Sie sind auf Seite 1von 10

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

Available online at www.sciencedirect.com

ScienceDirect
journal homepage: www.intl.elsevierhealth.com/journals/dema

Qualitative and quantitative characterization of


monomers of uncured bulk-ll and conventional
resin-composites using liquid
chromatography/mass spectrometry
Ruwaida Z. Alshali a,b , Nesreen A. Salim c , Rehana Sung d ,
Julian D. Satterthwaite a , Nick Silikas a,
a

School of Dentistry, The University of Manchester, Manchester, UK


Department of Oral and Maxillofacial Rehabilitation, King Abdulaziz University, Jeddah, Saudi Arabia
c Prosthodontic Department, University of Jordan, Amman, Jordan
d Manchester Institute of Biotechnology, The University of Manchester, Manchester, UK
b

a r t i c l e

i n f o

a b s t r a c t

Article history:

Objectives. The aim of this study was to assess the resin matrix monomer composition of

Received 26 January 2015

selected bulk-ll and conventional resin-composite materials using reverse phase liquid

Received in revised form

chromatography coupled with electron spray ionization mass spectrometry.

2 March 2015

Material and methods. Six bulk-ll (SureFil SDR, Venus Bulk Fill, X-tra base, Filtek Bulk Fill

Accepted 24 March 2015

owable, Sonic Fill, and Tetric EvoCeram Bulk Fill) and eight conventional resin-composites
(Grandioso Flow, Venus Diamond Flow, X-Flow, Filtek Supreme XTE, Grandioso, Venus Diamond, TPH Spectrum, and Filtek Z250) were tested. For assessment of resin composition and

Keywords:

relative monomer amounts, uncured resin-composites were analysed with reverse phase

Monomer

liquid chromatography/electron spray ionization mass spectrometry. log P values (a mea-

Liquid chromatography

sure of hydrophobicity) of detected compounds were calculated and their correlation to

Mass spectrometry

reverse phase liquid chromatography retention time was explored. Data were analysed with

Bulk-ll

one-way ANOVA, Tukey post hoc test, Pearson correlation and regression analyses at = 0.05.

Resin-composites

Results. The main monomers detected were BisGMA, UDMA, TEGDMA, and BisEMA.
Monomers were detected at variable combinations in different materials with signicantly
different relative amounts. Other monomers were detected including HDDMA, DEGDMA,
TCD-DI-HEA, and SDR-UDMA in Grandioso ow, X-ow, Venus Diamond, and SureFil SDR
respectively. A positive correlation between log P and reverse phase liquid chromatography
retention time was detected (r2 = 0.62, p = 0.004).
Conclusions. Resin composition of bulk-ll resin-composites is comparable to that of conventional materials with the exception of SureFil SDR. The relative hydrophobicity of dental
monomers can be determined by their reverse phase liquid chromatography retention time.
2015 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

Corresponding author at: School of Dentistry, The University of Manchester, Manchester M13 9PL, UK. Tel.: +44 1612756747.
E-mail address: nick.silikas@manchester.ac.uk (N. Silikas).

http://dx.doi.org/10.1016/j.dental.2015.03.010
0109-5641/ 2015 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

712

1.

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

Introduction

Resin-composites are widely used dental materials since they


restore both function and esthetics of dental tissues [1,2].
The main components of a resin-composite are the inorganic
llers, and the organic resin matrix. The resin matrix constitutes about 2040 wt% of a resin-composite material and is
mainly composed of monomeric compounds including base
monomers, co-monomers and additives [3]. Bowen in the
early 1960s introduced the dimethacrylate monomer Bisphenol A Glycidyl Methacrylate (BisGMA) as the rst mononomer
suitable to be incorporated into a resin-composite formulation intended for use as a direct dental restorative material
[4]. Dental resin-composites have evolved signicantly since
they were rst introduced to improve their strength, wear
resistance, and polymerization shrinkage properties; although
most of the changes have involved the inorganic ller fraction
[5]. The resin fraction; comprised mainly of the base monomer
BisGMA; has largely remained unchanged. BisGMA is a highly
viscous monomer due to its hydroxyl groups which increase
its polarity and result in strong intermolecular attraction [6]. It
is usually diluted with a low molecular weight linear monomer
triethylene glycol dimethacrylate (TEGDMA). Dilution of BisGMA with TEGDMA reduces its viscosity, improves its handling
properties, degree of conversion [79], and allows more ller
fraction to be loaded into the resin. However, BisGMA dilution
with TEGDMA has been shown to be associated with increased
polymerization shrinkage and water sorption [10,11].
Urethane dimethacrylate monomer (UDMA) is a low viscosity base monomer that has molecular weight close to
that of BisGMA, however, the absence of the phenol rings
in UMDA results in a more exible structure with higher
toughness in comparison to BisGMA [12,13]. Ethoxylated
bisphenol A dimethacrylate (BisEMA) is a BisGMA analogue
with low viscosity due to the absence of hydroxyl groups
[6]. BisEMA is available in different molar weights depending on the ethylene oxide chain length between the aromatic
core and the functional methacrylate groups and increasing the degree of ethoxylation of BisEMA further reduces
its viscosity and increases conversion, but also increases
water sorption and decreases exural strength [6,14]. Variable combinations and proportions of monomers are used
in current resin-composites resulting in different copolymer
systems. The long-term hygroscopic and hydrolytic stability of these materials has been shown to be dictated by the
hydrophobicity of their constituent monomers, and by the
nal network characteristics; specically, the degree of conversion, cross-link density and structural heterogeneity of their
copolymer structure [15]. To have a better understanding and
assessment of their effect on the clinical and laboratory performance, adequate qualitative and quantitative data about
a resin-composite constituent monomers and co-monomers
is mandatory. Unfortunately, due to commercial protectionism, these data are usually either missing or not completely
disclosed by the manufacturers [3].
The synthesis of new dental monomers is mainly driven by
the aim of reducing polymerization shrinkage, increasing the
biocompatibility of the materials, and maximizing polymerization potential. Recently, a new class of materials described

as bulk-ll resin-composites has been introduced. Bulk-ll


resin-composites are mainly characterized by their enhanced
depth of cure allowing thick increments (45 mm) to be placed
in a single step instead of the conventional layering technique [16,17]. One such material (SureFil SDR ) is based on a
modied UDMA stress-decreasing resin system with an incorporated polymerization modulator photoactive group [18].
Determination and quantication of different dental resin
components and eluted species has been carried out using the
sensitive techniques of gas chromatography (GC) and high performance liquid chromatography (HPLC) [19]. Both HPLC and
GC can be coupled to a mass spectrometer (MS) as a detector to
increase the sensitivity and selectivity of the technique. This
also allows identication of unknown substances and degradation products from both uncured resins and extracts from
cured materials [20]. Less commonly used techniques include
uorescence spectrophotometry [21] and micellar electrokinetic chromatography [22].
The aim of this study was to assess the resin-matrix
monomer composition of selected bulk-ll and conventional resin-composite materials, using liquid chromatography/mass spectrometry to identify the constituent monomers
of the uncured resin of these materials and their relative proportions. The rst null hypothesis was that there would be no
difference in the type and relative proportions of monomers
that constitute the uncured resin of bulk-ll and conventional
resin-composite materials. The hydrophobicity of identied
monomers and compounds was also assessed by generating
their predicted octanolwater partition coefcient (log P), and
the correlation between predicted log P values of monomers
and their reverse phase liquid chromatography retention time
was explored. Hence, the second null hypothesis was that
there would be no correlation between predicted log P values
and reverse phase liquid chromatography retention time of
the identied compounds.

2.

Materials and methods

2.1.

Solvents and reagents

All solvents in this study were HPLC grade. Water, methanol,


acetonitrile, caffeine, BisEMA, DEGDMA, EBPADA, and MMA
were from SigmaAldrich, UK. BisGMA, TEGDMA, UDMA,
HEMA were supplied by Rhm GmbH, Germany (Table 2).

2.2.

Characterization of uncured resin

Fourteen commercial resin-composite materials including six


bulk-ll materials and eight conventional resin-composite
materials were tested. A list of the resin-composites studied
is given in Table 1. Three samples of each uncured resincomposite material were prepared. For each sample 0.1 g of
uncured resin-composite was dissolved in 1.5 ml methanol
(containing 0.1 mg/ml caffeine as an internal standard). The
samples were shaken for 10 min until the composite was
completely dissolved in the methanol solution. The samples
were then centrifuged at 15,000 g for one minute to separate
the ller particles from the solution (IEC MicroCL 17 centrifuge, Thermo Electron Corporation, USA). The supernatant

713

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

Table 1 Test resin-composite materials and manufacturers: six bulk-ll (light gray) and eight conventional
resin-composite materials (dark gray).
Material

Code

Type

Shade

Filler (wt%)

Lot number

SureFil SDR ow

SDR

Bulk ll, owable

Universal

68

10211

Venus Bulk Fill

VBF

Bulk ll, owable

Universal

65

010101

X-tra base

XB

Bulk ll, owable

Universal

75

1208392

Filtek Bulk Fill owable

FBF

Bulk ll, owable

Universal

64

N370958

Tetric EvoCeram Bulk Fill


SonicFilllTM

TEC
SF

Bulk ll, regular


Bulk ll, regular

Universal A shade(IVA)
A2

77
83

R04686
4964921

Grandioso Flow

GRF

Conventional, owable

A2

81

1305362

Venus Diamond Flow

VDF

Conventional, owable

A3

65

010104

X-Flow

XF

Conventional, owable

A2

60

1267

FiltekTM Supreme XTE

FF

Conventional, owable

A3

65

N522058

Grandioso

GR

Conventional, regular

A2

89

1304304

Venus Diamond

VD

Conventional, regular

A2

81

010046

TPH 3 Spectrum

TPH

Conventional, regular

A2

75

1301000713

FiltekTM Z250

Z250

Conventional

A2

79

N458477

methanol solution containing the dissolved resin was then


quickly withdrawn with a syringe without disturbing the
precipitated llers, passed through a 0.2 m membrane lter, and transferred into an HPLC vial. Resulting solutions
were then analysed without any further work-up by high
performance liquid chromatography and mass spectrometry
(LC/MS). Samples of reference monomers (BisGMA, UDMA,
TEGDMA, BisEMA, EBPADA, HEMA, MMA, and DEGDMA) dissolved in methanol were also analysed to facilitate their
identication in the uncured resin solutions prepared from
different composite materials.
The analysis was carried out using an Agilent 1100LC/MSD
system. This consisted of an Agilent 1100 high performance
liquid chromatograph with diode array detection coupled to
an Agilent 1100 ion trap mass spectrometer (Agilent Technologies, Germany). Chromatography was performed using a
reverse phase Phenomenex SphereClone 5 m ODS (2) column
of dimensions 4.6 mm 250 mm (Phenomenex, USA). An isocratic method of acetonitrile and water (50:50) and ow rate of
0.5 ml/min was used and the UV detection was set at 205 nm.
The column temperature was 22 C with a run time of 70 min
for each sample. The analytes were ionized using electrospray
(ESI) in positive ion mode with an applied capillary voltage of
3 kV. Nebulizing nitrogen gas pressure was 60 psi, drying gas
ow was 10 l/min, and the drying gas temperature was 350 C.
The mass range was set at 501000 m/z. The LC/MS system
was controlled by Chemstation (version B.01.03) and LC/MSD
Trap (version 3.3) software.
Identication of the various compounds was achieved
by comparison of their mass spectra (mass-to-charge-ratio
(m/z)) and chromatographic retention times with those of

Manufacturer
DENTSPLY Caulk,
USA
Heraeus Kulzer
GmbH, Germany
VOCO GmbH,
Germany
3M ESPE GmbH,
Germany
Ivoclar Vivadent
Kerr Corporation,
USA
VOCO GmbH,
Germany
Heraeus Kulzer
GmbH, Germany
DENTSPLY Caulk,
USA
3M ESPE GmbH,
Germany
VOCO GmbH,
Germany
Heraeus Kulzer
GmbH, Germany
DENTSPLY Caulk,
USA
3M ESPE GmbH,
Germany

reference compounds. For compounds with no available reference, the molecular weight was calculated from the highest
intensity m/z value and compared with data in the literature.
To compare the quantity of each compound in different
resin-composite materials, the chromatogram peak area%
of each compound in relation to the total chromatogram
area was calculated. Log P values (octanolwater partition
coefcient) of identied non-fragmented compounds with
known structural formula were computationally obtained
using the ACD/ChemSketch (version 11.02) software.

2.3.

Statistical analysis

Data were entered into statistical analysis software (SPSS,


V20, Chicago, USA), were checked for normality using a
ShapiroWilks test, and analysed using one-way ANOVA to
assess differences in specic monomer content between different uncured materials. A Tukey post hoc test was used for
multiple comparisons. Pearson correlation and linear regression analyses were carried out to explore the relationship
between logP values and reverse phase liquid chromatography
retention time. All tests were carried out at = 0.05.

3.

Results

3.1.

LC/MS of reference monomers

Data of LC/MS analysis of reference monomers and detected


compounds in resin-composites and their structural formulas are summarized in Table 2 and Fig. 1. In full scan mass

714

Table 2 Characteristics of substances used or detected in LC/MS analysis.


Substance
abbreviation

Trivial name

MW
[M+Na]+

Bisphenol A
Glycidyl
Methacrylate

512

UDMA

Urethane
dimethacrylate
Triethylene
glycol dimethacrylate
Diethylene glycol
di-methacrylate
1,6-Hexanediol
di-methacrylate
2-Hydroxyethyl methacrylate
Bis-(acryloyloxymethyl)
tricyclo[5.2.1.0.sup.2,6] decane
Ethoxylated bis-phenol A
dimethacrylate
Ethoxylated Bis-phenol A
Diacrylate
SDR-urethane dimethacrylate
4-Dimethyl aminobenzoic acid
ethyl ester
Methyl methacrylate
1,3,7-Trimethylxanthine
(caffeine)

470

DEGDMA
HDDMA
HEMA
TCD-DI-HEA
BISEMA
EBPADA
SDR-UDMA
DMABEE
MMA
CF (internal standard)
a
b

Values in this row only appear in SDR.


values in this row only appear in FBF and FF.

242

535
476 [MC4 H5 O+Na]+
399 [MC8 H10 O2 +Na]+
519 [MOH+Na]+
493
425 [MC4 H5 O+Na]+
479 [MCH2 +Na]+ a
309
241 [MC4 H5 O+Na]+
353 [M+C2 H4 O+Na]+ b
265

254

277

130
478

131
501
447 [MC3 H3 O+Na]+
381 [MC6 H10 O+Na]+
536 (CH2 CH2 O)n
427 (CH2 CH2 O)n
481 (CH2 CH2 O)n
871

286

1700
468
848
193
100
194

[M+H]+

[M+NH4 ]+

Predicted
log P ()

48.2
13.0
6.3
11.4
34.6
10.2
2629
16.3
5.7
15.7
17.3

5.53 (0.48)

255

68.0

4.06 (0.3)
0.5 (0.28)
4.18 (0.42)

194

6.7
18.526.6
8.510.2
6, 69.4
2869
10.2
2429
48.6
26.2

195

10.1
6.1

530

471
474 [MCH2 +NH4 ]+
304

Reverse phase LC
average retention
time (min)

457[MCH2 +H]+

153

422 (CH2 CH2 O)n


476 (CH2 CH2 O)n

Not detected by MS

5.06 (0.46)

1.81 (0.43)

2.17 (0.35)

6.99 (0.4)
5.89 (0.35)
NA
3.14 (0.24)
1.35 (0.25)
0.13 (0.37)

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

BisGMA

TEGDMA

MS numbers (m/z)

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

715

Fig. 1 Structural formula of substances used or detected in the analysis.

spectra, the sodium adduct [M+Na]+ constituted the major and


highest intensity signal. Signals of both ammonium [M+NH4 ]+
and sodium adducts were detected for some compounds while
only the protonated adducts [M+H]+ were detected for low
molecular weight compounds.
Analysis of BisGMA showed multiple chromatographic
peaks with different mass to charge ratio (m/z) values on the
mass spectrum due to fragmentation of the molecule. The
sodium adduct of the parent BisGMA molecule was observed
at 535 m/z value at 4650 min. Sodium adducts of fragmented
BisGMA were observed at 476 [BisGMA CH2 C(CH3 )CO],
399 [BisGMA 2[CH2 C(CH3 )CO]], and 519 [BisGMA OH] m/z
values. The Sodium adduct of UDMA parent molecule was
observed at 493 m/z, and its protonated adduct at 471 m/z.
Sodium adduct of fragmented UDMA was observed at 425
m/z [UDMA CH2 C(CH3 )CO]. Sodium adduct of TEGDMA and
DEGDMA were observed at 309 and 256 m/z respectively with
fragmented TEGDMA molecule [TEGDMA CH2 C(CH3 )CO]
observed at 241 m/z. A low intensity signal of sodium adduct
of HEMA was observed at 153 m/z with a much lower intensity
signal of the protonated adduct at 131 m/z. MMA monomer
could not be detected by the MS.
BisEMA showed a unique mass spectrum and chromatogram compared to other monomers. The sodium adduct
of fragmented BisEMA was observed at 381 m/z [BisEMA
CH2 C(CH3 )CO,CH3 ,CH2 ] at 6 min and 69.4 min. Multiple chromatographic peaks with their corresponding mass spectra
were observed between 28 min and 69 min. The average m/z
value of the main signal was 536 (44)n (Fig. 2). The different
m/z values on the mass spectrum represent BisEMA molecules
having different number of ethoxy groups ( CH2 CH2 O )
(BisEMA homologues). The same was observed for EBPADA.
It showed 427 (44)n m/z value on the mass spectrum with a
high intensity HPLC peak at 10.2 min. Between 24 and 29 min,
multiple chromatographic peaks with m/z values of 481 (44)n
were shown representing EBPADA homologues with different
degree of ethoxylation. The highest molecular weights that
could be detected for BisEMA and EBPADA were 892 and 820
representing BisEMA and EBPADA homologues with twelve
and eleven ethoxy groups respectively. TEGDMA was detected

as an ingredient in all of BisGMA, BisEMA, and EBPADA reference solutions.

3.2.

LC/MS of uncured resin composites

LC/MS analysis of the methanol extract of different resincomposite materials showed variable combinations of
monomers with different relative amounts between materials (Table 3 and Fig. 3). In addition to the reference monomers
that were analysed, other monomers and compounds were
detected.
All materials showed TEGDMA on LC/MS except XF. TPH
showed a signicantly higher chromatogram peak area% of
TEGDMA (75.35%, SD 7.36) than other materials. TEC, FBF, VDF,
VBF, XB, and Z250 showed comparable TEGDMA area% which
was signicantly lower than other materials (0.196.26%).
Ethoxylated TEGDMA (TEGDMA+ CH2 CH2 O) with a molecular weight of 330 was detected in both FBF and FF. DEGDMA
was detected as a main component of XF with 67.97% chromatogram peak area. Only a small amount of DEGDMA was
detected in XB with a peak area of 2.82%.
BisGMA was detected in TEC, FF, GR, SF, Z250, FBF, GRF, and
TPH. The highest chromatogram peak area% of BisGMA was
shown by TEC (70.27%) followed by FF (62.59%) and GR and
(62.38%). TPH showed the lowest BisGMA chromatogram peak
area of 4.59%. UDMA was detected in VDF, VBF, FBF, SDR, Z250,
VD, TEC, and XB. VDF and VBF showed signicantly higher
UDMA area% than other materials (80.12% and 78.07% respectively). XB showed the lowest UDMA area% of 3.75%.
BisEMA was detected in all materials but was not quantiable for TEC. BisEMA chromatogram peak areas for GRF, FBF, FF,
GR, XF, Z250, VD, and TPH were not signicantly different and
ranged from 0.33 to 3.24%. SF, VBF, VDF, and SDR showed higher
BisEMA chromatogram peak area% which ranged from 11.60
to 16.91%. XB showed a signicantly higher chromatogram
peak area% of BisEMA (55.42%) compared to other materials.
EBPADA was detected in GRF, SF, and GR with a peak area
that ranged from 0.39% to 0.47%. EBPADA was also detected
in XB with a signicantly higher peak area (9.09%) than other
materials.

716

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

Fig. 2 LC/MS of BisEMA showing total ion and UV chromatograms (top). Note the multiple peaks in both chromatograms.
Some peaks are dened and their corresponding mass spectra are shown at the bottom. At 6.0 and 69.4 min, m/z value of
381 (sodium adduct of BisEMA CH2 = C(CH3 )CO,CH3 ,CH2 ) are observed on the mass spectrum. Note the caffeine (internal
standard) m/z value of 195 and corresponding UV chromatographic peak at 6 min. Total ion and UV chromatographic peaks
between 28 and 69 min shows mass spectra with an average m/z value of 536 (44)n . They represent BisEMA molecules
having different number of ethoxy groups ( CH2 CH2 O ). Peak at 16.6 min shows a mass spectrum with m/z value of 309
which represent the sodium adduct of TEGDMA (used as a diluent for BisEMA).

Some further compounds were detected in specic materials. HDDMA was detected in GRF at 68 min with a
chromatographic peak area of 43.35%. The sodium adduct of
the co-monomer HDDMA was observed at 277 m/z value on the
mass spectrum. In SDR, UDMA was observed at a different m/z
value than the reference UDMA monomer. Sodium, ammonium, and protonated adducts of [UDMA-CH2 ] were observed
at 479, 474, and 457 m/z values respectively at 2629 min. Also,
an m/z value of 871 was detected in SDR at 48.6 min with
a chromatographic peak area of 11.37%. This high molecular weight compound could represent the sodium adduct of
two modied UDMA molecules joined by an SDR photoactive group (polymerization modulator). In VD, m/z values of
501 (18.526.6 min) and 447 (8.510.2 min) were detected. They
represent the sodium adduct of TCD-DI-HEA molecule and
TCD-DI-HEA [COCH CH2 ] respectively and account for the
56.34% peak area on VD chromatogram. In XF, an unknown

compound was detected at 5.25.4 min with m/z value of 497


and peak area of about 6% on the chromatogram. It may represent the sodium adduct of a di- or multifunctional acrylate
or methacrylate resin with a molecular weight of 474.
The protonated adduct of DMABEE (co-initiator) was
observed at 194 m/z value at 26.2 min. It was detected in variable amounts in all materials except SDR, VBF, VDF, VD, and
TEC (0.511.89%). XB showed a statistically signicant higher
DMABEE peak area of 6.21% compared to other materials.

3.3.

log P and reverse phase LC retention time

The chromatographic retention time of non-fragmented


monomer molecules was in the following order from the
earliest to the latest: HEMA, MMA, DEGDMA, TEGDMA TCDDI-HEA, UDMA, BisGMA, SDR UDMA, and HDDMA. BisEMA
retention time took a range between UDMA and HDDMA while

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

717

Fig. 3 The main monomer content of the different resin-composites (bulk-ll and conventional) detected by LC/MS (mean
area% of the chromatogram). Monomers with values <1 chromatogram area % are not shown.

EBPADA took a range between TEGDMA and UDMA retention


times. HEMA showed the lowest predicted log P (0.5 0.28)
while BisGMA showed the highest value (5.53 0.48).log P
(octanolwater partition coefcient) could not be predicted
for SDR-UDMA due to its unknown structural formula. Similarly, a precise log P value could not be determined for both
BisEMA and EBPADA due to variable degree of ethoxylation in their structure, however, values of (6.99 0.4) and
(5.89 0.35) were obtained for BisEMA and EBPADA with two
ethoxy groups respectively (Table 2). Predicted log P values
of detected compounds showed a strong positive correlation
with chromatographic retention time (0.79 Pearson correlation
coefcient, p value = 0.004). Linear regression analysis gave r2
value of 0.62 (p = 0.004) (Fig. 4).

Fig. 4 A scatter plot showing a positive correlation and


linear regression between reverse phase HPLC retention
time of identied monomers and their predicted log P
(octanolwater partition coefcient). r2 = 0.62 (p = 0.004) at
= 0.05.

4.

Discussion

The results of the present study showed variable monomer


combinations and different relative amounts in the tested
bulk-ll and conventional resin-composites. Resin composition was material-specic with no particular monomers constituting the resin component of bulk-ll resin-composites as
a group of materials that differentiate them from conventional
resin-composites (with the exception of SDR). Thus, the rst
null hypothesis was accepted. Importantly, the composition
of uncured resin-composite materials that were investigated
in the current study using LC/MS (ESI) technique has not been
previously explored. The LC/MS detection method operated
with a reverse phase liquid chromatography separation procedure proved to be a valid technique in determining the relative
hydrophobicity of different monomers and compounds.
Retention times for identied compounds were positively correlated to their log P values (a measure of hydrophobicity);
therefore, the second null hypothesis was rejected.
GC/MS technique was used in previous studies, sometimes along with LC/MS, to identify resin composition and
eluates from several conventional resin-composites [2325].
Although conventional GC/MS has been shown to be a very
sensitive detection technique, its applicability in characterization of dental resins and leachable compounds is limited
for the detection of thermally stable, volatile, and low molecular weight compounds, namely additives, contaminants, and
degradation products [23]. Also, the electron impact ionization source that is usually used with this technique results in
extensive molecular fragmentation which makes it difcult
to identify compounds without matching them to a reference. Only recently, high temperature GC/MS could be used
to identify high molecular weight non fragmented BisEMA
homologues when direct on-column injection technique was

718

91.38
83.92
89.54
91.81
89.14
70.65
96.09
89.79
83.80
90.68
89.95

62.38
(4.24)
NDa
26.73d (1.74)
NDa
NDa
NDa
1.08a (0.52)
0.47a (0.01)
NDa
1.15b,c,d (0.31)
62.59 (10.23)
NDa
23.94d (1.70)
NDa
NDa
NDa
0.88a (0.39)
NDa
NDa
1.73d (0.43)
ND
NDa
NDa
67.97b (4.38)
NDa
NDa
2.17a (0.81)
NDa
NDa
0.51a,b (0.01)
ND
80.12e (2.75)
1.44a (0.17)
NDa
NDa
NDa
14.03b (0.33)
NDa
NDa
NDa
30.68 (4.01)
NDa
14.54c (1.48)
NDa
43.35b (4.67)
NDa
0.33a (0.15)
0.39a (0.16)
NDa
0.50a,b (0.03)
55.75 (4.27)
NDa
14.56c (1.40)
NDa
NDa
NDa
11.60b (3.63)
0.41a (0.18)
NDa
1.48c,d (0.14)
70.27 (0.67)
20.22b (3.73)
0.19a (0.08)
NDa
NDa
NDa
Not quantiable
NDa
NDa
NDa
44.43 (1.00)
43.53d (4.98)
0.37a (0.26)
NDa
NDa
NDa
0.39a (0.10)
NDa
NDa
1.23b,c,d (0.23)
ND
3.75a (0.93)
2.14a (0.13)
2.82a (1.23)
NDa
NDa
55.42c (3.15)
9.09b (1.26)
NDa
6.21e (0.33)

91.90
88.28
Total

ND
78.07e (1.18)
1.47a (0.06)
NDa
NDa
NDa
12.35b (1.43)
NDa
NDa
NDa

ND
35.61c (0.70)
24.39d (0.61)
0.00a
NDa
NDa
16.91b (4.73)
NDa
11.37b (4.68)
NDa
BisGMA
UDMA
TEGDMA
DEGDMA
HDDMA
TCD-DI-HEA
BISEMA
EBPADA
SDR-UDMA
DMABEE

79.43

Z250

52.06c,d (6.54)
28.12c (1.68)
6.26a,b (0.72)
NDa
NDa
0.00a
3.05a (0.14)
NDa
NDa
1.89d (0.75)
4.59 (0.74)
NDa
75.35e (7.36)
NDa
NDa
NDa
3.24a (0.39)
NDa
NDa
0.74a,b,c (0.15)
ND
24.07b,c (1.31)
9.13b,c (1.41)
NDa
NDa
56.34b (3.53)
3.06a (0.18)
NDa
NDa
NDa

TPH
VD

GR

c.d,e
d,e

FF
XF

a
a

VDF
GRF

b
c,d

SF
TEC

e
c

FBF
XB

VBF

SDR

Compound

Table 3 Main monomer content of the different resin-composites detected by LC/MS (mean area% of the chromatogram and standard deviation). Values with same
superscript letters per row indicate statistically homogenous groups (Tukey test, = 0.05).

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

used [14]. In a few studies, LC/MS equipped with a soft ionization source such as electro spray ionization (ESI) has been
reliably used for the detection of high molecular weight nonvolatile cross-linking monomers such as UDMA, Bis-GMA,
and BisEMA [24,2628]. In the current study, higher molecular weight compounds and new monomers (like SDR-UDMA
(848 MW), TDC-DI-HEA (478 MW), and homologous series of
BisEMA and EBPADA) were detected intact for the rst time
using LC/MS with ESI. Also, different m/z values of UDMA and
TEGDMA were detected in some materials like SDR, FF, and
FBF compared to other resin-composites which reect variations in the chemical structure of monomers used by different
manufacturers as has been previously reported [29].
In this study, LC/MS was operated using an internal
standard (caffeine) incorporated into all samples. The use of a
suitable internal standard in conjunction with LC/MS, LC, and
GC procedures provides a means for evaluating method performance, verication, and assessing the validity of the detection
technique.
Reverse phase liquid chromatography technique is the
best separation method for non-polar compounds that constitute the majority resin-composite monomers, and also has
the advantage of separating components according to their
order of hydrophobicity [30]. This has been shown in the
current study by the positive correlation between predicted
log P values of compounds with known structural formula and
their reverse phase liquid chromatography retention time. The
same relationship was shown in a previous study when HPLC
was used for identication of monomers composing different
dentin primers and bonding agents [31]. log P (octanolwater
partition coefcient) is a widely used partition coefcient used
to measure hydrophobicity in pharmaceutical and medicinal chemistry [32]. Hydrophobicity of constituent monomers
and compounds is considered as a key factor that determines the diffusibility of a storage solution into a polymeric
structure and its solubility [33]. In a previous study, log P
was found to be a good predictor of water sorption for polymerized experimental dental resins when they were grouped
according to compositional and structural similarities [34]. In
addition, from a biological point of view, hydrophobicity of
a substance has been found to correlate with its genotoxicity [35]. In a previous study using GC/MS, the octanolwater
partition coefcient for several detected leachable compounds
from resin-composite materials were laboratory measured as
a mean for assessing their lipophilicity. Values in that study
ranged from 0.5 to 4.1, however, values for BisGMA, UDMA, and
TEGDMA were not comparable to those obtained in the current study which is attributed to the use of different partition
solutions [25].
According to the determined relative hydrophobicity of
compounds detected in this study, higher solubility and elution into polar aqueous media is expected from XF compared
to other materials since its resin is mainly composed of the
hydrophilic monomer DEGDMA. The log P values of BisEMA
and EBPADA could not be obtained because of their variable
ethylene oxide chain length resulting in different molecular
weights and structural formulae for these compounds. However, predicted log P for BisEMA and EBPADA with two ethoxy
groups were the highest compared to other materials which
indicates high hydrophobicity of these monomers.

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

Bisphenol A (BPA) is a dental resin component usually


a contaminant from the manufacturing process of aromatic
monomers [19] that has been widely studied because of its
estrogenic parahormonal activity shown by in vivo studies
[36]. Although some studies have shown BPA to be extracted
from aromatic and BisGMA-based resin-composite materials (including Z250) [23,37], BPA was not detected at all as a
component of any of the currently studied resin-composite
materials. Possible reasons for undetected BPA in this study
include: the actual absence of BPA as a contaminant of the
resin component of the tested materials, the concentration of
BPA is below the detection threshold, or the detection method
is not suitable [19]. However, when cured, BPA might be present
as a by-product due to the breakdown of monomers.
Composition analysis of resin-composite components is
critical not only for assessment of their physical and mechanical performance but also for determining biological reactions
to these materials. All constituent monomers and compounds
detected in unpolymerized composites have been shown to
be extracted into organic solutions after curing [24]. Many
of these resin-composite monomers (such as TEGDMA) have
been known for a long time to act as sensitizers that elicit
various allergic reactions including contact dermatitis, lichen
planus, gingivitis, and ulcerations [38]. In addition, all of those
detected extracted monomers and compounds have shown
variable levels of cytotoxicity [39,40] and thus may contribute
to localized irritation of oral cells and tissues and may predispose to undesirable systemic effects.
Any research in dental materials is more meaningful when
it can provide an insight into the behavior of materials. This
forensic ability is critical in explaining why material A is
performing better than material B. In order to do so, an accurate knowledge of the composition of materials is required.
The results of the current study should provide a reference
for researchers concerned about the effect of resin composition on properties of a wide range of resin-composites
including the recently introduced bulk-ll materials. Among
these properties are: the degree of monomer conversion, polymer network cross-link density, degree of chemical softening,
sorption, solubility, and monomer elution into storage media
[15,41]. The provided data about types and relative amounts
of monomers (qualitative and quantitative) among tested
materials can be used for explaining differences in material
performance and properties which are mainly inuenced by
the resin matrix composition.

5.

Conclusions

Within the limitations of the current study the following can


be concluded:

1. Resin composition of bulk-ll resin-composites is generally comparable to that of conventional materials with the
exception of SDR.
2. The relative hydrophobicity of dental monomers can be
determined by their reverse phase liquid chromatography
retention time.

719

references

[1] Sadowsky SJ. An overview of treatment considerations for


esthetic restorations: a review of the literature. J Prosthet
Dent 2006;96:43342.
[2] Lynch CD, Opdam NJ, Hickel R, Brunton PA, Gurgan S,
Kakaboura A, et al. Guidance on posterior resin composites:
Academy of Operative Dentistry European section. J Dent
2014;42:37783.
[3] Van Landuyt K, Nawrot T, Geebelen B, De Munck J,
Snauwaert J, Yoshihara K, et al. How much do resin-based
dental materials release? A meta-analytical approach. Dent
Mater 2011;27:72347.
[4] Bowen RL, Dental lling material comprising vinyl silane
treated fused silica and a binder consisting of the reaction
product of bis phenol and glycidyl acrylate. 1962, Google
Patents.
[5] Ferracane JL. Resin composite state of the art. Dent Mater
2011;27:2938.
[6] Ogliari FA, Ely C, Zanchi CH, Fortes CB, Samuel SM, Demarco
FF, et al. Inuence of chain extender length of aromatic
dimethacrylates on polymer network development. Dent
Mater 2008;24:16571.
[7] Sideridou I, Tserki V, Papanastasiou G. Effect of chemical
structure on degree of conversion in light-cured
dimethacrylate-based dental resins. Biomaterials
2002;23:181929.
[8] Papakonstantinou AE, Eliades T, Cellesi F, Watts DC, Silikas
N. Evaluation of UDMAs potential as a substitute for
Bis-GMA in orthodontic adhesives. Dent Mater
2013;29:898905.
[9] Al-Ahdal K, Silikas N, Watts DC. Rheological properties of
resin composites according to variations in composition and
temperature. Dent Mater 2014;30:51724.
[10] Feilzer A, Dauvillier B. Effect of TEGDMA/BisGMA ratio on
stress development and viscoelastic properties of
experimental two-paste composites. J Dent Res
2003;82:8248.
[11] Kalachandra S, Turner DT. Water sorption of
polymethacrylate networks: bis-GMA/TEGDMA copolymers.
J Biomed Mater Res 1987:32938.
[12] Ferracane J. New polymer resins for dental restoratives. Oper
Dent 2001:199209.
[13] Hergenrother RW, Wabers HD, Cooper SL. Effect of hand
segment chemistry and strain on the stability of
polyurethanes: in vivo biostability. Biomaterials
1993;14:44958.
[14] Durner J, Schrickel K, Watts DC, Ilie N. Determination of
homologous distributions of bisEMA dimethacrylates in
bulk-ll resin composites by GCMS. Dent Mater
2015;31:47380.
[15] Ferracane JL. Hygroscopic and hydrolytic effects in dental
polymer networks. Dent Mater 2006;22:21122.
[16] Alrahlah A, Silikas N, Watts D. Post-cure depth of cure of
bulk ll dental resin-composites. Dent Mater 2014;30:14954.
[17] Alshali RZ, Salim NA, Satterthwaite JD, Silikas N.
Post-irradiation hardness development, chemical softening,
and thermal stability of bulk-ll and conventional
resin-composites. J Dent 2015;43:20918.
[18] Ilie N, Hickel R. Investigations on a methacrylate-based
owable composite based on the SDR (TM) technology. Dent
Mater 2011;27:34855.
[19] Noda M, Komatsu H, Sano H. HPLC analysis of dental resin
composites components. J Biomed Mater Res 1999;47:3748.
[20] Siuzdak G. The expanding role of mass spectrometry in
biotechnology. San Diego: Mcc Press; 2003.

720

d e n t a l m a t e r i a l s 3 1 ( 2 0 1 5 ) 711720

[21] rtengren U, Langer S, Gransson A, Lundgren T. Inuence


of pH and time on organic substance release from a model
dental composite: a uorescence spectrophotometry and
gas chromatography/mass spectrometry analysis. Eur J Oral
Sci 2004;112:5307.
[22] Schedle A, Ivanova M, Kenndler E. Determination of
ethoxylated bisphenol A dimethacrylate monomers in
dental composites by micellar electrokinetic
chromatography. J Chromatogr A 2003;990:2317.
[23] Michelsen VB, Lygre H, Sklevik R, Tveit AB, Solheim E.
Identication of organic eluates from four polymer-based
dental lling materials. Eur J Oral Sci 2003;111:26371.
[24] Spahl W, Budzikiewicz H, Geurtsen W. Determination of
leachable components from four commercial dental
composites by gas and liquid chromatography/mass
spectrometry. J Dent 1998;26:13745.
[25] Durner J, Spahl W, Zaspel J, Schweikl H, Hickel R, Reichl F-X.
Eluted substances from unpolymerized and polymerized
dental restorative materials and their Nernst partition
coefcient. Dent Mater 2010;26:919.
[26] Hsu WY, Wang VS, Lai CC, Tsai FJ. Simultaneous
determination of components released from dental
composite resins in human saliva by liquid
chromatography/multiple-stage ion trap mass spectrometry.
Electrophoresis 2012;33:71925.
[27] Tsitrou E, Kelogrigoris S, Koulaouzidou E,
Antoniades-Halvatjoglou M, Koliniotou-Koumpia E, van
Noort R. Effect of extraction media and storage time on the
elution of monomers from four contemporary resin
composite materials. Toxicol Int 2014;21:89.
[28] Mazzaoui S, Burrow M, Tyas M, Rooney F, Capon R.
Long-term quantication of the release of monomers from
dental resin composites and a resin-modied glass ionomer
cement. J Biomed Mater Res 2002;63:299305.
[29] Polydorou O, Knig A, Hellwig E, Kmmerer K. Uthethane
dimethacrylate: a molecule that may cause confusion in
dental research. J Biomed Mater Res B: Appl Biomater
2009;91:14.

[30] Valk K. Application of high-performance liquid


chromatography based measurements of lipophilicity to
model biological distribution. J Chromatogr A
2004;1037:299310.
[31] Silikas N, Watts D. High pressure liquid chromatography of
dentin primers and bonding agents. Dent Mater
2000;16:818.
[32] Leo A, Hansch C, Elkins D. Partition coefcients and their
uses. Chem Rev 1971;71:525616.
[33] George SC, Thomas S. Transport phenomena through
polymeric systems. Prog Polym Sci 2001;26:9851017.
[34] Dickens SH, Flaim GM, Floyd CJ. Effects of adhesive, base
and diluent monomers on water sorption and conversion of
experimental resins. Dent Mater 2010;26:67581.
[35] Dorn SB, Degen GH, Bolt HM, van der Louw J, van Acker FA,
den Dobbelsteen DJ, et al. Some molecular descriptors for
non-specic chromosomal genotoxicity based on
hydrophobic interactions. Arch Toxicol 2008;82:3338.
[36] Eramo S, Urbani G, Sfasciotti GL, Brugnoletti O, Boss M,
Polimeni A. Estrogenicity of bisphenol A released from
sealants and composites: a review of the literature. Ann
Stomatol (Roma) 2010;1:14.
[37] Pulgar R, Olea-Serrano MF, Novillo-Fertrell A, Rivas A, Pazos
P, Pedraza V, et al. Determination of bisphenol A and related
aromatic compounds released from bis-GMA-based
composites and sealants by high performance liquid
chromatography. Environ Health Perspect 2000;108:21.
[38] Alanko K, Susitaival P, Jolanki R, Kanerva L. Occupational
skin diseases among dental nurses. Contact Dermatitis
2004;50:7782.
[39] Spahl W, Budzikiewicz H. Qualitative analysis of dental resin
composites by gas and liquid chromatography/mass
spectrometry. Fresenius J Anal Chem 1994;350:68491.
[40] Issa Y, Watts D, Brunton P, Waters C, Duxbury A. Resin
composite monomers alter MTT and LDH activity of human
gingival broblasts in vitro. Dent Mater 2004;20:1220.
[41] Vasudeva G. Monomer systems for dental composites and
their future: a review. J Calif Dent Assoc 2009;37:389.

Das könnte Ihnen auch gefallen