques andlight-curing methods. Journal of Contemporary DentalPractice 2009;10:01724.

    [67] Leprince J, Devaux J, Mullier T, Vreven J, Leloup G.Pulpal-temperature rise and polymerization efficiency ofLED curing lights. Operative Dentistry 2010;35:22030.

    [68] Cook WD. Photopolymerization kinetics of dimethacrylatesusing the camphorquinone amine initiator system.Polymer 1992;33:6009.

    [69] Jakubiak J, Allonas X, Fouassier JP, Sionkowska A,Andrzejewska E, Linden LA. Camphorquinone-aminesphotoinitating systems for the initiation of free radicalpolymerization. Polymer 2003;44:521926.

    [70] Musanje L, Ferracane JL, Sakaguchi RL. Determination ofthe optimal photoinitiator concentration in dentalcomposites based on essential material properties. DentalMaterials 2009;25:9941000.

    [71] Pfeifer CS, Ferracane JL, Sakaguchi RL, Braga RR.Photoinitiator content in restorative composites: influenceon degree of conversion, reaction kinetics, volumetricshrinkage and polymerization stress. American Journal ofDentistry 2009;22:20610.

    [72] Furuse AY, Mondelli J, Watts DC. Network structures ofbis-GMA/TEGDMA resins differ in DC, shrinkage-strain,hardness and optical properties as a function of reducingagent. Dental Materials 2011;27:497506.

    [73] Yoshida K, Greener EH. Effects of two amine reducingagents on the degree of conversion and physical propertiesof an unfilled light-cured resin. Dental Materials1993;9:24651.

    [74] Schneider LFJ, Cavalcante LM, Consani S, Ferracane JL.Effect of co-initiator ratio on the polymer properties of

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1007/s00784-012-0702-8http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1007/s00784-012-0702-8http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    16/18

    154 d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156

    experimental resin composites formulated withcamphorquinone and phenyl-propanedione. DentalMaterials 2009;25:36975.

    [75] Cook WD, Chen F. Enhanced photopolymerization ofdimethacrylates with ketones, amines, and iodonium salts:the CQ system. Journal of Polymer Science Part A: PolymerChemistry 2011;49:503041.

    [76] Park J, Ye Q, Topp EM, Misra A, Kieweg SL, Spencer P. Effect

    ofphotoinitiator system and water content on dynamicmechanical properties of a light-cured bisGMA/HEMAdental resin. Journal of Biomedical Materials Research Part A 2010;93:124551.

    [77] Shin DH, Rawls HR. Degree of conversion and color stabilityofthe light curing resin with new photoinitiator systems.Dental Materials 2009;25:10308.

    [78] Ogliari FA, Ely C, Petzhold CL, Demarco FF, Piva E. Oniumsalt improves the polymerization kinetics in anexperimental dental adhesive resin. Journal of Dentistry2007;35:5837.

    [79] Neumann MG, Miranda WG, Schmitt CC, Rueggeberg FA,Correa IC. Molar extinction coefficients and the photonabsorption efficiency of dental photoinitiators and lightcuring units. Journal of Dentistry 2005;33:

    52532.[80] Neumann MG, Schmitt CC, Ferreira GC, Correa IC. The

    initiating radical yields and the efficiency of polymerizationfor various dental photoinitiators excited by different lightcuring units. Dental Materials 2006;22:57684.

    [81] Ogunyinka A, Palin WM, Shortall AC, Marquis PM.Photoinitiation chemistry affects light transmission anddegree of conversion of curing experimental dental resincomposites. Dental Materials 2007;23:80713.

    [82] Moszner N, Fischer UK, Ganster B, Liska R, Rheinberger V.Benzoyl germanium derivatives as novel visible lightphotoinitiators for dental materials. Dental Materials2008;24:9017.

    [83] Leprince JG, Hadis M, Shortall AC, Ferracane JL, Devaux J,Leloup G, et al. Photoinitiator type and applicability of

    exposure reciprocity law in filled and unfilled photoactiveresins. Dental Materials 2011;27:15764.

    [84] Arikawa H, Takahashi H, Kanie T, Ban S. Effect of variousvisible light photoinitiators on the polymerization andcolor of light-activated resins. Dental Materials Journal2009;28:45460.

    [85] Asmussen S, Vallo C. Light absorbing products duringpolymerization of methacrylate monomers photoinitiatedwith phenyl-1,2-propanedione/amine. Journal ofPhotochemistry and Photobiology A: Chemistry2009;202:22834.

    [86] Kameyama A, Hatayama H, Kato J, Haruyama A, Teraoka H,Takase Y, et al. Spectral characteristics of light-curing unitsand dental adhesives. Journal of Photopolymer Science andTechnology 2011;24:4116.

    [87] Price RBT, Felix CA. Effect of delivering light in specificnarrow bandwidths from 394 to 515 nm on themicro-hardness of resin composites. Dental Materials2009;25:899908.

    [88] Chen YC, Ferracane JL, Prahl SA. Quantum yield ofconversion of the photoinitiator camphorquinone. DentalMaterials 2007;23:65564.

    [89] Decker C. Kinetic study and new applications of UVradiation curing. Macromolecular Rapid Communications2002;23:106793.

    [90] Alvim HH, Alecio AC, Vasconcellos WA, Furlan M, deOliveira JE, Saad JR. Analysis of camphorquinone incomposite resins as a function of shade. Dental Materials2007;23:12459.

    [91] Miletic V, Santini A. Micro-Raman spectroscopic analysis ofthe degree of conversion of composite resins containingdifferent initiators cured by polywave or monowave LEDunits. Journal of Dentistry 2012;40:10613.

    [92] Schneider LFJ, Cavalcante LM, Prahl SA, Pfeifer CS,Ferracane JL. Curing efficiency of dental resin compositesformulated with camphorquinone ortrimethylbenzoyl-diphenyl-phosphine oxide. Dental

    Materials 2012;28:3927.[93] Neumann MG, Schmitt CC, Horn MA. The effect of the

    mixtures of photoinitiators in polymerization efficiencies.Journal of Applied Polymer Science 2009;112:12934.

    [94] Stepuk A, Mohn D, Grass RN, Zehnder M, Kramer KW, PelleF, et al. Use of NIR light and upconversion phosphors inlight-curable polymers. Dental Materials 2012;28:30411.

    [95] Sim JS, Seol HJ, Park JK, Garcia-Godoy F, Kim HI, Kwon YH.Interaction of LED light with coinitiator-containingcomposite resins: Effect of dual peaks. Journal of Dentistry2012;40:83642.

    [96] Lovell LG, Newman SM, Bowman CN. The effects of lightintensity, temperature, and comonomer composition onthe polymerization behavior of dimethacrylate dentalresins. Journal of Dental Research 1999;78:146976.

    [97] Ogliari FA, Ely C, Zanchi CH, Fortes CB, Samuel SM,Demarco FF, et al. Influence of chain extender length ofaromatic dimethacrylates on polymer networkdevelopment. Dental Materials 2008;24:16571.

    [98] Zhu X, Zhou Y, Yan D. Influence of branching architectureon polymer properties. Journal of Polymer Science, Part B:Polymer Physics 2011;49:127786.

    [99] Dewaele M, Leprince JG, Fallais I, Devaux J, Leloup G.Benefits and limitations of adding hyperbranchedpolymers to dental resins. Journal of Dental Research2012;91(12):117883.

    [100] Garoushi S, Vallittu PK, Watts DC, Lassila LV. Effect ofnanofiller fractions and temperature on polymerizationshrinkage on glass fiber reinforced filling material. DentalMaterials 2008;24:60610.

    [101] Turssi CP, Ferracane JL, Vogel K. Filler features and theireffects on wear and degree of conversion of particulatedental resin composites. Biomaterials 2005;26:49327.

    [102] Beun S, Bailly C, Dabin A, Vreven J, Devaux J, Leloup G.Rheological properties of experimental bis-GMA/TEGDMAflowable resin composites with variousmacrofiller/microfiller ratio. Dental Materials2009;25:198205.

    [103] Trujillo-Lemon M, Ge J, Lu H, Tanaka J, Stansbury JW.Dimethacrylate derivatives of dimer acid. Journal ofPolymer Science, Part A: Polymer Chemistry 2006;44:39219.

    [104] Moraes RR, Garcia JW, Barros MD, Lewis SH, Pfeifer CS, Liu J,et al. Control of polymerization shrinkage and stress innanogel-modified monomer and composite materials.Dental Materials 2011;27:50919.

    [105] Pfeifer CS, Wilson ND, Shelton ZR, Stansbury JW. Delayedgelation through chain-transfer reactions: mechanism forstress reduction in methacrylate networks. Polymer2011;52:3295303.

    [106] Boulden JE, Cramer NB, Schreck KM, Couch CL,Bracho-Troconis C, Stansbury JW, et al.Thiol-ene-methacrylate composites as dental restorativematerials. Dental Materials 2011;27:26772.

    [107] Park HY, Kloxin CJ, Scott TF, Bowman CN. Covalentadaptable networks as dental restorative resins: stressrelaxation by addition-fragmentation chain transfer in allylsulfide-containing resins. Dental Materials 2010;26:10106.

    [108] Leung D, Bowman CN. Reducing shrinkage stress ofdimethacrylate networks by reversible

    http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005http://localhost/var/www/apps/conversion/tmp/scratch_10/dx.doi.org/10.1016/j.dental.2012.11.005
  • 8/10/2019 Progress in Dimethacrylate-based Dental Composite Technology and Curing Efficiency

    17/18

    d en ta l mat er ia ls 2 9 ( 2 0 1 3 ) 139156 155

    addition-fragmentation chain transfer. MacromolecularChemistry and Physics 2012;213:198204.

    [109] Weinmann W, Thalacker C, Guggenberger R. Siloranes indental composites. Dental Materials 2005;21:6874.

    [110] Gao BT, Lin H, Zheng G, Xu YX, Yang JL. Comparisonbetween a silorane-based composite andmethacrylate-based composites: shrinkage characteristics,thermal properties, gel point and vitrification point. Dental

    Materials Journal 2012;31:7685.[111] Boaro LC, Goncalves F, Guimaraes TC, Ferracane JL, Versluis

    A, Braga RR. Polymerization stress, shrinkage and elasticmodulus of current low-shrinkage restorative composites.Dental Materials 2010;26:114450.

    [112] Palin WM, Fleming GJ, Nathwani H, Burke FJ, Randall RC.In vitro cuspal deflection and microleakage of maxillarypremolars restored with novel low-shrink dentalcomposites. Dental Materials 2005;21:32435.

    [113] Van Ende A, De Munck J, Mine A, Lambrechts P, VanMeerbeek B. Does a low-shrinking composite induce lessstress at the adhesive interface? Dental Materials2010;26:21522.

    [114] Watts DC, Cash AJ. Analysis of optical transmission by400500 nm visible light into aesthetic dental biomaterials.

    Journal of Dentistry 1994;22:1127.[115] Shortall AC. How light source and product shade influence

    cure depth for a contemporary composite. Journal of OralRehabilitation 2005;32:90611.

    [116] Emami N, Sjodahl M, Soderholm KJ. How filler properties,filler fraction, sample thickness and light source affect lightattenuation in particulate filled resin composites. DentalMaterials 2005;21:72130.

    [117] dos Santos GB, Alto RV, Filho HR, da Silva EM, Fellows CE.Light transmission on dental resin composites. DentalMaterials 2008;24:5716.

    [118] Masotti AS, Onofrio AB, Conceicao EN, Spohr AM. UVvisspectrophotometric direct transmittance analysis ofcomposite resins. Dental Materials 2007;23:72430.

    [119] Fujita K, Ikemi T, Nishiyama N. Effects of particle size of

    silica filler on polymerization conversion in a light-curingresin composite. Dental Materials 2011;27:107985.

    [120] Shortall AC, Palin WM, Burtscher P. Refractive indexmismatch and monomer reactivity influence compositecuring depth. Journal of Dental Research 2008;87:848.

    [121] Howard B, Wilson ND, Newman SM, Pfeifer CS, StansburyJW. Relationships between conversion, temperature andoptical properties during composite photopolymerization.Acta Biomaterialia 2010;6:20539.

    [122] Hadis MA, Tomlins PH, Shortall AC, Palin WM. Dynamicmonitoring of refractive index change through photoactiveresins. Dental Materials 2010;26:110612.

    [123] Feng L, Suh BI, Shortall AC. Formation of gaps at thefillerresin interface induced by polymerizationcontraction stress: gaps at the interface. Dental Materials2010;26:71929.

    [124] Hadis MA, Shortall AC, Palin WM. Competitive lightabsorbers in photoactive dental resin-based materials.Dental Ma