28.01.2015 Views

Long-term release of monomers from modern dental ... - Dentsply

Long-term release of monomers from modern dental ... - Dentsply

Long-term release of monomers from modern dental ... - Dentsply

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Eur J Oral Sci 2009; 117: 68–75<br />

Printed in Singapore. All rights reserved<br />

<strong>Long</strong>-<strong>term</strong> <strong>release</strong> <strong>of</strong> <strong>monomers</strong> <strong>from</strong><br />

<strong>modern</strong> <strong>dental</strong>-composite materials<br />

Polydorou O, Ko¨nig A, Hellwig E, Ku¨mmerer K. <strong>Long</strong>-<strong>term</strong> <strong>release</strong> <strong>of</strong> <strong>monomers</strong> <strong>from</strong><br />

<strong>modern</strong> <strong>dental</strong>-composite materials. Eur J Oral Sci 2009; 117: 68–75. Ó 2009 The<br />

Authors. Journal compilation Ó 2009 Eur J Oral Sci<br />

The elution <strong>of</strong> <strong>monomers</strong> <strong>from</strong> composite materials influences the biocompatibility <strong>of</strong><br />

<strong>dental</strong> restorations. The purpose <strong>of</strong> the present study was to investigate the elution <strong>of</strong><br />

<strong>monomers</strong> [bisphenol A glycidyl methacrylate (BisGMA), triethylene glycol dimethacrylate<br />

(TEGDMA), urethane dimethacrylate (UDMA), and bisphenol A (BPA)]<br />

<strong>from</strong> two light-cured materials (nanohybrid and ormocer) and <strong>from</strong> a chemically cured<br />

composite material, after different curing times (0, 20, 40 and 80 s) and different<br />

storage periods (24 h, 7 d, 28 d, and 1 yr after curing). Each specimen was stored in<br />

1 ml <strong>of</strong> 75% ethanol. This medium was renewed after 24 h, 7 d, 28 d, and 1 yr. The<br />

ethanol samples were analyzed using liquid chromatography tandem mass spectrometry<br />

(LC–MS/MS). The amount <strong>of</strong> <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> the nanohybrid and the<br />

chemically cured composite was significantly higher than <strong>release</strong>d <strong>from</strong> the ormocer.<br />

The curing time exerted a significant effect on the <strong>release</strong> <strong>of</strong> <strong>monomers</strong>. For the<br />

nanohybrid, less monomer was <strong>release</strong>d after increasing the curing time. For the<br />

ormocer, 80 s <strong>of</strong> curing resulted in a higher degree <strong>of</strong> monomer <strong>release</strong>. The effect <strong>of</strong><br />

storage differed between the <strong>monomers</strong>. Although the elution <strong>of</strong> TEGDMA was significantly<br />

decreased after storage for 28 d and 1 yr, a similar amount <strong>of</strong> BisGMA was<br />

<strong>release</strong>d at each storage time-point analyzed, even after 1 yr. The present study<br />

showed that ormocer <strong>release</strong>d a very small amount <strong>of</strong> <strong>monomers</strong> compared with the<br />

other materials.<br />

Ó 2009 The Authors.<br />

Journal compilation Ó 2009 Eur J Oral Sci<br />

European Journal <strong>of</strong><br />

Oral Sciences<br />

Olga Polydorou 1 , Armin Kçnig 2 ,<br />

Elmar Hellwig 1 , Klaus Kümmerer 2<br />

1 Department <strong>of</strong> Operative Dentistry and<br />

Periodontology, Dental School and Hospital,<br />

University Medical Center Freiburg, Freiburg;<br />

2 Department <strong>of</strong> Environmental Health<br />

Sciences, University Medical Center Freiburg,<br />

Freiburg, Germany<br />

Olga Polydorou, Department <strong>of</strong> Operative<br />

Dentistry and Periodontology, Dental School<br />

and Hospital, University Medical Center<br />

Freiburg, Hugstetter Straße 55, 79106 Freiburg<br />

i. Br., Germany<br />

Telefax: +49–761–2704762<br />

E-mail: olga.polydorou@uniklinik-freiburg.de<br />

Key words: composite; <strong>dental</strong> restorative<br />

material; monomer; photopolymerization<br />

Accepted for publication November 2008<br />

The increased demand for esthetic <strong>dental</strong> restorations<br />

during the last few years in the general <strong>dental</strong> practice<br />

has led to an increased use <strong>of</strong> composite materials. It has<br />

been assumed that composite materials have excellent<br />

mechanical properties, good esthetic quality, are easy to<br />

handle, and are able to bond with the enamel surface.<br />

However, a major concern exists about their toxicity and<br />

consequently their effect on human health. Different<br />

substances (such as residual <strong>monomers</strong>, additives, and<br />

degradation products) may irritate the oral s<strong>of</strong>t tissues,<br />

stimulate the growth <strong>of</strong> bacteria, and promote allergic<br />

reactions (1).<br />

The cross-linking dimethacrylates bisphenol A glycidyl<br />

methacrylate (BisGMA), urethane dimethacrylate<br />

(UDMA), triethylene glycol dimethacrylate (TEG-<br />

DMA), and bisphenol A ethoxylated dimethacrylate<br />

(BisEMA) are most widely used for the preparation <strong>of</strong><br />

composite materials. Moreover, resin composites contain<br />

various additives, such asphotoinitiators (e.g.<br />

camphoroquinone), co-initiators (e.g. dimethyl-aminobenzoic-acid-ester),<br />

inhibitors <strong>of</strong> polymerization (e.g.<br />

BHT [butylated hydroxytoluene]), and photostabilizers<br />

(e.g. benzophenone). The differences in composition <strong>of</strong><br />

the <strong>monomers</strong> used for each composite material impact<br />

the reactivity, the viscosity and the polymerization<br />

shrinkage <strong>of</strong> the monomer as well as the mechanical<br />

properties <strong>of</strong> the composite material (2). Therefore, the<br />

elution <strong>of</strong> components (residual <strong>monomers</strong>, oligomers,<br />

and degradation products) <strong>from</strong> resin composites may<br />

influence the biocompatibility <strong>of</strong> the restorations (3).<br />

Several factors contribute to the process <strong>of</strong> elution<br />

<strong>from</strong> resin-based <strong>dental</strong> materials. According to Ferracane<br />

(4), the first relates to the amount <strong>of</strong> components<br />

<strong>release</strong>d, and this should be directly related to the extent<br />

<strong>of</strong> the polymerization reaction (i.e. the degree <strong>of</strong> doublebond<br />

conversion). Second, the chemistry <strong>of</strong> the solvent<br />

has a significant effect on the elution. Third, the size and<br />

the chemical nature <strong>of</strong> the <strong>release</strong>d components play a<br />

role. Additionally, the composition (filler loading) <strong>of</strong> the<br />

composite materials directly influences the elution <strong>of</strong><br />

<strong>monomers</strong> (5). The filler content, the filler size, and the<br />

distribution <strong>of</strong> the filler particles influence the physical<br />

and mechanical properties <strong>of</strong> the composite resins. It has<br />

been shown that the filler volume fraction and filler load<br />

level <strong>of</strong> the composites correlate with the material<br />

strength and elastic modulus, as well as the fracture<br />

toughness <strong>of</strong> the material (6, 7). Increasing the filler<br />

content and reducing the average filler size have been one<br />

approach used to produce composite resins for posterior<br />

restorations. Advanced, highly filled composite resins<br />

exhibit excellent material properties and a good clinical<br />

performance (8).


Monomer <strong>release</strong> <strong>from</strong> composite materials 69<br />

In various studies (4, 9) the degree <strong>of</strong> monomer–<br />

polymer conversion in <strong>dental</strong>-composite resins has been<br />

found to vary between approximately 35 and 77%.<br />

Several authors investigated a possible correlation<br />

between the degree <strong>of</strong> conversion <strong>from</strong> the monomer to<br />

polymer which de<strong>term</strong>ine the chemical stability and<br />

solubility. Pearson & <strong>Long</strong>man (10) showed that<br />

reduced irradiation significantly increases the solubility.<br />

Tanaka et al. (1) found that increasing the irradiation<br />

time <strong>from</strong> 30 to 50 s resulted in a significant decrease in<br />

residual monomer contents and the quantities that were<br />

<strong>release</strong>d into water. In contrast to the above studies,<br />

Ferracane (4) found a very poor correlation between<br />

the degree <strong>of</strong> conversion and solubility.<br />

The elution <strong>of</strong> substances <strong>from</strong> composite materials<br />

has been widely studied (11–15). Geurtsen et al. (13)<br />

evaluated the <strong>release</strong> <strong>of</strong> substances <strong>from</strong> restorative<br />

materials and also the cytotoxicity <strong>of</strong> the <strong>release</strong>d<br />

<strong>monomers</strong> on fibroblasts. In their study, substances were<br />

identified that caused severe or minor cytotoxic effects in<br />

all permanent and primary cell lines. Additionally, their<br />

results indicate that for most <strong>of</strong> the severely toxic composite<br />

components, few cytotoxic alternatives are available.<br />

Spahl et al. (16) identified (co)<strong>monomers</strong>,<br />

additives, and contaminants <strong>from</strong> the manufacturing<br />

process to be <strong>release</strong>d <strong>from</strong> resin composites. Several<br />

in vitro studies (12, 17) have shown the cytotoxic, genotoxic,<br />

mutagenic or estrogenic effects and pulpal and<br />

gingival/oral mucosa reactions <strong>of</strong> some <strong>of</strong> the <strong>monomers</strong><br />

<strong>release</strong>d <strong>from</strong> the composite materials. However, most <strong>of</strong><br />

the in vitro studies (14, 18) have used experimental<br />

composite materials to investigate the process <strong>of</strong> monomer<br />

elution, leading to useful information about commercial<br />

products.<br />

Several studies have been conducted in water and in<br />

oral/food-simulating liquids, among which 75% ethanol/water<br />

solution has been used. This solution is<br />

recommended by the Food and Drug Administration<br />

(FDA) Guidelines <strong>of</strong> the USA (1976, 1988) as a food<br />

simulator and might be considered clinically relevant<br />

(19). It simulates the influence <strong>of</strong> certain beverages<br />

(including alcoholic), vegetables, fruits, and syrup.<br />

Different studies (9, 12) indicate the important role <strong>of</strong><br />

the type <strong>of</strong> the solvent, and that irreversible processes,<br />

such as leaching <strong>of</strong> components, may occur (e.g. in the<br />

presence <strong>of</strong> ethanol). A close match in the solubility<br />

parameter results in maximum s<strong>of</strong>tening <strong>of</strong> the resin<br />

(20, 21). This phenomenon may contribute to irreversible<br />

material degradation provoked by ethanol because<br />

it penetrates the matrix and expands the space between<br />

polymer chains into which soluble substances, such as<br />

unreacted <strong>monomers</strong>, may diffuse. A 75% ethanol<br />

solution has a solubility parameter which matches that<br />

<strong>of</strong> BisGMA (20, 21).<br />

Another point <strong>of</strong> concern about the elution process <strong>of</strong><br />

<strong>monomers</strong> <strong>of</strong> the resin composites is the method used to<br />

de<strong>term</strong>ine the quality and the quantity <strong>of</strong> the eluted<br />

<strong>monomers</strong>. Analysis is usually performed with highperformance<br />

liquid chromatography (HPLC) (14, 22). In<br />

this case, pure components are used as external standards<br />

for the identification and quantification <strong>of</strong> the<br />

eluted components. However, as reported previously<br />

(12), this method has some disadvantages and it is not<br />

suitable to give an exact identification <strong>of</strong> the substances<br />

eluted. Liquid chromatography–mass spectrometry<br />

(LC–MS) is an analytical technique that combines the<br />

advantages <strong>of</strong> an LC instrument with those <strong>of</strong> a mass<br />

spectrometer. High-performance liquid chromatography<br />

is a means <strong>of</strong> separating components <strong>of</strong> mixtures by<br />

passing them through a chromatographic column so that<br />

they emerge sequentially. Because LC operates in the<br />

liquid phase but MS is a gas-phase method, it is not a<br />

simple matter to connect the two. In order to allow the<br />

separated components <strong>of</strong> a mixture to be passed<br />

sequentially <strong>from</strong> the LC into the MS, an interface is<br />

needed. Having interfaced the two, LC–MS is a very<br />

powerful technique (23). A combination <strong>of</strong> HPLC (or<br />

LC) with MS can be used in these cases for a very specific<br />

identification <strong>of</strong> the compounds eluted (15). The fact that<br />

most <strong>of</strong> the studies made in this field were performed<br />

without LC–MS highlights the need for further research<br />

in this field.<br />

Although the advantages <strong>of</strong> the light-cured composites<br />

cannot be questioned, the chemically cured composites<br />

still have important applications in contemporary<br />

restorative dentistry, especially in areas that are not<br />

easily penetrable by light (24). This low conversion rate<br />

(17–24) <strong>of</strong> these materials could result in an increased<br />

<strong>release</strong> <strong>of</strong> <strong>monomers</strong>. Core build-up materials may be in<br />

contact with the oral cavity for a long period <strong>of</strong> time<br />

before crown preparation is performed. No information<br />

on the elution <strong>of</strong> <strong>monomers</strong> <strong>from</strong> core build-up materials<br />

exists in the literature.<br />

Although much literature has been published on the<br />

<strong>release</strong> <strong>of</strong> <strong>monomers</strong> (3, 11–13, 15, 25), information is<br />

still lacking with respect to the elution <strong>of</strong> <strong>monomers</strong><br />

<strong>from</strong> <strong>modern</strong> composite materials such as ormocers,<br />

nanohybrid composite resins, and core build-up materials<br />

that are used daily in the <strong>dental</strong> clinical practice.<br />

Therefore, new studies are necessary to evaluate these<br />

materials under clinically relevant conditions. Additionally,<br />

no information is available on the <strong>release</strong> <strong>of</strong><br />

<strong>monomers</strong> over a post-cure period <strong>of</strong> longer than<br />

3 months. A longer storage time period simulating clinical<br />

oral conditions would be very useful in order to<br />

evaluate the composite materials. However, to correspond<br />

to the period that the composite restorations are<br />

expected to remain in the mouth, a storage time <strong>of</strong><br />

5–10 yr would be necessary. This would not be particularly<br />

useful for studying composite materials because<br />

each year new composite materials become commercially<br />

available and therefore results concerning 10-yr-old<br />

composite materials would not be clinically relevant.<br />

Therefore, the aim <strong>of</strong> the present in vitro study was to<br />

identify and quantify the elution <strong>of</strong> <strong>monomers</strong> <strong>of</strong> three<br />

different resin composites using liquid chromatography<br />

tandem mass spectrometry (LC–MS/MS), after different<br />

curing times and storage periods up to 1 yr after curing.<br />

Additionally, the following hypothesis was examined:<br />

composite materials with new chemistry, such as ormocers,<br />

<strong>release</strong> fewer <strong>monomers</strong> than materials with traditional<br />

chemistry.


70 Polydorou et al.<br />

Table 1<br />

Composite materials tested<br />

Material Type Main monomer composition Manufacturer<br />

Filtek<br />

Supreme XT<br />

Ceram X<br />

Clearfil Core<br />

Universal restorative,<br />

nanohybrid<br />

ÔOrmocerÕ, universal nano-ceramic<br />

restorative<br />

Self-curing composite core<br />

build-up material<br />

BisGMA, TEGDMA, UDMA, BisEMA<br />

Methacrylate modified polysiloxane,<br />

dimethacrylate resin<br />

TEGDMA, BisGMA<br />

3M ESPE Dental Products,<br />

Seefeld, Germany<br />

<strong>Dentsply</strong> DeTrey, Konstanz, Germany<br />

Kuraray Europe,<br />

Du¨ sseldorf, Germany<br />

Material and methods<br />

Three different resin-composite materials were used: a<br />

nanohybrid (Filtek Supreme XT; 3M ESPE Dental Products,<br />

Seefeld, Germany), an ormocer (Ceram X; <strong>Dentsply</strong><br />

DeTrey, Konstanz, Germany) and a chemically cured<br />

(Clearfil Core; Kuraray Europe, Du¨ sseldorf, Germany)<br />

resin composite. Detailed information about the composition<br />

<strong>of</strong> the composite materials and the manufacturers are<br />

given in Table 1.<br />

For curing the samples <strong>of</strong> the two light-cured resin<br />

composites (Filtek Supreme XT and Ceram X), a halogen<br />

unit (Elipar Highlight; 3M ESPE) was used. The light<br />

intensity <strong>of</strong> this unit was 780–800 mW cm )2 . The spectral<br />

irradiance was de<strong>term</strong>ined using a visible curing light meter<br />

(Cure Rite; <strong>Dentsply</strong>, York, USA).<br />

From each <strong>of</strong> the two light-cured composite materials<br />

(Filtek Supreme XT and Ceram X), four groups (n = 10)<br />

were prepared, one for each curing time tested (0, 20, 40, and<br />

80 s). From the chemically cured resin composite (Clearfil<br />

Core) 10 specimens were fabricated. The specimens were<br />

prepared using forms, given <strong>from</strong> <strong>Dentsply</strong> DeTrey, allowing<br />

the production <strong>of</strong> standardized cylindrical specimens<br />

(4.5 mm in diameter and 2 mm thick). The forms were<br />

positioned on a transparent plastic matrix strip lying on a<br />

glass plate. Then, they were filled with the composite materials.<br />

The samples were built up in one increment (2 mm).<br />

After inserting the materials into the discs, a transparent<br />

plastic matrix strip (Kerr Hawe, Bioggio, Switzerland) was<br />

placed on top <strong>of</strong> them in order to avoid creating an oxygeninhibited<br />

superficial layer. In addition, a glass slide was used<br />

in order to flatten the surface. Every sample was then cured<br />

according to the group to which it belonged.<br />

Directly after curing, each sample was immediately<br />

immersed in 1 ml <strong>of</strong> 75% ethanol. The samples were stored<br />

in a dark box at room temperature (’21°C) and the storage<br />

medium (1 ml) was renewed 24 h, 7 d, 28 d, and 1 yr after<br />

the polymerization. From the storage medium removed,<br />

ethanol samples were prepared and stored until analysis at<br />

4°C in the dark.<br />

Analytical method<br />

Information about the substances used as standards for the<br />

analysis are given in Table 2. The reference standards <strong>of</strong><br />

BPA, BisGMA, one UDMA, and TEGDMA were<br />

purchased <strong>from</strong> <strong>Dentsply</strong> DeTrey. A second type <strong>of</strong> UDMA<br />

was obtained <strong>from</strong> Deltamed (Freidberg, Germany). The<br />

two types <strong>of</strong> UDMA differ in chemical composition and<br />

molecular weight. Usually, different types <strong>of</strong> UDMA are<br />

used for the production <strong>of</strong> composite materials. The analysis<br />

was performed as described in detail by Polydorou et al.<br />

(15) with the difference being that the limits <strong>of</strong> quantification<br />

were improved: 1 lg ml )1 for UDMA, 0.5 lg ml )1 for<br />

TEGDMA,1 lg ml )1 for BisGMA, and 0.5 lg ml )1 for<br />

BPA. Values lower than these levels could not be qualified.<br />

For the analysis, the ethanol samples were investigated<br />

using LC–MS/MS. A Bruker Esquire 3000 plus (Bruker-<br />

Franzen Analytik, Bremen, Germany) was used. Separation<br />

<strong>of</strong> the <strong>monomers</strong> was performed using a CC 125/4 Nucleodur<br />

100-5 C18ec HLC-Column (Machery & Mayiel,<br />

Du¨ ren, Germany) and a solvent gradient <strong>of</strong> 0.1% formic<br />

acid and acetonitrile. For the analysis, external calibrations<br />

with ethanol standards (up to a 1,000 lg ml )1 ) were carried<br />

out with the help <strong>of</strong> the peak areas <strong>of</strong> the extracted ion<br />

chromatograms (EICs). Table 3 presents the identification<br />

masses that were used for the qualification and the quantification<br />

<strong>of</strong> the substances during LC–MS/MS.<br />

Statistical analysis<br />

The significance level <strong>of</strong> the statistical analysis was 0.05.<br />

For the statistical analysis, the repeated-measures analysis<br />

<strong>of</strong> variance (anova) was used with two between-factors<br />

(materials and <strong>monomers</strong>) and two within-factors (curing<br />

time and storage time). Because the data were not normally<br />

distributed, the logarithms were used for the statistical<br />

analysis. According to the Ôcorrelation procedureÕ, the<br />

log-transformed data were normally distributed<br />

(P = 0.27).<br />

Table 2<br />

Monomers used as reference standards<br />

Substances Name Chemical type Mol. weight CAS-number<br />

BisGMA Bisphenol A glycidyl methacrylate C 29 H 36 O 8 513.0 1565-94-2<br />

TEGDMA Triethylene glycol dimethacrylate C 14 H 22 O 6 286.32 109-16-0<br />

UDMA 1 Urethane dimethacrylate Product C 26 H 42 O 8 N 2 498.0 –<br />

UDMA 2 Urethane dimethacrylate C 23 H 38 N 2 O 8 470.56 41137-60-4 or 72869-86-4<br />

BPA Bisphenol A C 15 H 16 O 2 228.29 80-05-7


Monomer <strong>release</strong> <strong>from</strong> composite materials 71<br />

Table 3<br />

Identification masses <strong>of</strong> <strong>monomers</strong> in liquid chromatography<br />

tandem mass spectrometry (LC–MS/MS)<br />

Substances Qualification Quantification<br />

BisGMA 512 495<br />

TEGDMA 287 113<br />

UDMA 1 499 499<br />

UDMA 2 471 471<br />

Bisphenol A 227 227<br />

BisGMA, bisphenol A glycidyl methacrylate; TEGDMA,<br />

triethylene glycol dimethacrylate; UDMA, urethane dimethacrylate.<br />

Results<br />

The logarithmic values and the standard deviations <strong>of</strong><br />

the substances <strong>release</strong>d <strong>from</strong> the three composite materials,<br />

for each curing time and storage period, are given<br />

in Tables 4–6.<br />

All <strong>monomers</strong> used as reference standards could be<br />

identified in the ethanol samples <strong>of</strong> the composite<br />

materials. However, differences were found in the type <strong>of</strong><br />

<strong>monomers</strong> <strong>release</strong>d <strong>from</strong> each material. BisGMA,<br />

TEGDMA, BPA, and one form <strong>of</strong> UDMA (UDMA 2)<br />

could be detected in the ethanol samples <strong>of</strong> Filtek<br />

Supreme XT. BisGMA, TEGDMA, and Bisphenol A<br />

were <strong>release</strong>d <strong>from</strong> Ceram X; and only BisGMA and<br />

TEGDMA were detected <strong>from</strong> samples <strong>of</strong> Clearfil Core.<br />

The amount <strong>of</strong> BisGMA eluted was always significantly<br />

higher (P < 0.0001) than the amount <strong>of</strong> TEG-<br />

DMA eluted, regardless <strong>of</strong> curing time, type <strong>of</strong> material,<br />

and storage period. According to the statistical analysis<br />

there was a significant difference between the two lightcured<br />

resin composites (P < 0.0001). More BisGMA<br />

and TEGDMA were <strong>release</strong>d <strong>from</strong> Filtek Supreme XT<br />

compared with Ceram X for each storage time tested.<br />

For these two <strong>monomers</strong>, the following results were<br />

obtained.<br />

For Filtek Supreme XT, the curing time had a significant<br />

effect on the elution <strong>of</strong> <strong>monomers</strong> (P < 0.0001). A<br />

smaller amount <strong>of</strong> <strong>monomers</strong> was <strong>release</strong>d after<br />

increasing the polymerization time. However, a curing<br />

time <strong>of</strong> 80 s did not seem to reduce the elution <strong>of</strong><br />

<strong>monomers</strong> compared with a curing time <strong>of</strong> 40 s. The<br />

elution <strong>of</strong> <strong>monomers</strong> differed significantly during the<br />

storage periods (24 h until 1 yr) (P < 0.0001). During<br />

storage, the <strong>release</strong> <strong>of</strong> TEGDMA decreased significantly,<br />

whereas the amount <strong>of</strong> BisGMA <strong>release</strong>d <strong>from</strong> the<br />

polymerized samples remained high, even 1 yr after<br />

curing.<br />

For Ceram X, the curing time exerted a significant<br />

effect on the elution <strong>of</strong> BisGMA and TEGDMA<br />

(P = 0.0004). Curing for 80 s resulted in the <strong>release</strong> <strong>of</strong> a<br />

significantly higher amount <strong>of</strong> <strong>monomers</strong>. The amount<br />

<strong>of</strong> BisGMA and TEGDMA <strong>release</strong>d <strong>from</strong> Ceram X was<br />

significantly influenced by the storage time <strong>of</strong> the composite<br />

samples (P = 0.01). However, the effect <strong>of</strong> the<br />

storage time was significantly different for the two<br />

Table 4<br />

Filtek Supreme XT: log values <strong>of</strong> the concentration ± standard deviation (SD)<br />

24 h 7 d 28 d 1 yr<br />

0s<br />

BisGMA 3.52 ± 0.03 2.22 ± 0.1 1.94 ± 0.26 1.13 ± 0.14<br />

TEGDMA 2.7 ± 0.025 2.67 ± 0.08 1.12 ± 0.54 0.48 ± 0.15<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 2.14 ± 0.03 1.64 ± 0.03 0.7 ± 0.35 1.07 ± 0.18<br />

Bisphenol A ND ND ND<br />

20 s<br />

BisGMA 2.01 ± 0.14 1.59 ± 0.1 1.59 ± 0.07 1.95 ± 0.07<br />

TEGDMA 1.4 ± 0.25 1.16 ± 0.42 0.75 ± 0.05 0.59 ± 0.07<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 1.3 ± 0.08 0.65 ± 0.06 0.67 ± 0.06 2.16 ± 0.04<br />

Bisphenol A ND ND ND ND<br />

40 s<br />

BisGMA 1.82 ± 0.06 1.36 ± 0.11 1.43 ± 0.08 1.9 ± 0.06<br />

TEGDMA 1.16 ± 0.42 1.21 ± 0.07 0.66 ± 0.03 0.55 ± 0.09<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 1.14 ± 0.06 0.44 ± 0.1 0.54 ± 0.05 2.12 ± 0.08<br />

Bisphenol A ND ND ND ND<br />

80 s<br />

BisGMA 1.76 ± 0.06 1.22 ± 0.09 1.16 ± 0.3 1.69 ± 0.28<br />

TEGDMA 1.1 ± 0.11 1.07 ± 0.13 0.55 ± 0.19 0.41 ± 0.18<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 1.04 ± 0.06 0.29 ± 0.09 0.36 ± 0.02 1.84 ± 0.47<br />

Bisphenol A ND ND ND ND<br />

Logarithmic (log) values <strong>of</strong> the concentration and SDs <strong>of</strong> the <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> Filtek Supreme XT.<br />

BisGMA, bisphenol A glycidyl methacrylate; ND, not detected; TEGDMA, triethylene glycol dimethacrylate; UDMA, urethane<br />

dimethacrylate.


72 Polydorou et al.<br />

Table 5<br />

Ceram X: log values <strong>of</strong> the concentration ± standard deviation (SD)<br />

24 h 7 d 28 d 1 yr<br />

0s<br />

BisGMA 0.59 ± 0.1 0.99 ± 0.05 0.61 ± 0.02 0.51 ± 0.07<br />

TEGDMA 0.09 ± 0.14 0.38 ± 0.08 0.05 ± 0.1 ND<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 ND ND ND ND<br />

Bisphenol A 0.65 ± 0.03 1.37 ± 0.06 0.96 ± 0.08 ND<br />

20 s<br />

BisGMA 0.58 ± 0.06 0.66 ± 0.02 0.56 ± 0.02 0.84 ± 0.04<br />

TEGDMA 0.05 ± 0.1 0.35 ± 0.06 ND ND<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 ND ND ND ND<br />

Bisphenol A 0.72 ± 0.03 0.76 ± 0.04 0.85 ± 0.1 ND<br />

40 s<br />

BisGMA 0.64 ± 0.14 0.65 ± 0.04 0.54 ± 0.19 0.91 ± 0.1<br />

TEGDMA 0.13 ± 0.15 0.35 ± 0.09 ND ND<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 ND ND ND ND<br />

Bisphenol A 0.71 ± 0.06 0.59 ± 0.31 0.32 ± 0.14 ND<br />

80 s<br />

BisGMA 1.28 ± 0.02 0.6 ± 0.01 0.53 ± 0.19 0.84 ± 0.04<br />

TEGDMA 0.42 ± 0.13 0.24 ± 0.04 ND ND<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 ND ND ND ND<br />

Bisphenol A ND 0.5 ± 0.27 0.28 ± 0.16 ND<br />

Logarithmic (Log) values <strong>of</strong> the concentration and SDs <strong>of</strong> the <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> Ceram X.<br />

BisGMA, bisphenol A glycidyl methacrylate; ND, not detected; TEGDMA, triethylene glycol dimethacrylate; UDMA, urethane<br />

dimethacrylate.<br />

<strong>monomers</strong> (P < 0.0001). No TEGDMA could be<br />

detected 1 yr after curing, while at the same time-point<br />

the <strong>release</strong> <strong>of</strong> BisGMA remained at detectable levels.<br />

Bisphenol A was detected only in the polymerized<br />

samples <strong>of</strong> Ceram X. The amount <strong>of</strong> BPA was not<br />

influenced by the curing time (P > 0.05). The second<br />

type <strong>of</strong> UDMA (UDMA 2) was <strong>release</strong>d by the samples<br />

<strong>of</strong> Filtek Supreme XT. Its <strong>release</strong> was influenced by<br />

curing and storage time (P £ 0.05). Increasing the<br />

curing time reduced the amount <strong>of</strong> UDMA 2 eluted,<br />

whereas storage for 1 yr resulted in a significantly higher<br />

<strong>release</strong>.<br />

The amount <strong>of</strong> BisGMA <strong>release</strong>d <strong>from</strong> Clearfil Core<br />

was higher compared with the amount <strong>of</strong> TEGDMA<br />

<strong>release</strong>d. The amount <strong>of</strong> BisGMA and TEGDMA eluted<br />

<strong>from</strong> Clearfil Core was compared with the amount <strong>of</strong><br />

BisGMA and TEGDMA <strong>release</strong>d <strong>from</strong> Filtek Supreme<br />

XT and Ceram X cured for 40 s and 80 s. For both<br />

curing times, a significant difference (P £ 0.05) was<br />

found between the three composite materials concerning<br />

the elution <strong>of</strong> BisGMA and TEGDMA, namely that the<br />

amounts <strong>of</strong> BisGMA and TEGDMA <strong>release</strong>d <strong>from</strong><br />

Clearfil Core were higher than the amounts <strong>of</strong> BisGMA<br />

and TEGDMA <strong>release</strong>d <strong>from</strong> Filtek Supreme XT and<br />

Ceram X.<br />

Discussion<br />

In the present study, monomer <strong>release</strong> <strong>from</strong> three different<br />

<strong>dental</strong>-composite materials was investigated. The<br />

highest elution <strong>of</strong> <strong>monomers</strong> occurred <strong>from</strong> Filtek<br />

Table 6<br />

Clearfil Core: log values <strong>of</strong> the concentration ± standard deviation (SD)<br />

24 h 7 d 28 d 1 yr<br />

BisGMA 2.68 ± 0.06 1.74 ± 0.14 1.29 ± 0.03 2.24 ± 0.09<br />

TEGDMA 2.39 ± 0.45 1.42 ± 0.29 0.86 ± 0.03 0.96 ± 0.07<br />

UDMA 1 ND ND ND ND<br />

UDMA 2 ND ND ND ND<br />

Bisphenol A ND ND ND ND<br />

Logarithmic (log) values <strong>of</strong> the concentration and SDs <strong>of</strong> the <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> Clearfil Core.<br />

BisGMA, bisphenol A glycidyl methacrylate; ND, not detected; TEGDMA, triethylene glycol dimethacrylate; UDMA, urethane<br />

dimethacrylate.


Monomer <strong>release</strong> <strong>from</strong> composite materials 73<br />

Supreme XT, followed closely by Clearfil Core. Very<br />

small amounts <strong>of</strong> <strong>monomers</strong> were eluted <strong>from</strong> Ceram X.<br />

These findings support the results <strong>of</strong> Franz et al. (26),<br />

who demonstrated that Filtek Supreme XT exhibits<br />

higher cytotoxicity on L-929 fibroblasts than the other<br />

composite materials tested.<br />

According to the manufacturersÕ information, the filler<br />

content <strong>of</strong> these materials is as follows: Ceram X, filler<br />

content 76% by weight; Clearfil Core, filler content 78%<br />

by weight; and Filtek Supreme XT, filler content 78.5%<br />

by weight. These small differences in the filler loading<br />

between the three materials could not be the reason for<br />

our findings. It is more likely that the difference in the<br />

chemistry between the three materials, the filler size, and<br />

the distribution <strong>of</strong> the filler particles could have influenced<br />

our results.<br />

As far as Clearfil Core is concerned, no published data<br />

exist concerning the elution <strong>of</strong> <strong>monomers</strong>. The decreased<br />

degree <strong>of</strong> conversion <strong>of</strong> the chemically cured composites<br />

compared with the light-cured materials (27) could<br />

explain the high degree <strong>of</strong> elution <strong>of</strong> <strong>monomers</strong> <strong>from</strong><br />

Clearfil Core. Storage <strong>of</strong> the samples in ethanol immediately<br />

after their manufacture could have influenced the<br />

<strong>release</strong> <strong>of</strong> TEGDMA and BisGMA because <strong>of</strong> the fact<br />

that the degree <strong>of</strong> conversion Clearfil Core at the end <strong>of</strong><br />

mixing is markedly lower than that <strong>of</strong> photoinitiated<br />

systems. Although such a comparison <strong>of</strong> the composite<br />

materials directly after mixing might be considered to be<br />

quite unfair for the chemically cured composite, as<br />

mentioned above, the use <strong>of</strong> this solution for sample<br />

storage simulates the situation in the mouth. Chemicallycured<br />

materials are generally used as core build-up<br />

materials and therefore they should not come into direct<br />

contact with saliva or oral s<strong>of</strong>t tissues because they are<br />

covered with a crown. However, these materials very<br />

<strong>of</strong>ten remain in the oral cavity for a considerable period<br />

<strong>of</strong> time before they are covered with the final restoration.<br />

Additionally, it is possible that <strong>monomers</strong> can diffuse<br />

through dentin into the pulp (28, 29).<br />

In the present study, the ethanol samples were analysed<br />

using LC–MS/MS. Using this method, substances<br />

with the same mass and retention time can be identified.<br />

The use <strong>of</strong> MS/MS has the benefits <strong>of</strong> separating different<br />

molecular ions, generating fragment ions <strong>from</strong> a<br />

selected ion, and then mass measuring the fragmented<br />

ions. The fragmental ions are used for structural de<strong>term</strong>ination<br />

<strong>of</strong> original molecular ions (23). Therefore, a<br />

combination <strong>of</strong> LC and MS in the form <strong>of</strong> LC–MS/MS<br />

can be very helpful to identify specific compounds with<br />

similar structures, which is one <strong>of</strong> the greatest problems<br />

encountered in analytical methods. In the present study,<br />

the identification <strong>of</strong> specific molecules was studied.<br />

However, it should be mentioned that beside these<br />

molecules, other substances and degradation products<br />

could be <strong>release</strong>d <strong>from</strong> the composite materials. The use<br />

<strong>of</strong> LC–MS/MS could be very helpful in their identification.<br />

An important finding <strong>of</strong> the present study was the<br />

small amount <strong>of</strong> <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> Ceram X.<br />

Ormocer contains inorganic–organic copolymers in<br />

addition to the inorganic silanized filler particles. It is<br />

synthesized through solution and gelation processes (sol–<br />

gel process) <strong>from</strong> multifunctional urethane and thioether(meth)acrylate<br />

alkoxysilanes. By hydrolysis and<br />

polycondensation reactions, the alkoxysil groups <strong>of</strong> the<br />

silane promote the formation <strong>of</strong> an inorganic Si–O–Si<br />

network, while the methacrylate groups are still available<br />

for light-activated organic polymerization. Ormocers are<br />

characterized by the novel inorganic–organic copolymers<br />

in the formulation that allow the modification <strong>of</strong><br />

mechanical properties over a wide range (30). Hickel<br />

et al. (31) suggested that the advantages <strong>of</strong> ormocer<br />

include low shrinkage, high abrasion resistance, biocompatibility,<br />

and caries protection. However, only<br />

limited information is available on the elution <strong>of</strong><br />

<strong>monomers</strong> <strong>from</strong> ormocers and their toxicity. Brackett<br />

et al. (32) found that materials with traditional compositions,<br />

such as Filtek Supreme XT, were severely cytotoxic<br />

throughout an 8-wk interval, but materials with<br />

newer chemistries or filling strategies, such as Ceram X,<br />

improved over time <strong>of</strong> aging in artificial saliva. This<br />

appears to support the claims <strong>of</strong> better polymerization<br />

and lower leaching in newer chemistries (33). The difference<br />

in the chemical structure <strong>of</strong> the composite, and<br />

the variation in the ratio <strong>of</strong> filler and monomer, have a<br />

significant effect on the element <strong>release</strong> and cytotoxicity<br />

level <strong>of</strong> the materials (34).<br />

Regardless <strong>of</strong> the polymerization time, storage time,<br />

and material, a greater amount <strong>of</strong> BisGMA was <strong>release</strong>d<br />

compared with the other <strong>monomers</strong>. As mentioned<br />

above, 75% ethanol has a solubility parameter which<br />

matches that <strong>of</strong> BisGMA and this could be the reason for<br />

the higher amount <strong>of</strong> BisGMA <strong>release</strong>d. Additionally,<br />

BisGMA and TEGDMA differ markedly as far as their<br />

chemical properties and reactive potentials are concerned.<br />

(35). BisGMA is more reactive than TEGDMA,<br />

but the fact that it presents a sterically hindered structure<br />

precludes a higher degree <strong>of</strong> conversion (36). It has been<br />

shown that the quantity <strong>of</strong> unreacted unsaturations in<br />

the BisGMA/TEGDMA system can be qualitatively<br />

correlated with the amount <strong>of</strong> BisGMA in the polymerizing<br />

system (37, 38) and that neat TEGDMA<br />

exhibits a degree <strong>of</strong> conversion that is about 10% higher<br />

than that <strong>of</strong> neat BisGMA (37, 38).<br />

In the present study, BPA was found to be eluted by<br />

unpolymerized composite samples <strong>of</strong> Filtek Supreme XT<br />

and Ceram X and by polymerized samples <strong>of</strong> Ceram X.<br />

This is in contrast to previous studies where no BPA was<br />

detected (15, 16, 25). However, detectable amounts <strong>of</strong><br />

BPA have been mentioned, also by other authors (34,<br />

39), to be extracted <strong>from</strong> composite materials. Dental<br />

composites typically contain <strong>monomers</strong> [such as<br />

BisGMA and bisphenol A dimethacrylate (BisDMA)]<br />

that are derived <strong>from</strong> BPA, but there is no known use for<br />

BPA itself in <strong>dental</strong> composites. As these <strong>monomers</strong> may<br />

leach <strong>from</strong> <strong>dental</strong> composites, their stability has been<br />

studied under a variety <strong>of</strong> conditions to de<strong>term</strong>ine whether<br />

they may hydrolyze to form BPA. BisGMA, the<br />

base monomer for many resin composites, has been<br />

found to be stable under various hydrolytic conditions<br />

(25). However, two researchers have reported that Bis-<br />

DMA is hydrolyzed to BPA, which probably accounts


74 Polydorou et al.<br />

for the BPA detected in extracts <strong>from</strong> certain composites<br />

(25–40). Further research is necessary to evaluate the<br />

source <strong>of</strong> BPA <strong>release</strong>d <strong>from</strong> the composite materials.<br />

With respect to the different polymerization times used<br />

in the present study, it could be shown that an increase in<br />

the curing time <strong>from</strong> 20 to 40 s resulted in a decrease in<br />

the amount <strong>of</strong> <strong>monomers</strong> <strong>release</strong>d <strong>from</strong> Filtek Supreme<br />

XT. The increased polymerization time <strong>of</strong> 80 s did<br />

decrease the monomer <strong>release</strong> compared with the polymerization<br />

time <strong>of</strong> 40 s. For Ceram X, the amount <strong>of</strong><br />

<strong>monomers</strong> eluted was shown to increase after polymerization<br />

for 80 s. However, the amount <strong>of</strong> <strong>monomers</strong><br />

<strong>release</strong>d <strong>from</strong> this material was very low, especially when<br />

the samples were polymerized only for 20 s. This result<br />

suggests that Ceram X can be used with short curing<br />

times without exposing the patients to high levels <strong>of</strong><br />

<strong>monomers</strong>. However, further research is necessary to<br />

evaluate the effect <strong>of</strong> the increased curing time on the<br />

chemistry <strong>of</strong> this material. Although this type <strong>of</strong> negative<br />

effect on composite materials after a longer curing time<br />

has not been previously reported, it may be assumed that<br />

the longer exposure time could have a negative effect on<br />

the network <strong>of</strong> the materials, affecting their polymerization<br />

and, as a result, making the special chemistry <strong>of</strong><br />

these materials a disadvantage. The results <strong>of</strong> the present<br />

experiments did underline the conclusion that for different<br />

composites, polymerization times <strong>of</strong> < 80 s are<br />

not sufficient to prohibit monomer <strong>release</strong>.<br />

Within the limitations <strong>of</strong> the present study, the<br />

composite materials tested as a result <strong>of</strong> having marked<br />

differences in their composition could be considered as a<br />

model for a wide variety <strong>of</strong> materials in clinical use. As<br />

far as the hypothesis made at the start <strong>of</strong> the study is<br />

concerned, it can be stated that materials with new<br />

chemistries, like ormocer, can show, after a short timeperiod<br />

<strong>of</strong> curing, a low elution <strong>of</strong> <strong>monomers</strong> compared<br />

to composites with traditional compositions and chemically<br />

cured materials. However, this hypothesis cannot<br />

be totally accepted because the behavior <strong>of</strong> the ormocer<br />

was not linear and therefore it cannot be considered as<br />

chemically stable without further research.<br />

Acknowledgements – This work was supported by the<br />

Research Commission <strong>of</strong> the Medical Faculty <strong>of</strong> the University<br />

Freiburg (Forschungskommission), Germany; number<br />

KU¨ M431/05.<br />

References<br />

1. Tanaka K, Taira M, Shintani H, Wasaka K, Yamaki M.<br />

Residual <strong>monomers</strong> (TEG-DMA and Bis-GMA) <strong>of</strong> a set visible-light-cured<br />

<strong>dental</strong> resin composite when immersed in water.<br />

J Oral Rehabil 1991; 18: 353–362.<br />

2. Moszner N, Salz U. New developments <strong>of</strong> polymer <strong>dental</strong><br />

composites. Prog Polym Sci 2001; 26: 535–576.<br />

3. Moharamzadeh K, Van Noort R, Brook IM, Scutt AM.<br />

HPLC analysis <strong>of</strong> components <strong>release</strong>d <strong>from</strong> <strong>dental</strong> composites<br />

with different resin compositions using different extraction<br />

media. J Mater Sci Mater Med 2007; 18: 133–137.<br />

4. Ferracane JL. Elution <strong>of</strong> leachable components <strong>from</strong> composites.<br />

J Oral Rehabil 1994; 21: 441–452.<br />

5. Wataha JC, Rueggeberg FA, Lapp CA, Lewis JB, Lockwood<br />

PE, Ergle JW, Mettenburg DJ. In vitro cytotoxicity <strong>of</strong> resincontaining<br />

restorative materials after aging in artificial saliva.<br />

Clin Oral Invest 1999; 3: 144–149.<br />

6. St Germain H, Swartz ML, Phillips RW, Moore BK, Roberts<br />

TA. Properties <strong>of</strong> micr<strong>of</strong>illed composite resins as influenced<br />

by filler content. J Dent Res 1985; 64: 155–160.<br />

7. Li Y, Swartz ML, Phillips RW, Moore BK, Roberts TA.<br />

Effect <strong>of</strong> filler content and size on properties <strong>of</strong> composites.<br />

J Dent Res 1985; 64: 1396–1401.<br />

8. Leinfelder KF. Posterior composite resins: the materials and<br />

their clinical performance. J Am Dent Assoc 1995; 126: 663–664,<br />

667–778, 671–672.Review.<br />

9. Ferracane JL, Condon JR. Rate <strong>of</strong> elution <strong>of</strong> leachable<br />

components <strong>from</strong> composite. Dent Mater 1990; 6: 282–287.<br />

10. Pearson GJ, <strong>Long</strong>man CM. Water sorption and solubility <strong>of</strong><br />

resin-based materials following inadequate polymerization by a<br />

visible-light curing system. J Oral Rehabil 1989; 16: 57–61.<br />

11. Lee SY, Greener EH, Menis DL. Detection <strong>of</strong> leached moieties<br />

<strong>from</strong> <strong>dental</strong> composites in fluids stimulating food and<br />

saliva. Dent Mater 1995; 11: 348–353.<br />

12. Geurtsen W. Substances <strong>release</strong>d <strong>from</strong> <strong>dental</strong> composites and<br />

glass ionomer cements. Eur J Oral Sci 1998; 106: 687–695.<br />

13. Geurtsen W, Lehmann F, Spahl W, Leyhausen G. Cytotoxicity<br />

<strong>of</strong> 35 composite <strong>monomers</strong>/additives in permanent<br />

3T3- and three human oral primary fibroblast cultures. J Biomed<br />

Mater Res 1998; 41: 474–480.<br />

14. Munksgaard EC, Peutzfeldt A, Asmussen E. Elution <strong>of</strong><br />

TEGDMA and BisGMA <strong>from</strong> a resin and a resin composite<br />

cured with halogen or plasma light. Eur J Oral Sci 2000; 108:<br />

341–345.<br />

15. Polydorou O, Trittler R, Hellwig E, Kummerer K. Elution<br />

<strong>of</strong> <strong>monomers</strong> <strong>from</strong> two conventional <strong>dental</strong> composite materials.<br />

Dent Mater 2007; 23: 1535–1541.<br />

16. Spahl W, Budzikiewicz H, Geurtsen W. De<strong>term</strong>ination <strong>of</strong><br />

leachable components <strong>from</strong> four commercial <strong>dental</strong> composites<br />

by gas and liquid chromatography/mass spectrometry. J Dent<br />

1998; 26: 137–145.<br />

17. Kleinsasser NH, Schmid K, Sassen AW, Harreus UA,<br />

Staudenmaier R, Folwaczny M, Glas J, Reichl FX. Cytotoxic<br />

and genotoxic effects <strong>of</strong> resin <strong>monomers</strong> in human salivary<br />

gland tissue and lymphocytes as assessed by the single cell<br />

microgel electrophoresis (Comet) assay. Biomaterials 2006; 27:<br />

1762–1770.<br />

18. Leung D, Spratt DA, Pratten J, Gulabivala K, Mordan<br />

NJ, Young AM. Chlorhexidine-releasing methacrylate <strong>dental</strong><br />

composite materials. Biomaterials 2005; 26: 7145–7153.<br />

19. Sideridou ID, Achilias DS, Karabela MM. Sorption kinetics<br />

<strong>of</strong> ethanol/water solution by dimethacrylate-based <strong>dental</strong> resins<br />

and resin composites. J Biomed Mater Res B Appl Biomater<br />

2007; 81: 207–218.<br />

20. Wu W, McKinney JE. Influence <strong>of</strong> chemicals on wear <strong>of</strong> <strong>dental</strong><br />

composites. J Dent Res 1982; 61: 1180–1183.<br />

21. McKinney JE, Wu W. Chemical s<strong>of</strong>tening and wear <strong>of</strong> <strong>dental</strong><br />

composites. J Dent Res 1985; 64: 1326–1331.<br />

22. Lin BA, Jaffer F, Duff MD, Tang YW, Santerre JP.<br />

Identifying enzyme activities within human saliva which are<br />

relevant to <strong>dental</strong> resin composite biodegradation. Biomaterials<br />

2005; 26: 4259–4264.<br />

23. Siuzdak G. The expanding role <strong>of</strong> mass spectrometry in biotechnology.<br />

San Diego, CA: MCC Press, 2003.<br />

24. Bolhuis PB, De Gee AJ, Kleverlaan CJ, El Zohairy AA,<br />

Feilzer AJ. Contraction stress and bond strength to dentinfor<br />

compatible and incompatible combinations <strong>of</strong> bonding systems<br />

and chemical and light-cured core build-up resin composites.<br />

Dent Mater 2006; 22: 223–233.<br />

25. Schmalz G, Preiss A, Arenholt-Bindslev D. Bisphenol-A<br />

content <strong>of</strong> resin <strong>monomers</strong> and related degradation products.<br />

Clin Oral Invest 1999; 3: 114–119.<br />

26. Franz A, Konradsson K, Konig F, Van Dijken JW, Schedle<br />

A. Cytotoxicity <strong>of</strong> a calcium aluminate cement in comparison<br />

with other <strong>dental</strong> cements and resin-based materials. Acta<br />

Odontol Scand 2006; 64: 1–8.<br />

27. Lutz F, Setcos JC, Phillips RW, Roulet JF. Dental restorative<br />

resins. Types and characteristics. Dent Clin N Am 1983;<br />

27: 697–712.


Monomer <strong>release</strong> <strong>from</strong> composite materials 75<br />

28. Hume WR, Gerzina TM. Bioavailability <strong>of</strong> components <strong>of</strong><br />

resin-based materials which are applied to teeth. Review. Crit<br />

Rev Oral Biol Med 1996; 7: 172–179.<br />

29. Gerzina TM, Hume WR. Diffusion <strong>of</strong> <strong>monomers</strong> <strong>from</strong><br />

bonding resin–resin composite combinations through dentine in<br />

vitro. J Dent 1996; 24: 125–128.<br />

30. Wolter H, Storch W, Ott H. New inorganic/organic<br />

copolymers (ORMOCERs) for <strong>dental</strong> applications. Mater Res<br />

Soc Symp Proc 1994; 346: 143–149.<br />

31. Hickel R, Dasch W, Janda R, Tyas M, Anusavice K. New<br />

direct restorative materials. FDI Commission Project. Int Dent<br />

J 1998; 48: 3–16.<br />

32. Brackett MG, Bouillaguet S, Lockwood PE, Rotenberg S,<br />

Lewis JB, Messer RL, Wataha JC. In vitro cytotoxicity <strong>of</strong><br />

<strong>dental</strong> composites based on new and traditional polymerization<br />

chemistries. J Biomed Mater Res B Appl Biomater 2007; 81:<br />

397–402.<br />

33. Eick JD, Kostoryz EL, Rozzi SM, Jacobs DW, Oxman JD,<br />

Chappelow CC, Glaros AG, Yourtee DM. In vitro<br />

biocompatibility <strong>of</strong> oxirane/polyol <strong>dental</strong> composites<br />

with promising physical properties. Dent Mater 2002; 18: 413–<br />

421.<br />

34. Al-Hiyasat AS, Darmani H, Elbetieha AM. Leached<br />

components <strong>from</strong> <strong>dental</strong> composites their effects on fertility<br />

<strong>of</strong> female mice. Eur J Oral Sci 2004; 112: 267–272.<br />

35. Stanbury JW, Dickens SH. De<strong>term</strong>ination <strong>of</strong> double bond<br />

conversion in <strong>dental</strong> resins by near infrared spectroscopy. Dent<br />

Mater 2001; 17: 71–79.<br />

36. Kilambi H, Cramer NB, Schneidewind LH, Shah P, Stansbury<br />

JW, Bowman CN. Evaluation <strong>of</strong> highly reactive monomethacrylates<br />

as reactive diluents for BisGMA-based <strong>dental</strong><br />

composites. Dent Mater 2009; 25: 33–38.<br />

37. Gebeleim CC, Koblitz PP. Polymer science and technology,<br />

Volume 14 – Biomedical and <strong>dental</strong> applications <strong>of</strong> polymers.<br />

New York, NY: Plenum Press, 1981; 379.<br />

38. Jancar J, Wang W, Dibenedetto AT. On the heterogeneous<br />

structure <strong>of</strong> thermally cured bis-GMA/TEGDMA resins.<br />

J Mater Sci Mater Med 2000; 11: 675–682.<br />

39. Pulgar R, Olea-Serrano MF, Novillo-Fertell A, Rivas A,<br />

Pazos P, Pedraza V, Navajas JM, Olea N. De<strong>term</strong>ination <strong>of</strong><br />

bisphenol A and related aromatic <strong>release</strong>d <strong>from</strong> Bis-GMAbased<br />

composites and sealants by high performance liquid<br />

chromatography. Environ Health Perspect 2000; 108: 21–28.<br />

40. Atkinson JC, Diamond F, Eichmiller F, Selwitz R, Jones G.<br />

Stability <strong>of</strong> bisphenol A, triethylene-glycol dimethacrylate, and<br />

bisphenol A dimethacrylate in whole saliva. Dent Mater 2002;<br />

18: 128–135.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!