Composite Manufacture

Embed Size (px)

Citation preview

  • 7/29/2019 Composite Manufacture

    1/10

    INTRODUCTION

    The field of dental restorative resin composites is still

    open for improvement and thus a much active area of

    research1). In fact, despite the recent advancements

    both in chemistry of the resins and in development

    of novel nanoparticles and related technologies2,3),

    many combinations thereof have to be explored and

    the respective recipes have to be optimized yet. The

    main issue with the dental resin composites is that

    still their mechanical properties can only barely

    approach those of dental amalgam4). Unfortunately, the

    industrial manufacturers of dental composites, which

    have the resources in both man-power and accessible

    instrumental techniques, on the one hand are forced to

    focus on the health issues of the novel materials, and

    on the other hand are pushed to deliver new products

    on the market as soon as possible to win the race with

    their competitors. As a result, materials that are poorly

    characterized in physical and mechanical properties

    are sometimes released and promoted beyond their real

    performances, as it has been the case in the last decade for

    several flowable composites5,6). In this scenario, a deeper

    understanding of the combined effects of the variousparameters of composite materials is to be attained, by

    preliminary investigations carried out in the laboratory.

    In this work we investigated simple composite systems

    without filler-matrix bonding agent, and within this

    limitation we studied the effect of different filler particle

    materials on the respective elastic properties. Three

    different types offillers were selected, namely ball-milled

    glass, and commercial particles of titania and silica. The

    ball-milled glass was chosen for the potential control

    of the particle size in our laboratory in view of future

    developments, and the commercial silica particles were

    chosen as the standard counterpart for comparisons.

    Titania has been chosen additionally as a novel filler

    material, thanks to its mechanical stiffness due to

    the crystalline particles nature, and to the interesting

    possible photo-activity. In fact, it is known that titania is

    a good absorber of UV light potentially driving chemical

    reactions (photocatalytic effect7)), which could be used to

    trigger e.g. periodic biofilm removal and/or changing its

    surface wettability to superhydrophobic character8).

    In most existing literature the mechanical properties

    of dental composites are investigated by means of static

    loading methods, such as nanoindentation to measure

    elastic modulus and hardness, or universal tester to

    measure flexural strength and fracture toughness.

    However, indentation only reports about the response

    to compressive stress, whereas flexural stress loading is

    also important in restorations, which exhibit interfaces

    bonded to the native tooth cavity. On the other hand,

    strength and toughness tests are destructive, making

    it impossible to repeat testing of the same specimensover the time. Furthermore, both methods are static and

    can only partially account for the viscoelastic nature

    intrinsic in the polymeric matrix phase of the dental

    restorative composites. Therefore, we decided to use

    dynamic mechanical testing as the main experimental

    technique, to provide a more realistic characterization of

    the materials considered.

    Preparation and characterization of a BisGMA-resin dental restorativecomposites with glass, silica and titania fillers

    Sanjay THORAT 1,2, Niranjan PATRA 1, Roberta RUFFILLI 3, Alberto DIASPRO 1 and Marco SALERNO 1

    1 University of Genova, viale Causa 13, I-16145 Genova, Italy2Istituto Italiano di Tecnologia, Department of Nanophysics, via Morego 30, I-16163 Genova, Italy3 Istituto Italiano di Tecnologia, Department of Nanochemistry, via Morego 30, I-16163 Genova, Italy

    Corresponding author, Marco SALERNO; E-mail: [email protected]

    A photo-polymerizable Bisphenol-A diglycidylether methacrylate resin was characterized by Fourier transform infrared spectroscopy

    after its irradiation under different conditions to identify the best curing. Bonding-agent free composites with particles of ball-milled

    glass, silica and titania at loading of 10 and 50%wt were prepared, and their viscoelastic properties investigated by dynamic

    mechanical analysis, in experimental conditions close to the working environment in the mouth. All composites showed good stability

    at the considered conditions. The stiffest composite was the silica one, which was based on the smallest primary particles. The

    storage moduli close to room temperature (25C) and mastication frequency (1 Hz) were extracted as reference bending moduli for

    the materials, and compared to static compressive moduli measured by nanoindentation performed by atomic force microscopy.

    Nanoindentation showed qualitative results in agreement with dynamic mechanical analysis as to the ranking of different materials,

    while resulting in approximately two-fold elastic modulus.

    Keywords: Dental restorative composites, Inorganic fillers, Dynamic mechanical analysis, Nanoindentation, Elastic modulus

    Color figures can be viewed in the online issue, which is avail-

    able at J-STAGE.

    Received Dec 12, 2011: Accepted Apr 2, 2012

    doi:10.4012/dmj.2011-251 JOI JST.JSTAGE/dmj/2011-251

    Dental Materials Journal 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    2/10

    MATERIALS AND METHODS

    Resin matrix

    2,2-bis[4-3-(methacryloxy-2-hydroxy-propoxy)-phenyl]-

    propane (BisGMA) and triethylene glycol dimethacrylate

    (TEGDMA) were mixed in 7:3 ratio by weight. The

    former is the traditional resin monomer used in dental

    restorative composites since more than half a century9).

    The latter is one among several monomers usually

    mixed with BisGMA to make the paste more fluid

    and handy during manipulation (viscosity decreasing

    co-monomer)1). The system was carefully blended by

    spatulation for 3 min. The co-monomer mixture was

    further added with the photopolymerization system,

    consisting in camphorquinone (CQ) as a photoinitiator

    and dimethyl amino ethyl methacrylate (DMAEMA) as

    an amine reducing agent. CQ and DMAEMA were added

    in 1:1 weight ratio, such that their total amount was 0.5%

    wt of the total co-monomer. All products were supplied

    by Sigma-Aldrich (Milan, Italy). The overall system was

    blended again by spatulation for additional 3 min. Incase offiller loading, the respective particles were added

    in 10 or 50% wt proportion of the overall organic matrix

    paste, and the system was spatulated again for up to

    additional 15 min for the highest loading cases. The

    paste was then poured into a clean antisticking mold

    of Teflon, and placed in a bell rest chamber pumped to

    low vacuum (~100 mbar) to remove air bubbles formed

    during spatulation.

    Filler materials

    Two of the three types offillers used were commercial

    materials, namely silica and titania particles

    (Sigma-Aldrich products no. 718483 and 232033,

    respectively). The silica particles are described as

    nanopowder with average diameter of 12 nm, and

    are amorphous in character. The titania particles are

    described as powder without any size specification,

    and are crystalline in character (anatase). The third

    type offillers used was ball-milled glass. For the source,

    optical microscopy specimen coverslips were used

    (Menzel-Glser, Germany), of size 20200.3 mm3.

    These coverslips are made of borosilicate glass (D 263 M

    type, Schott AG, Germany). Despite the large planar

    dimensions, these slides are so thin that it was possible

    to load them directly into the ball-miller jar without

    preliminary fragmentation. The ball-miller used was

    a PM100 (Retsch, Germany), operated with a 50 mLzirconia jar, filled with 100 g of glass and zirconia balls.

    We used 30 balls of 5 mm diameter together with 60 balls

    of 3 mm diameter. The milling was carried out in 20 mL

    isopropanol (IPA, Sigma-Aldrich) at a rotating speed of

    450 RPM, with alternating cycles of 1 min clockwise and

    1 min counterclockwise rotation without resting time,

    for a total milling time of 5 h. After milling, the hot jar

    (~80C) was let to cool down to room temperature (RT)

    in place, before opening it. The suspension was then

    poured into a previously weighted glass beaker, and

    was let to dry in an oven at 90C for 4 h. After drying,

    the glass appeared to be agglomerated in large plates

    sticking to the jar bottom, but these could be easily

    broken down into fine powder by simply touching them

    with steel tweezers. It was thus possible to mix this

    powder thoroughly with the resin paste, similarly to the

    commercial particles.

    In order to compare the commercial nanosilica and

    titania powders with the glass ball-milled in IPA, the

    particle size of each filler material was first measured by

    dynamic light scattering (DLS) in IPA suspensions. To

    this goal, polystyrene cuvettes were used in a Nano-ZS

    setup (Malvern Instruments, UK). We started from

    concentrated IPA suspensions (~5 g/L for the commercial

    powders, and as collected material from the ball-milled

    glass), and then moved on to a more diluted suspension

    of the same material, decreasing the concentration to

    50% at each step. We went on with dilution until the

    measurement quality resulting from the instrument

    report remained acceptable, as to the sufficient optical

    density required for a good statistical analysis. Also, for

    each suspension 3 series of 3 measurements each were

    repeated and averaged.In addition to DLS measurements in IPA, after

    drying the IPA suspension of ball-milled glass the filler

    particles were measured again in DLS upon mixing them

    into the resin. Since the mixtures had to be manipulated

    in the light for several minutes, for these measurements

    we used BisGMA and TEGDMA only without

    photopolymerizing system (CQ-DMAEMA). The missing

    CQ-DMAEMA part should not affect significantly the

    rheology of the system, due to the low percentage (0.5

    wt% of the co-monomer). Finally, microscopic imaging

    was also used to further assess the size of bare primary

    particles and particle aggregates.

    Photo-curing conditions

    The elongated beam shape of the specimens (see

    Dyamic mechanical analysis for the size) required

    three irradiation cycles for each specimen, which

    were applied starting from the central region first and

    moving to the two side regions later on. For selection of

    the most appropriate photo-curing lamp, the intensity

    of various light sources available in our laboratory was

    preliminary measured with a power meter Nova II

    (Ophir, USA). Since CQ has a peak of light absorbance

    at 470 nm wavelength, which efficiently starts the

    polymerization reaction after amplification by the

    electrons extracted from DMAEMA, we measured both

    the full spectrum power (white light) and the power ata selected blue region window only (blue light). To this

    goal, a filter was taken from a fluorescence cube GFP

    (Semrock, USA), with ~95% transmittance at a 455

    490 nm wavelength pass band. As a result of this step,

    we selected a X-Cite 120 lamp (EXFO, Canada), which

    was placed at a distance of ~3 cm from the specimens to

    be cured. This lamp is a high output power white light

    source with a broad smooth spectrum, normally used

    for optical microscopy imaging of specimens stained

    with fluorescent dyes. The irradiance of this lamp was

    evaluated in comparison with other light sources used in

    both literature and recent clinical practice10-12), on which

    636 Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    3/10

    basis we decided to set our irradiation time to 105 and

    210 min for white and blue light, respectively, which

    methods were both tested for curing the bare resin.

    Fourier transformed infrared spectroscopy

    In order to verify the conversion of the co-monomer paste

    into a polymerized resin, Fourier transformed infrared

    spectroscopy (FTIR) was carried out for all the specimens

    at RT in ambient air, both soon after pouring the paste

    into the Teflon molds and after its irradiation. The FTIR

    spectra were acquired by a Vertex 70 spectrometer

    (Bruker, USA), in the range of 4004,000 cm1. The

    samples were analyzed in attenuated total reflection

    configuration, with an aperture diameter of 3 mm and a

    spectral resolution of 4 cm1. For optimal signal-to-noise

    ratio, 50 scans were averaged per sample spectrum,

    and apodized by applying the Blackman-Harris 3-term

    correction function for the Fourier transformation.

    The interferograms were corrected using a zero-filling

    factor of 2. All the spectra were baseline-corrected by

    third order polynomial and were normalized thereafter.

    Dynamic mechanical analysis

    All the materials were shaped with the Teflon mold to

    be rectangular beams of 13352 mm3 in size. On these

    samples we performed dynamic mechanical analysis

    (DMA) by means of a Q800 setup (TA Instruments,

    USA), with instrument compliance of less than

    0.2 m/N, as determined by a prior calibration in staticloading mode. We carried out DMA measurements

    in single-cantilever mode, under strain control in the

    materials linear regime. The maximum applied strain

    was 35 m. Temperature sweeps at strain frequency of1 Hz were carried out, in a range of +2 to +62C (with

    5C steps, 5C/min rate), since it should represent well

    the limit values occurring in human mouth in normal

    operating conditions, when ingesting from hot food to

    icy drinks. For reaching the lowest temperatures in this

    range, liquid nitrogen was used as a coolant. During

    the temperature scans both the storage modulus E and

    the loss modulus E were recorded, representing the

    in-phase (real) and out-of-phase (imaginary) parts of a

    complex modulus E*=E+iE, respectively, occurring due

    to the stress lagging behind the applied strain with some

    phase angle delay.

    AFM nanoindentation

    The same samples as prepared for the DMA were used,before carrying out the respective tests. We performed

    nanoindentation by means of an AFM used in so-called

    force spectroscopy mode, i.e. collecting force-distance

    curves on given specimen sites. We used a MFP-3D AFM

    (Asylum Research, USA) equipped with gold coated

    silicon probes NSG20 (NT-MDT, Russia), made each of a

    cantilever with nominal spring constant and resonance

    frequency of ~60 N/m and ~450 kHz, respectively, and of

    a terminal transverse pyramidal tip with apex diameter

    and full aperture angle of ~20 nm and ~22, respectively.

    The typical lateral scan size was 5 m, with spatial mapsampling of 302 pixels. The AFM probe was calibrated in

    air for determination of the cantilever spring constant,

    and again in de-ionized water for determination of

    the appropriate optical lever sensitivity. In fact, the

    force-distance curves were acquired in water to remove

    the effect of aspecific tip-surface adhesion due to

    ambient moisture. The curves had 1,024 datapoints, with

    z actuation loops of 1 m range and 0.5 Hz frequency,low enough to minimize viscous drag effects, (data not

    shown).

    Preliminary to the AFM nanoindentation

    experiments, occasionally AFM imaging was carried

    out with the same probe (in Tapping mode, with 30 mscan size and 2562 pixels), which showed evidence of the

    presence offillers in the composites, as compared to the

    bare resin surface (data not shown). However, all the

    composite samples were still flat and smooth enough at

    the surface, such that they could be properly investigated

    by nanoindentation at the considered low indentations

    (up to limited maximum values ofmax~100 nm only).

    After acquisition, the indentation was calculated from

    the z movement of the actuator z and from the changein cantilever deflection D (partially compensating theformer) as =zD. The force-indentation data werefinally fit to the Hertz model of elastic contact, using the

    unload part of the force loop to find the elastic modulus

    values EAFM. Given the high number of parameters in the

    Hertz model, (indenting tip size and shape, sample and

    tip materials Poissons ratios, tip elastic modulus, actual

    contact point (, F)=(0, 0)), both offsets of indentation and

    force were let to fit automatically by the AFM software

    within a broad range of forces, 25%75% of Fmax, such

    as to minimize the deviations between fitting curve

    and data-points (reduced 21,000). Additionally, theremaining parameters were adjusted on the bare resin

    reference material, such that for this the compression

    modulus EAFM resulting from nanoindentation was equal

    to the respective flexural modulus E resulting from the

    DMA experiments. After this initial setup, the same

    values of the remaining working parameters were used

    later also for the composites.

    RESULTS

    Resin curing evaluation

    In Fig. 1a) the FTIR spectra of the uncured components

    of the organic paste are presented, both individually

    and blended in the resulting mixture (red line). Clearly,

    BisGMA and TEGDMA dominate the spectrum of themixture, with some features, such as e.g. the peaks at

    2,874 cm1 (symmetric stretching of CH3) and 1,126 cm1

    (symmetric vibration of -C-O-C-) of TEGDMA increasing

    the adjacent shoulders of BisGMA. The peaks of

    DMAEMA at 2,771 cm1 (N(CH3)2 band) and of CQ at

    1,747 cm1 (C=O vibration) do not appear in the mixture,

    due to the low relative contribution of the respective

    components. The 3,3503,550 cm1 band (OH stretching)

    is slightly depressed.

    In Fig. 1b), the effect of photo-curing is reported.

    In particular, a significant change in the bands

    centered around 1,637 and 1,580 cm1 is observed, with

    637Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    4/10

    Fig. 1 a) FTIR spectra of individual BisGMA, TEGDMA, CQ and DMAEMA, as well as of the (uncured) mixture

    of the same compounds (red line). b) FTIR spectra of the resin mixture before (uncured, red) and after

    irradiation, according to different conditions.

    a relative increase of the second one, corresponding

    to the stretching of the aliphatic and aromatic (i.e.

    ring) C=C bonds, respectively. In fact, as a measure of

    photo-polymerization, the degree of conversion (DC) of

    the monomer mixture into the polymer (photo-cured

    mixture) can be evaluated from the following equation,

    according to a peak intensity method14):

    (I1637 / I1580) polDC= 100 [ 1 - ] (1)

    (I1637 / I1580) mon

    where I1637 and I1580 are the peak intensities of thebands at the respective wavenumber positions, and the

    subscripts outside the parentheses refer to the spectra

    before (mon) and after (pol) photo-curing, respectively.

    From eq. 1 applied to the spectra in Fig. 1b), it turns

    out that the highest conversion is found for white light

    curing (see Fig. 4). This condition (gray curve in Fig. 1b))

    provided a DC of ~75% soon after cure, which also

    showed the highest time-delayed increased one week

    later (black curve in Fig. 1b), reaching ~94% (see Fig. 4).

    Thus, spectral measurement repeated on white light

    exposed resin after one week of storage in ambient light

    at RT showed still significant ongoing conversion on that

    time scale, (+25% in DC). However, when repeated again

    after one month since irradiation, no further change was

    observed, showing no effect of possible absorbance of

    moisture or other ambient contaminants.

    The blue exposed resin, on the contrary (blue curve

    in Fig. 1b)), only reached ~57% soon after curing (Fig. 4),

    and also showed lesser improvement at one week time

    (~63%, i.e. +11%, curve not shown).

    Filler size characterization

    The apparent size of the filler particles was determined

    by means of DLS, which was first applied on suspensionsof the respective powder in IPA. The adopted method of

    progressively more diluted suspension should minimize

    the effect of agglomeration, allowing for a determination

    of particle size as close as possible to the primary particles.

    In fact, after 3 to 5 dilution steps, for all materials the

    apparent average particle size had decreased, reaching

    either a more stable value or the limit of lowest

    acceptable measurement quality mentioned above. The

    final distributions of particles size in IPA are shown in

    red in Fig. 2a)c). The apparent mean particle size in

    IPA is ~650 nm for glass (Fig. 2a) red bars) and ~950 nm

    for titania (Fig. 2b) red bars), respectively. For silica

    638 Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    5/10

    Fig. 2 a)c): distributions of particle size (populations by particle number) obtained in IPA (red bars) and in

    resin co-monomer (green bars), for the particles of: a) glass, b) titania, and c) silica. d)e): images offiller

    particles drop-cast onto glass slides from IPA: d) AFM of glass particles, and e) SEM of titania particles.

    (Fig. 2c) red bars) two populations showed up, centered

    around ~180 and ~610 nm size, respectively.

    In order to characterize a system closer to the final

    samples, we also carried out DLS of the filler particles

    dispersed in the resin. Since the 50 wt% filler loaded

    mixtures were too viscous to be properly poured into

    the DLS cuvettes, we restricted our measurements to

    the 10 wt% samples. These results are also included in

    Fig.2a)c), with size population distributions in green

    color. The glass particles distribution in the resin (Fig.2a)

    green bars) showed three peaks, with the intermediate

    one (~700 nm) well overlapped with the distribution inIPA (red bars), and two more peaks on both the small

    and large size sides, centered at ~240 and ~2000 nm,

    respectively. The titania particles distribution in

    the resin (Fig.2b) green bars) still showed a single

    population peak same as in IPA, yet significantly shifted

    to lower size, centered around ~200 nm. Finally, the

    silica particles distribution (Fig.2c) green bars) is again

    partly overlapped with the IPA one, but with the middle

    shifted towards smaller size. In particular, two peaks at

    ~9 and ~24 nm appear, showing overall good consistency

    with the nominal mean particle size of 12 nm.

    In fact, for those fillers without any nominal

    reference value for the primary particles size, i.e. titania

    and glass, additional determination of particle size was

    also carried out by microscopy. To this goal, the same

    IPA solutions used for DLS were drop-cast onto glass

    slides, and the surfaces were imaged. We first tried

    AFM, which worked for the relatively larger and more

    irregular glass particles (see Fig.2d)), whereas for the

    titania particles it turned out into unstable images,

    probably due to particle aggregates loosely bound to

    the substrates and swept during the scan. Therefore

    for titania we used SEM, after 5 nm platinum coating,

    (see Fig.2e)). The images showed a general agreementwith DLS results in the resin. Indeed, for the glass the

    presence of mostly submicrometer sized particles is also

    accompanied by some larger ones (up to 23 m in atleast one direction). Similarly, the image of the titania

    particles shows a primary particle size of 21060 nm

    (meanstandard deviation), in agreement with the green

    distribution in Fig.2b).

    Samples mechanical properties

    We first carried out DMA on the bare resin samples,

    to assess the mechanical properties resulting from

    the different curing conditions. The respective

    639Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    6/10

    Fig. 3 a) Storage modulus E and b) damping factor tan=E/E of bare resin samples cured in different conditions.

    Fig. 4 Comparison between DC from the FTIR spectra

    and storage modulus E at RT from the 1 Hz DMA

    frequency scan in Fig. 3 (blue light curing curve),

    stressing the possible correlation between the two

    quantities.

    measurements are presented in Fig. 3. In Fig. 3a)

    the storage modulus E decreases monotonously in all

    cases, as expected due to the increased fluidity of the

    resin at higher temperatures. The white cured storage

    modulus curve shows not only the highest values at all

    temperatures, but also the slowest decrease slope than

    the other curves. For an easier comparison with DC, the

    E values at RT have been extracted from Fig. 3a) and

    plotted again along with the respective DC values in

    Fig. 4.

    In Fig. 3b) we decided to plot the damping factor

    tan=E/E rather than the loss modulus E or the total

    complex modulus |E*|=(E 2+E2), as done sometimesin the literature15). In fact, only one out of the three

    mentioned parameters is independent in addition to

    E, such that each choice is allowed. However, whereas

    |E*| is usually quite close to E (being E

  • 7/29/2019 Composite Manufacture

    7/10

    Fig. 5 a) Storage modulus E and b) damping factor tan=E/E of resin composites with differentfiller materials

    and loading, when available.

    Fig. 6 Flexural (red bars) and compression modulus (green

    bars) of cured specimens of all considered materials,

    from left to right: bare resin, and composites with

    glass, titania and silica fillers. The hardness from

    the nanoindentation experiments (blue bars) is

    also plotted.

    the lowest decreasing slope with increasing temperature.

    In fact, this curves starts at the lowest temperatures withE values only lower than for the glass composites, and

    ends at the highest temperatures with higher E than

    those samples. The cross-point is approximately placed

    between the RT and the body temperature values.

    For the damping factor plotted in Fig. 5b), this control

    information shows again, similarly to the bare resin case

    (Fig. 3b)), that no glass transition occurs in the considered

    temperature range. Also in this parameter, describing

    the amount of viscous character of the material, does the

    silica composite perform better than the others, as its

    curve is the most flat, showing high thermal stability

    in the operating temperature range. Again, the glass

    composites with different loading are very close to each

    other (even more than for E in Fig. 5a)).

    Finally, the results of the nanoindentation analysis

    are summarized in Fig. 6, where the values of EAFM

    modulus are compared with the E modulus from

    DMA, for all the composites. The error bars of the

    AFM quantities represent one standard deviations

    of the populations of 302 data-points from the maps of

    force-distance curves on the respective samples. These

    error bars are approximately ~24% and ~33% of the

    respective mean values, for the 50%wt and the 10%wt

    composites, respectively. In fact, a larger deviation is

    expected for the low (10%wt) loading composites, where

    regions of dominating bare resin stiffness effect can be

    found on the surface more likely than in the average

    high loading (50%wt) composites. In Fig. 6 bars for the

    measured hardness H have also been plot, which is an

    additional information obtained from the Hertz fits. In

    general, the relative error on HAFM is higher than that on

    EAFM, due to the higher sensitivity of H to the uncertainty

    in contact area with respect to E. Additionally, an even

    higher error is found for both HAFM and EAFM for the low

    loading composites, where the relative error with respect

    to the mean is increased from ~32% to ~42%.Overall, the elastic modulus reported in Fig. 6

    is in good qualitative agreement with DMA results.

    Indeed, in both techniques the glass composites show

    values significantly higher than the bare resin, and

    not significantly different between them, for the two

    different loadings considered. Also similarly to DMA,

    the lowest moduli from nanoindentation come from the

    titania composites. However, whereas in DMA at RT the

    10%wt silica composite scores the same as the two glass

    composites, in nanoindentation it is significantly higher,

    and clearly the stiffest composite of all. Additionally,

    whereas for the bad cured 50%wt titania composite DMA

    641Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    8/10

    showed a modulus even lower than the bare resin, for

    compressive nanoindentation test the presence offillers

    makes the resulting modulus higher than the resin

    even in that negative case. In fact, it is known that for

    tensile measurement the presence of non-bonded fillers

    can even act as detrimental defect sites, rather than as

    reinforcing agents16).

    DISCUSSION

    Current commercial dentist lamps are based on LEDs

    which present a typical emission light power of ~1,000 mW

    in the spectral region of 460480 nm wavelength. In

    common practice two 10 s irradiation cycles of these

    lamps are used, for maximum 2 mm thick restorations.

    The lamps are usually placed close to contact (~2 mm)

    with the dental composite restoration, which has typical

    diameter of 3 mm and is completely covered by a light

    spot of approximately the same size, at the specified

    distance. Therefore, a used irradiance of ~14,000 mW/

    cm2

    with a total delivered dose surface density of ~280 J/cm2 can be estimated.

    The lamp used by us had a power measured in the

    blue region of ~60 mW at 3 cm distance from the sensor

    on a circular area of 13 mm diameter. This lamp was

    the most powerful continuous wave source of blue light

    available in our laboratory. Alternatively, we could use

    pulsed lasers with high peak power, borrowed from

    a spectroscopy laboratory close by, but we suspected

    that the pulsed irradiation regime could modify the

    photo-curing process dynamics or chemistry, after

    heating effects. Therefore, we stuck to our continuous

    light source and rather extended the irradiation times.

    The selected times of 105 and 210 min for white and

    blue light, respectively, were chosen to reach the same

    dose surface density of the dentists lamp. We assumed

    100% absorption of the delivered dose, and a simple

    exposure reciprocity law between incident irradiance

    at specimen pinc and exposure time t, pinca t=constant,

    with a=1. Whereas this assumption can have limited

    validity in some cases also depending on parameters

    such as photoinitiator system, filler loading, and desired

    depth of cure, it can still be considered valid in a first

    approximation. In particular on the long time side of

    this law, it has been shown that there is no minimum

    irradiance under which no photopolymerization

    starts13).

    When discussing the FTIR spectra in Fig. 1 and therespective DC calculated and plotted in Fig. 4, it should

    be kept in mind that it is not totally clear to what extent

    thermal effects played a role in the curing, by eventually

    modifying the reaction path and the resulting material

    properties. In fact, when exposed to the white light

    the resin was locally heated above 80C. Similarly,

    thermal effects also applied in the FTIR measurement

    repeated after DMA study (magenta curve in Fig. 1b)).

    In this case, the heating cycle decreased the DC to ~26%

    (Fig. 4). One possible reason for this effect could be the

    partial vaporization of water molecules trapped inside

    the resin, which hinders chain propagation during the

    polymerization.

    Independent on the highest DC obtained with the

    white light exposure, for the curing of the composites

    we decided to adopt the blue light exposure, such as to

    follow more closely the conventional photo-curing adopted

    in normal dental practice. In fact, one more problem

    appears in the composites due to the fillers scattering

    the light off its original straight path through the

    organic phase. Actually, this can have two counteracting

    effects. On one hand, some light exits the specimen and

    is lost. On the other hand, most light incoming on the

    fillers is not absorbed by them but rather redirected to

    the locally surrounding matrix, which is also reduced in

    quantity. Therefore, low loading composites should be

    cured even faster than bare resin specimens. However,

    in high loading composites the light could even not reach

    at all the deepest matrix levels. Overall, we assumed a

    general compensation of these effects on the considered

    specimens (thickness ~2 mm), and applied to the

    composites the same blue irradiation procedure used

    for the bare resin specimens.Regarding the filler particle size from Fig. 2 it

    clearly appears that all the average particles in IPA

    are at maximum of micrometer size. In particular,

    the ball-milled glass particles stay below that limit,

    appearing even smaller than the titania particles. On

    the other hand, for the only particles of known size, i.e.

    the silica nanoparticles, the apparent size in IPA is much

    larger than the nominal value. Clearly, some degree of

    aggregation in suspension is present in IPA at least for

    silica.

    When repeating the DLS measurements of the

    filler particles in the resin medium, obviously, due to

    the higher viscosity as compared to solvent, the glass

    particles were less free to move, and a lower degree of

    both aggregation and sedimentation appears, resulting

    in the presence of some primary and some very large

    particles, respectively. For the titania particles, in turn,

    clearly aggregation was less effective in the resin, and

    the resulting size can be supposed to be the real primary

    particle size.

    Concerning the direct microscopic imaging of the

    fillers, in Fig. 2d)e) we have shown images of bare

    stand-alone particles only, and not of particles embedded

    inside the cured composites. However, also given the

    probably slow reaction rate following the low used

    irradiance, there is no particular reason why the size of

    either the primary particles or the particle aggregatesshould undergo major changes during photocuring with

    respect to the size found in the DLS measurements in

    the resin.

    In Fig. 3a), the white cured sample was the one

    with the best polymerized resin, which is in agreement

    with the highest DC value observed (Fig. 4). In fact, the

    two curves of the blue cured resin in Fig. 3a) present an

    inversion of ranking when compared with the respective

    DC (Fig. 4). Indeed, the decrease in DC resulting from

    the DMA thermal scan did not correspond to a decreased

    modulus but rather increased it, instead. Obviously,

    polymer cross-linking is not the only mechanism

    642 Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    9/10

    accounting for the increase in the mechanical properties

    of the polymer, and thus only specimens processed in

    exactly the same way should be compared in expected

    mechanical properties with respect to their DC.

    In Fig. 3b), the observed minor transition (curve

    peak) for the blue cured resin (blue line) can partly

    account for the inversion of E ranking with respect

    to DC observed in Fig. 4. However, on repeating the

    temperature scan, this weak maximum disappears

    (magenta curve). Obviously, the respective transition, if

    any, was related to orientation of some side groups of the

    polymer or localized backbone motions (beta transition).

    Finally, we conclude that no glass transition occurs for

    the resin in the considered temperature range.

    When discussing the results in Fig. 5, reporting

    DMA measurements of the composites, it should be

    mentioned that the respective DC values were similar

    for all samples but for the 50%wt titania sample, which

    showed much lower DC, as low as ~10%. In fact, the

    50%wt titania sample also showed direct evidence of

    incomplete curing as its surface was still gelly and stickyafter curing. On the contrary, the other composites were

    properly hardened and had DC values between 60 and

    75%, all higher than bare resin (~57% with blue light),

    as expected due to the dominating effect of less resin

    quantity to be cured in the composite for the same dose

    delivered.

    In Fig. 5a), the more limited decrease in storage

    modulus with increasing temperature for the silica

    samples than for all the other samples may be due to the

    smallest particle size, which provided better dispersion

    and higher uniformity of the composite, resulting in less

    regions of resin only domains. In this way, the decrease

    in modulus, a peculiarity of the resin matrix, is less

    pronounced, as the filler is maintaining the modulus

    at relatively high levels during all the thermal cycle.

    On the contrary, the glass fillers, similar to the titania

    fillers (in the 10%wt sample) exhibit higher plasticizing

    effect on the resin, probably due to the larger primary

    particle size.

    In Fig. 5b), the 10%wt titania sample, despite its

    high DC (~75% vs 60% and 70% of the glass samples) also

    shows high viscosity, probably due to some higher flow of

    the titania particles in the resin with respect to the glass

    ones, perhaps due to more difficult diffusion of the latter

    after the higher size. The final maximum appearing for

    the 50%wt titania sample is probably associated with

    secondary effects such as the incomplete curing in caseof this sample, as mentioned above. The apparent glass

    transition is probably due only to a higher fluidity of the

    not polymerized material.

    From the nanoindentation results, overall the

    compressive modulus EAFM is approximately twice as

    much as the flexural modulus E, with a highest value

    of ~5.3 GPa for the nanosilica composite. Whereas this

    value is still far from the minimum requirement for

    dental restorative composite of posterior teeth, which can

    set to ~10 GPa, it can already be considered a good result

    for the relatively low loading achieved (10%wt, which is

    less by volume, due to the filler being heavier than the

    resin), in the absence of a bonding agent. In fact, the

    latter effect can be the reason of a higher modulus found

    in compressive mode of nanoindentation with respect to

    flexural mode of DMA, where one side of the specimen

    is under compression as well whereas the opposite side

    is on the contrary under tension. Actually, it is known

    that dynamic test mode should result in a higher value

    of the measured modulus, due to the higher strain rate

    applied17,18). Indeed, in our nanoindentation, the rate

    of applied load is such that the maximum load Fmax is

    reached in half the time period of the force-distance

    loop. Being the frequency 0.5 Hz and the overall z sweep

    1 m, this corresponds to 2 /s. In DMA, on the otherhand, the strain rate is such that a cycle with amplitude

    of 35 m is performed with a frequency of 1 Hz, whichmeans 140 m/s, more than one order of magnitudefaster than during indentation. However, in the present

    experiment, obviously the effect due to the different type

    of loading scheme dominates, making the AFM moduli

    always higher than the DMA ones.

    Usually, fillers of glass or other milled materialhave relatively large size population, spanning at least

    one order of magnitude (110 m and often 0.110 m),which makes the respective dental restorative materials

    belong to the class of hybrids. These are normally

    considered to allow for a better distribution of stress,

    and increase strength and toughness and eventually

    elastic modulus1). However, in our case the ball-milling

    provided rather narrow particle size distribution (31%

    of the mean, as from Fig. 2a)). This was also indirectly

    confirmed by the AFM nanoindentation measurements,

    which found monomodal distributions of material

    stiffness (i.e. uniform composites) even at the low loading

    of 10%wt.

    Between the composites with silica and titania

    fillers, one could expect a higher modulus for the titania

    ones, due to the crystalline phase of the particles.

    However, obviously, at the 10% loading the mostly

    probed character of the composite is still assigned to the

    resin matrix more than the filler particles. On the other

    hand, at 50% the titania sample could not be properly

    cured, due to the high reflectance of the titania, which is

    well known for its use as a white pigment in paints. In

    fact, previous investigations of titania filled composites

    used thermal curing instead of photo-curing19). Indeed,

    the crystalline titania composite with 50%wt loading as

    measured by DMA showed elastic modulus even lower

    than the bare resin.Finally, when comparing composites of differentfiller

    materials as in the present case, it should be considered

    that additional effects also arise from the different

    effective values of loading by volume. In fact, the densities

    of the various filler materials are rather different: 2.6 g/

    cm3 for silica and 3.9 g/cm3 for titania, according to the

    manufacturers data sheets, and 1.1 g/cm3 and 2.4 g/cm3

    for bare resin and glass, as determined by us by weighting

    and differential volumetric measurements of water in

    graduated cylinders. Therefore, 10%wt corresponds to

    volume loadings of 4.7, 4.4 and 2.7% for glass, silica

    and titania, respectively, whereas 50%wt corresponds

    643Dent Mater J 2012; 31(4): 635644

  • 7/29/2019 Composite Manufacture

    10/10

    to 31, 29 and 20% for the same sequence of materials.

    This calculation is made according to nominal densities

    of bulk materials. However, the volume loading of our

    nanosilica was obviously much higher, due to its fluffy

    form, probably around 20% already for the 10%wt sample

    and above 90% for the 50%wt sample, which could not

    be made at all. One way to overcome this problem could

    be using surfactant coated nanosilica, or suspend it in

    a solvent also good for the resin phase, before mixing

    everything together. However, this method will probably

    raise other questions, regarding the solvent effects20).

    CONCLUSION

    After identifying the most appropriate conditions for

    light curing our resin matrix, we carried out DMA

    characterization of bonding-agent free composites

    with different filler materials under flexural stress,

    and compared it to static nanoindentation carried out

    by AFM. The elastic modulus of the composites was

    always improved or at least remained the same as thebare resin in all compression tests. In flexural tests the

    elastic modulus was always lower than the compression

    test values, and even decreased with respect to bare

    resin in the 50%wt titania sample. Additionally, in glass

    composite no increase in elastic modulus was observed

    on increase of loading from 10 to 50%wt. In compression,

    which is in many cases the most important loading mode

    of dental restoration materials, the highest modulus was

    observed for the nanosilicafiller, even if in 10%wt loading

    only, corresponding in turn to ~4.4%vol. In this case,

    an elastic modulus as high as 5.31.6 GPa was found.

    Additionally, no thermal transitions were observed at

    the investigated temperatures for any material, showing

    good stability at operating conditions. Titania fillers

    could be of interest in future applications where their

    photocatalytic effect could be useful to promote local

    antibacterial or remineralization reactions.

    REFERENCES

    1) Chen M-H. Update on Dental Nanocomposites. J Dent Res

    2010; 89; 549-560.

    2) Uskokovi V, Bertassoni LE. Nanotechnology in dental

    sciences: moving towards a finer way of doing dentistry.

    Materials 2010; 3: 1674-1691.

    3) Sharma S, Cross SE, Hsueh C, Wali RP, Stieg AZ, Gimzewski

    JK. Nanocharacterization in Dentistry. Int J Mol Sci 2010;

    11: 2523-2545.4) Wahl MJ. AmalgamResurrection and redemption. Part 1:

    the clinical and legal mythology of anti-amalgam.

    Quintessence Int 2001; 32: 525-535.

    5) Han L, Ishizaki H, Fukushima M, Okiji T. Morphological

    analysis offlowable resins after long-term storage or surface

    polishing with a mini-brush. Dent Mater J 2009; 28: 277-

    284.

    6) Salerno M, Derchi G, Thorat S, Ceseracciu L, Ruffilli R,

    Barone A. Surface morphology and mechanical properties

    of new-generation flowable resin composites for dental

    restoration. Dent Mater 2011; 27: 1221-1228.

    7) Chen X, and Mao SS. Titanium dioxide nanomaterials:synthesis, properties, modifications, and applications. Chem

    Rev 2007; 107: 2891-2959.

    8) Pignatelli F, Carzino R, Salerno M, Scotto M, Canale C,

    Distaso M, Rizzi F, Caputo G, Cozzoli PD, Cingolani R,

    Athanassiou A. Directional enhancement of refractive

    index and tunable wettability of polymeric coatings due to

    preferential dispersion of colloidal TiO2 nanorods towards

    their surface. Thin Solid Films 2010; 518: 4425-4431.

    9) Bowen RL. Use of epoxy resins in restorative materials. J

    Dent Res 1956; 35: 360-369.

    10) Kurachi C, Tuboy AM, Magalhaes DV, Bagnato VS.

    Hardness evaluation of a dental composite polymerized with

    experimental LED-based devices. Dental Mater 2001; 17:

    309-315.

    11) Silva Soares LE, Martin AA, Barbosa Pinheiro AL, PachecoMTT. Vickers hardness and Raman spectroscopy evaluation

    of a dental composite cured by an argon laser and a halogen

    lamp. J Biomed Opt 2004; 9: 601-608.

    12) Hosseinalipour M, Javadpour J, Rezaie H, Dadras T, Hayati

    AN. Investigation of mechanical properties of experimental

    Bis-GMA/TEGDMA dental composite resins containing

    various mass fractions of silica nanoparticles. J Prosthod

    2010; 19:112-117.

    13) Musanje L, Darvell BW. Polymerization of resin composite

    restorative materials: exposure reciprocity. Dent Mater 2003;

    19: 531-541.

    14) Rueggeberg FA, Hashinger DT, Fairhurst CW. Calibration

    of FTIR conversion analysis of contemporary dental resin

    composites. Dent Mater 1990; 6: 241-249.

    15) Ryou H, Niu L-N, Dai L, Pucci CR, Arola AA, Pashley DH,

    Tay FR. Effect of biomimetic remineralization on the dynamicnanomechanical properties of dentin hybrid layers. J Dent

    Res 2011; 90: 1122-1128.

    16) Darvell BW. In: Darvell BW, editor. Materials science for

    dentistry. 9th ed. Sawston, Cambridge: Woodhead Publishing

    Limited; 2011.

    17) Braem M, Davidson CL, Vanherle G, Van Doren V, Lambrechts

    P. The relationship between test methodology and elastic

    behavior of composites. J Dent Res 1986; 66: 1036-1039.

    18) Jacobsen PH, Darr AH. Static and dynamic moduli of

    composite restorative materials. J Oral Rehabil 1997; 24:

    265-273.

    19) Yoshida K, Taira Y, Atsuta M. Properties of opaque resin

    composite containing coated and silanized Titanium dioxide.

    J Dent Res 2001; 80: 864-868.

    20) Patra N, Salerno M, Diaspro A, Athanassiou A. Effect of

    solvents on the dynamic viscoelastic behavior of poly(methyl

    methacrylate) film prepared by solvent casting. J Mater Sci

    2011; 46: 5044-5049.

    644 Dent Mater J 2012; 31(4): 635644