Sei sulla pagina 1di 10

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

Available online at www.sciencedirect.com

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

A novel dentin bonding system containing poly(methacrylic


acid) grafted nanoclay: Synthesis, characterization and
properties
Laleh Solhi, Mohammad Atai , Azizollah Nodehi, Mohammad Imani
Iran Polymer and Petrochemical Institute (IPPI), P.O. Box 14965/115, Tehran, Iran

a r t i c l e

i n f o

a b s t r a c t

Article history:

Objectives. Developing a novel dentin bonding system containing poly(methacrylic acid)-

Received 7 April 2012

grafted-nanoclay (PMAA-g-nanoclay) as reinforcing ller, with high stability of nanoparticle

Received in revised form 4 June 2012

dispersion and improved bond strength and mechanical properties were the main objectives

Accepted 18 June 2012

of this study.
Materials and methods. Poly(methacrylic acid) (PMAA) was grafted onto the pristine sodium
montmorrillonite (Na-MMT) nanoclay surface and characterized using FTIR, TGA, and X-

Keywords:

ray diffraction (XRD). The PMAA-g-nanoclay was incorporated into an experimental dentin

Dentin bonding system

bonding system as ller in different concentrations and stability of nanoclay dispersion

Nanoclay

in the dilute adhesive, morphology of nanoclay layers in the photocured adhesive matrix,

Graft polymerization

shear bond strength to caries-free extracted human premolar teeth, and mode of failure

Poly(methacrylic acid)

were studied. The mechanical properties including diametral tensile strength (DTS), exural
strength (FS), and exural modulus (FM) were also investigated. The measured FM was also
compared to theoretical prediction models.
Results. The grafting of PMAA onto the nanoclay surface was conrmed and the results
revealed a partially exfoliated structure for PMAA-g-nanoclay. The dispersion stability of
the modied nanoparticles in the dilute adhesive increased more than 45 times in comparison with the pristine nanoclay. The incorporation of 0.5 wt.% PMAA-g-nanoclay to the
adhesive resulted in a signicant increase in microshear bond strength, DTS, and FS. Higher
PMAA-g-nanoclay contents resulted in increased exural modulus. The experimental exural modulus was in good agreement with the HalpinTsai theoretical model.
Signicance. Incorporation of PMAA-g-nanoclay particles as novel functional llers into dental adhesive could result in the development of bonding systems with improved physical,
mechanical, and adhesion properties.
2012 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

The main function of dental adhesives is to bond dental


restorative materials to the tooth structure [1]. Theadhesion

is necessary to avoid secondary carries, sensitivity after


restoration, color change, and microleakage [2]. Dentin is a
hydrated biological composite which consists of water, organic
matrix and mineral materials with dentinal tubules located
in the radius of dentin in which a uid ows from inside to

Corresponding author. Tel.: +98 21 48662446; fax: +98 21 44580023.


E-mail addresses: m.atai@ippi.ac.ir, m.atai@yahoo.com (M. Atai).
0109-5641/$ see front matter 2012 Academy of Dental Materials. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.dental.2012.06.004

1042

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

outside direction [35]. Due to the wet and dynamic structure, adhesion to dentin is more complicated than enamel [6].
Penetration of bonding monomers into the dentin structures,
which subsequently results in a hybrid layer of polymerized resin and collagen brils and strengthening of adhesion
to dentin, is increased by using water chasing solvents like
ethanol [7]. Formation of the hybrid layer along with the
resin tags provide a micromechanical retention between resin
and dentin [8,9]. The elastic modulus of the adhesive layer
formed at the resindentin interface is the lowest comparing
to dentin and restorative materials [2,10]. The low elastic modulus causes the failure of the restoration when the occlusal
loading is more than the strength of the layer [7,11]. It has been
shown that the incorporation of small amounts of nanollers
into polymeric matrices improves mechanical and fracture
properties of the composites [12,13]. Na-MMT is a nanoclay
which occurs in the nature as platelets which are 1 nm thick
and up to 1 m in diameter. Because of the high aspect ratio,
nanoclay is efciently used to improve mechanical properties
of polymeric systems [14]. Incorporation of clay nanoparticles into the dilute systems such as dentin bondings, however,
results in the fast sedimentation of the particles due to their
higher densities. Surface modication of nanoclay platelets
with a polymer, has been shown, to solve the problem due to
the decrease in the density of the particles and providing interactions between the newly attached functional groups on the
surface of particles with the solvent. Grafting is an efcient
method to introduce functional groups onto the surface of
nanoparticles [12,15,16]. In our previous works, Na-MMT pristine nanoclay surface, was modied via graft polymerization of
methyl methacrylate and acrylic acid to be used as reinforcing
llers for dental adhesives [15,17].
In this study, the Na-MMT pristine nanoclay surface was
modied with methacrylic acid monomers, which consist
of carboxylic groups. The carboxylic acid functional groups
could interact with Ca2+ ions of hydroxyapatite [18] promoting the adhesion between the restorative materials and dentin
[12,15,18]. Having grafted PMAA chains onto the surface of
nanoclay, the grafted nanoclay platelets were characterized
and their dispersion stability in an experimental dental adhesive was investigated. Mechanical properties and microshear
bond strength of the dentin bonding system containing PMAAg-nanoclay were investigated. The elastic properties of the
adhesives were also modeled using theoretical prediction
models.

2.

Experimental

2.1.

Materials

(AMPS), ammonium persulfate, and lithium chloride (LiCl)


were purchased from SigmaAldrich (Germany). AdperTM Single Bond 2, a commercially available nanoparticle containing
dentin bonding, was obtained from 3M ESPE (USA). The 37.5%
phosphoric acid gel (Kerr Gel Etchant) was obtained from
SDS Kerr (USA). Deionized water was used throughout all the
experiments.

2.2.

2.2.1. Graft polymerization of methacrylic acid onto the


surface of nanoclay platelets
The graft polymerization of methacrylic acid onto the surface of nanoclay was carried out according to the previous
works [15,17]. Briey, MAA was grafted onto the clay platelets
in an aqueous solution in the presence of AMPS as reactive
surfactant and TDM as a chain transfer agent to control the
molecular weight of the grafted polymer.

2.2.2.

Characterization of the PMAA grafted nanoclay

The pristine Na-MMT and PMAA-grafted nanoclay were analyzed by FTIR spectroscopy (EQUINOX 55, Bruker, Germany)
at a resolution of 4 cm1 and 32 scans in the range of
4000400 cm1 using KBr disc technique. Thermogravimetric analyses of the pristine Na-MMT, PMAA-g-nanoclay and
pure PMAA were performed (TGA-1500, Polymer Laboratories, UK) from room temperature to 600 C at a heating rate
of 10 C min1 and under N2 atmosphere. X-ray diffraction
patterns of pristine Na-MMT and PMAA-g-nanoclay were collected in the range of 2 = 210 and step size of 0.02 , using
a Philips X-ray diffractometer (Philips, Xpert, Netherlands)
operating at a voltage of
with copper target,  = 1.54056 A
40 kV and a current of 40 mA at the rate of 2 min1 . The
d0 0 1 spacings were calculated according to the Braggs equation: n = 2d sin . The PMAA-g-nanoclay was then stirred at
50 C for 5 days in a 10 g L1 LiCl solution to debond the PMAA
chains from the clay surface [19]. The debonded nanoclay was
ltered and dried, and the polymer in the solution was precipitated in acetone, washed and dried. Thermogravimetric
analyses of the debonded polymer and nanoclay were performed. The molecular weight of the debonded polymer and
the molecular weight of the unbonded homopolymer which
was extracted during Soxhelet-extraction were determined by
GPC (GPC, Agilent 1100 series, USA) using Aquagel column of
7.5 mm 300 mm (ID L) with the ow rate of 1 mL min1 at
23 C.

2.2.3.

2-Hydroxyethyl methacrylate (HEMA), 2,3-butanedione (BD),


trimethacrylate
2-ethyl-2-hydroxymethyl-1,3-propandiol
(TMPTMA), methacrylic acid (MAA), acetone, ethanol and
methanol was purchased from Merck (Germany). tert-Dodecyl
mercaptan (TDM) were obtained from Fluka (Germany). 2,2Bis[4-(2-hydroxy-3-methacryloxypropoxy)phenyl]
propane
(Bis-GMA) was kindly donated by Evonik (Germany). Na-MMT
(Cloisite Na+ ) was obtained from Southern Clays Product,
Inc. (USA). 2-Acrylamido-2-methyl-1-propane-sulfonic acid

Methods

Preparation of adhesives

The adhesive was prepared according to the formulation


shown in Table 1. Pristine NA-MMT and PMAA-g-nanoclay
were added to the adhesive in 0.2, 0.5, 1, 2 and 5 wt.%.
The llers were well dispersed in the adhesive solution by
ultra sonication using a probe sonication apparatus (Sonoplus UW2200, Bandelin, Germany) for 3 min in an ice bath.
The stability of the dispersions was investigated by a separation analyzer (LUMiReader 416.1, LUM, Germany) working
with visible light and intensities of 25% and tilt of 0, for 20 h
(PMAA-g-nanoclay) and 30 min (pristine nanoclay), including
256 intervals.

1043

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

Table 1 Adhesive composition.

Bis[4-(2-hydroxy-3methacryloxypropoxy)phenyl]
propane (Bis-GMA)
2-Hydroxyethyl methacrylate
(HEMA)
2-Ethyl-2-hydroxymethyl-1,3propandiol trimethacrylate
(TMPTMA)
Urethane dimethacrylate (UDMA)
Ethanol
2,3-Butanedione (BD)
a

pristine nanoclay

%Weight
14

26

PMAA-g-nanoclay

Carbonyl group
of PMAA

4000

12
39
1a

The photoinitiator was added after sonication to prevent unwanted polymerization before applying the adhesive.

2.2.4.

Transmittance

Materials

Measurement of degree of conversion

A droplet of the adhesives containing 1% BD as photoinitiator was placed on a polyethylene lm. The solvent of
the adhesive was gently evaporated for 10 s applying a lowpressure air stream and a second lm was then placed on
it to form a very thin layer. The sandwich was placed into
the FTIR spectrometers sample holder and the absorbance
peaks obtained by transmission mode of FTIR before and after
40 s photocuring using a dental light source with an irradiance of ca. 600 mW cm2 (Optilux 501, SDS Kerr, Germany).
The degree of conversion (DC%) was determined from the ratio
of absorbance intensities of aliphatic C C (peak at 1638 cm1 )
against internal reference of aromatic C C (peak at 1608 cm1 )
before and after curing of the specimen [17].

2.2.5. Transmission and scanning electron microscopies


(TEM and SEM)
Adhesive specimen containing 1 wt.% PMAA-g-nanoclay was
solvent evaporated and light cured in a mold. An ultramicrotome (DMU3, Reichert, Austria) was used to prepare
approximately 70 nm thick TEM specimen. A Philips TEM
(CM200, FEG, Netherlands) was used for TEM observations. The
compositedentin interface was observed using SEM (TESCAN,
VEGAII, XMU, Czech Republic) to study the mode of failure in
the microshear bond strength test. EDX elemental composition analyzer was used to map the distribution of nanoclay
in the photocured dentin bonding system. A probe current of
2.0 109 A, an accelerating voltage of 30 kV, and spot size of
500 nm with a collection time of 100 s, were used during the
mapping. The elemental compositions of the products were
quantitatively identied by energy dispersive X-ray analysis
(EDXA, Model: QX2, RONTEC Co.) which was coupled with the
SEM.

2.2.6. Mechanical tests


2.2.6.1. Diametral tensile strength. Diametral tensile strength
(DTS) test was performed adopting the procedure of ANSI/ADA
specication no. 27 for light cured resins [20] using a universal
testing machine (STM-20, Santam, Iran) at a cross-head speed
of 10 mm min1 with a 5 KN load cell.

2.2.6.2. Flexural strength. Flexural strength of the solvent-free


adhesives was conducted according to the 3-point bending
method suggested in ISO 4049 [21] using the same universal

3500

3000

2500

2000

1500

1000

500

-1

Wavenumber (cm )

Fig. 1 FTIR spectrums of pristine Na-MMT nanoclay and


PMAA-g-nanoclay.

testing machine at a cross-head speed of 1 mm min1 with a


200 N load cell.
A number of theories and models have been developed to
describe the elastic properties of polymeric composites based
on the volume fraction, shape, size, aspect ratio, and distribution of the reinforcing particles, the stiffness of the matrix
and ller phase, and llermatrix interactions. Among them
HalpinTsai equation has successfully been used to predict
the elastic properties of nano-particle containing polymeric
composites [22,23]. The experimental data were tted to the
model equations and compared.

2.2.6.3. Microshear bond test. The microshear bond test was


performed according to previous works [12,15]. A commercially available dentin bonding (AdperTM SingleBond 2, 3M
ESPE, USA), as control group, was also applied following
the manufacturers instruction and resin composite (Premise,
Kerr, USA) was used as the lling material [24].

2.2.7.

Statistical analyses

The results were analyzed and compared using one-way


ANOVA and Tukey test at the signicance level of 0.05. The
reported values are the average of 10 measurements for
microshear bond strength, 5 repeats for DTS, FS, and FM and
3 measurements for degree of conversion.

3.

Results

FTIR spectra of the pristine Na-MMT and PMAA-g-nanoclay


are shown in Fig. 1. Determined by GPC, the Soxheletextracted PMAA homopolymer shows a weight average
molecular weight (Mw ) of 7500 (PDI = 1.8), while the chromatogram of the reverse ion exchange debonded PMAA,
reveals a Mw of about 2300 (PDI = 1.4). The TGA curves
of the pristine Na-MMT, PMAA-g-nanoclay, nanoclay after
ion exchange process and neat PMAA are represented in
Fig. 2. The PMAA was synthesized in the same conditions
as the graft polymerization. Considering the ash content
of neat PMAA in the thermogravimetric analysis (5 wt.%
at 550 C), the grafting PMAA percentage was calculated
about 43 wt.%. Fig. 3 illustrates the XRD patterns of pristine
Na-MMT, PMAA-g-nanoclay and cured adhesive containing
1 wt.% PMAA-g-nanoclay. Comparison of the XRD patterns

1044

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

Weight loss (%)

100

nanoclay
nanoclay after ion exchange

80

60
PMAA-g-nanoclay

40

20
PMAA

0
0

100

200

300

400

500

600

Temperature (C)

Fig. 2 TGA thermograms of PMAA, PMAA-g-nanoclay,


pristine Na-MMT nanoclay and nanoclay after debonding
the grafted PMAA.

Intensity (arbitrary units)

Fig. 4 TEM micrograph of the adhesive containing 1 wt.%


PMAA-g-nanoclay showing partially delaminated clay
platelets.

PMAA-g-nanoclay

nanoclay

PMAA-g-nanolay containing adhesive

1.5

2.5

3.5

4.5

5.5

6.5

7.5

8.5

9.5

2Theta ()

Fig. 3 XRD patterns of pristine Na-MMT,


PMAA-g-nanoclay and adhesive containing 1 wt.%
PMAA-g-nanoclay.

shows that the peak position from 2 = 9.41 , correspond for pristine
ing to an interlayer spacing of d0 0 1 = 9.39 A
Na-MMT, is shifted to 2 = 6.15 , representing an interlayer
for PMAA-g-nanoclay; this peak is
spacing of d0 0 1 = 14.36 A
disappeared for the adhesive containing 1 wt.% PMAA-gnanoclay. TEM image of light cured adhesive containing 1 wt.%
PMAA-g-nanoclay is shown in Fig. 4, which indicates intercalation/exfoliation of PMAA-g-nanoclay particles in the adhesive
system. Fig. 5 is the separation analysis results which shows
the sedimentation behavior of the nanoparticles in the adhesives for the specimens containing 1 wt.% of pristine Na-MMT
and treated nanoclay. The transmission of the adhesive containing PMAA-g-nanoclay did not reach 100% during 20 h. The
transmission of 80% (at the middle point of testing tube) was
obtained after ca. 7 min for the pristine nanoclay containing
adhesive whereas the same transmission was reached after
ca. 315 min for the adhesive containing PMAA-g-nanoclay,
which shows about 45 times increase in the stability of the
modied particles in the adhesive solution. To investigate
the dispersion of modied particles in the adhesives, the
cross section of cured adhesives was examined by EDXA. The
Si map of the cured adhesive containing 0.5 wt.% PMAA-gnanoclay was collected (Fig. 6). Figs. 7 and 8 show DTS and

FS of the adhesive samples containing different concentrations of PMAA-g-nanoclay. DTS and FS show a maximum
corresponding to the adhesive containing 0.5 wt.% PMAA-gnanoclay signicantly higher than those of the unlled resin
(p = 0.000). Fig. 9 presents the FM of the adhesives containing
PMAA-g-nanoclay. An increasing trend is observed in the FM
of the adhesive composites upon the increase in ller content. The model predictions for the FM of adhesives containing
different amounts of PMAA-g-clay are also calculated and presented in Fig. 9. DC% values of the adhesives were in the range
of 7885% and there were no signicant difference between
the DC% of the adhesives with different ller contents and
SingleBond (p > 0.05).
Fig. 10 shows the microshear bond strength of the adhesives containing different percentages of PMAA-g-nanoclay.
It reveals a signicant increase (p = 0.025) in the microshear
bond strength of the adhesive containing 0.5 wt.% PMAA-gnanoclay. In Fig. 11 the penetration of the adhesive into the
dentinal tubules is clearly observed in low ller concentrations (0.5 wt.%) while most of the tubules are left unlled in
high concentrations (5 wt.%). Agglomeration of the particles
in higher concentrations of the llers results in the decrease
of the adhesive penetration into the dentin tubules. The
failure mode in the shear bond test was mostly adhesive
from the adhesivedentin interface which is representatively
illustrated in Fig. 12.

4.

Discussion

In the graft polymerization, the pristine Na-MMT is dispersed


in water in the presence of a reactive surfactant, AMPS, which
contains amido and sulfonic acid groups in its structure. It
has been suggested that the interaction of AMPS with clay
occurs by proton transfer from sulfonic acid to nitrogen, to
form a protonated amido group and ion exchange with sodium
ions on the silicate layers [25]. AMPS can also be adsorbed
onto the surface of the clay galleries by formation of hydrogen bonds between its amido groups and water molecules

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

1045

Fig. 5 Separation analysis in LUMi Reader . Sedimentation behavior of the adhesive containing: (a) 1 wt.% pristine
Na-MMT sonicated for 3 min (total test duration: 30 min) and (b) 1 wt.% PMAA-g nanoclay sonicated for 3 min (total test
duration: 20 h).

surrounding the exchangeable cations, and iondipole interactions between its sulfonate groups and the interlayer
exchangeable cations [26]. These interactions lead to an
increase in the basal spacing of pristine Na-MMT. The adsorption of AMPS onto clay by formation of hydrogen bonding
between its amido group and the hydroxyl groups on the
edges of the clay platelets, however, has no effect on the
basal spacing. Polymerization of methacrylic acid monomers
in the presence of nanoclay platelets, AMPS, and a water
soluble free radical initiator, ammonium persulfate, leads
to both graft polymerization onto clay surface involving
the vinyl groups of AMPS, and the homopolymerization of
MAA in aqueous phase. Therefore, the reaction product is a

mixture of PMAA-g-nanoclay and PMAA homopolymers. Having Soxhelet-extracted with water for 72 h to remove PMAA
homopolymers completely, the remained PMAA-g-nanoclay,
was characterized using different analytical techniques. The
grafting of PMAA onto the pristine Na-MMT was conrmed
by FTIR spectroscopy (Fig. 1). The appearance of two PMAA
characteristic peaks at 1736 cm1 and 2955 cm1 which are
assigned to the stretching vibrations of carbonyl and C H
groups in poly(methacrylic acid) structure, respectively, [27]
is the indication of the presence of grafted PMAA.
TDM, which is a chain transfer agent, results in the
formation of low molecular weight grafted PMAA and
the homopolymer chains. Steric hindrance adjacent to the

1046

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

Fig. 8 Flexural strength (FS) of the adhesives containing


different PMAA-g-nanoclay content. Y-Error bars represent
the standard deviations.

Fig. 6 Si map of the cured adhesive containing 0.5%


PMAA-g-nanoclay.

Flexural modulus (GPa)

4.5
4
3.5
3
2.5
2
1.5
1
0.5
0
0.00

0.01

0.02

0.03

0.04

0.05

0.06

PMAA-g-nanoclay volume fraction


Experimental

Law of mixtures

Transverse law of mixtures

Halpin-Tsai

Fig. 9 Flexural modulus (FM) of the adhesives containing


different PMAA-g-nanoclay content. Y-Error bars represent
the standard deviations.

Fig. 7 Diametral tensile strength (DTS) of the adhesives


containing different PMAA-g-nanoclay contents. Y-Error
bars represent the standard deviations.

surface of nanoclay platelets is probably responsible for the


low molecular weight of the grafted chains in comparison to
the PMAA homopolymers. According to the thermogravimetric results in Fig. 2, the percentage of grafted PMAA is about
43 wt.%. As the PMAA-g-nanoclay was Soxhelet-extracted for
72 h with water, which is a good solvent for PMAA, the weight
loss in the grafted clay is due to the thermal degradation of
bonded PMAA polymer chains. There is no sharp weight loss
for degradation of PMAA in the grafted clay, which is probably due to the thermal stabilization effect of nanoclay on the
polymer [28]. TGA curve for neat PMAA, indicates two stages
of breakdown starting at ca. 240 C and 460 C, respectively.

Fig. 10 Microshear bond strength of the adhesives


containing different percentages of PMAA-g-nanoclay.
SingleBond (3M ESPE, USA) is a commercially available
dentin bonding agent. Y-Error bars represent the standard
deviations.

1047

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

An initial weight loss below 100 C is due to the release of


absorbed water. The rst stage of break down (up to ca. 300 C)
accounts for about 25% weight loss. According to the degradation mechanism, anhydropoly(methacrylic acid) is formed at
the rst stage. The dehydration reaction occurs by intramolecular cyclisation of adjacent monomer units to lose water and
give six-membered anhydride ring structures. At higher temperatures, the second stage, the polymer undergoes massive
degradation evolving both water and carbon dioxide and the
anhydride structures are decomposing [29]. In Fig. 3 the characteristic XRD peak of nanoclay in the PMAA-g-nanoclay has
been broadened and also shifted toward lower angles, which is
an indication of the lower regularity and larger basal spacing
as a result of the intercalation of PMAA chains into the clay
galleries. The grafted PMAA chains prevent the clay sheets to
stack up after removing from the aqueous reaction medium.
The solvating power of ethanol of the adhesive causes the
grafted PMAA chains to swell and allow the diffusion of multifunctional monomers of the adhesive into the clay layers.
Removing the solvent and curing the adhesive leaves additional organic components between the layers and results in
increased basal spacing and disappearance of XRD peak in the
cured adhesive containing 1 wt.% PMAA-g-nanoclay conrming an exfoliated structure which is further supported by the
TEM observations.
Dispersion stability analysis of modied nanoclay containing adhesives showed a signicant increase in the
sedimentation time of the particles through the dilute
bonding system (Fig. 5). Ethanol is a good solvent for
PMAA (solubility parameters of ethanol and PMAA are 11.7
and 12.48 cal1/2 cm3/2 , respectively) [30], therefore polymer
chains, grafted onto the nanoclay surface, swell in the adhesive and force the nanoclay platelets to separate from each
other which results in decreasing the overall density of the
ller particles and increasing the sedimentation time. A uniform distribution of the particles in the adhesive resin matrix
is also observed in the Si-mapping of the cured adhesive
(Fig. 6).
Mechanical properties of the resin-based bonding systems
are strongly depend upon the degree of conversion (DC%).
Therefore, they should preferably be compared in the same
degree of conversion. DC% measurements showed that there
is no signicant difference between the DC% of the test specimens providing a reliable comparison and analysis (p > 0.05).
DTS test is valid for the brittle materials in which plastic deformation is negligible [31]. The forcedisplacement
curves in this test revealed a brittle behavior validating the
DTS results. The increased DTS and FS of adhesive containing 0.5% PMAA-g-nanoclay comparing to the unlled resin
(Figs. 7 and 8) are probably due to high strength of llers,
crack lengthening mechanism and good interaction of polymer grafted ller platelets with the polymeric matrix. In
specimens with higher ller content, however, some weak
points are formed due to agglomeration of PMAA-g-nanoclay
which can result in crack initiation and reduced DTS and FS,
subsequently.
The elastic modulus of a composite depends on the elastic properties of its components. Since the inorganic llers
generally have higher stiffness, the modulus of the polymeric
matrix of dental adhesives is improved by incorporation of

the rigid llers. The higher elastic modulus observed in the


lled adhesives (Fig. 9) is due to the higher elastic modulus of
the nanoclay particles in comparison to the polymeric resin
matrix.
Theoretical or semi-empirical models have been developed
to describe the effect of reinforcing phase on the elastic properties of the composites. In this study, different models have
been applied to describe the elastic modulus of the nanoparticle lled dental adhesives.
To apply the models on the experimental data, the following parameters were measured and/or calculated:
The aspect ratio of ultrasound dispersed montmorillonite
Cloisite Na+ , considered as 200 [32], montmorillonite density
as 2.86 g cm3 , montmorillonite modulus as 170 GPa (Eclay ) [33],
density of PMAA-g-nanoclay as 1.82 g cm3 (experimentally
measured), unlled resin density as 1.09 g cm3 (calculated
by rule of mixtures taking into account the density and volume fraction of the monomers and solvent in the adhesive),
unlled resin modulus as 1.11 GPa (Fig. 9), PMAA density as
1.23 g cm3 (experimentally measured), and PMAA modulus
as 0.79 GPa (EPMAA ) [34]. The modulus of the modied nanoclay particles was approximated as 62 GPa using the law of
mixtures:
EPMAA-g-clay = Eclay Vclay + EPMAA VPMAA

(1)

The volume fraction of the grafted PMAA and the nanoclay


in the modied particles were calculated according the following equation:
Vclay =

Wclay
Wclay + clay /PMAA (1 Wclay )

(2)

where Vclay , Wclay , and clay are the volume fraction, weight
fraction, and density of the nanoclay and PMAA is the density
of PMAA.
To calculate the aspect ratio of PMAA-g-clay particles, it is
assumed that the grafted PMAA forms two layers on both sides
of the clay platelets. So, the aspect ratio of grafted particles is
calculated as 73 considering the volume percent of the grafted
PMAA onto the surface of nanoclay (64 vol.% of each particle),
density of PMAA, and density of PMAA-g-nanoclay.
The law of mixtures equations are the rst simple models
which are used to describe the phenomena of composite materials. The upper and lower bounds for elastic modulus can be
estimated using the equations considering that the ller and
matrix phases in the adhesive undergo an average stress or
strain. Assuming uniform strain in the individual phases, the
upper bound is given by:
Ec = Ef Vf + Em (1 Vf )

(3)

When the stress is assumed to be uniform in both phased,


the lower bound is obtained from:
Ec =

Ef Em
Ef (1 Vf ) + Em Vf

(4)

where Ec , Em , and Ef are the modulus of the adhesive


nanocomposite, adhesive matrix, and the ller(nanoparticles),

1048

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

Fig. 11 SEM micrographs of the fracture area in microshear bond strength test of the adhesives containing 5 wt.% (a) and
5 wt.% PMAA-g nanoclay (b).

limits. In the law of mixtures, elastic modulus of the components and the volume fraction of the inclusions are the only
parameters which have been taken into account while the
elastic properties of composite systems depend not only on
the volume fraction and modulus of the reinforcing phase, but
also upon its size, shape, distribution, and the interfacial adhesion. In the more developed models, the other parameters are
also considered. HalpinTsai model is one of them which considers the llers shape and has been used in the prediction of
elastic properties of clay-containing nanocomposites [22,23].
The model comes from work by Halpin and Pagano [36], and
gives the modulus of the material as a function of the modulus
of matrix polymer, Em , and modulus of ller particles, Ef . The
model also considers the ller aspect ratio by the inclusion of
a shape factor. The predicted composite modulus is calculated
by Eq. (5).
Ec =

1 + Vf
1 Vf

Em

(5)

where  is the shape factor, Vf is the volume fraction of particles and:


Fig. 12 Typical SEM micrograph of the fracture area in
microshear bond strength test showing an adhesive failure
from adhesivedentin interface.

respectively. Vf stands for the volume fraction of the nanoparticles.


The equations are unable to predict the experimental values in this study (Fig. 9), implying that the assumption of
either a state of uniform strain or uniform stress is not sufcient to describe the modulus of the lled systems [35]. The
equations, however, may approximate the upper and lower

=

(Ef /Em ) 1
(Ef /Em ) + 


(6)

The value of  correlated with the geometry of the reinforcing phase, especially with the aspect ratio (w/t) of the
particles, where w is the length of the particle and t is its
thickness. By comparison of the predictions with the results
of a nite-element analysis, Halpin and Kardos [37] suggested
a shape factor of  = 2(w/t), here  = 2 73 = 146, for calculating the modulus of a polymer with the particles aligned with
the loading direction [13]. Although the nanoclay particles are
randomly distributed in the adhesive matrix, the model prediction is in good agreement with the experimental results.

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

The theory was chosen in this work, among the other similar models, because of its effectiveness in predicting stiffness
of nanoclay reinforced composites, its adaptability for different ller geometries, particularly disks, and for its prevalence
in the literature [38,39]. The deviation of the experimental
result of 5 wt.% ller from the prediction is probably due to the
agglomeration of the nanoclay platelets which dramatically
reduces the dispersion of the particles in the matrix and consequently the effectiveness of the reinforcing nanoparticles in
enhancing the elastic modulus. Overestimation of HalpinTsai
method in prediction of the elastic modulus of a nanoclay
composite for particle volume fractions of more than 3%, has
been reported [23].
In the models a number of assumptions are inherent;
the ller and matrix are linearly elastic, isotropic, and
rmly bonded. Particleparticle interactions, however, are not
explicitly considered. The HalpinTsai equation for Ec is independent of the Poissons ratio of the ller or the matrix.
The ultimate goal of a dental bonding system is providing a strong and durable bond between restoration and tooth.
The microshear test (Fig. 10) shows a signicant increase
in the bond strength of the adhesive containing 0.5 wt.%
PMAA-g-nanoclay (p = 0.025). It has already been reported that
the incorporation of an optimum amount of ller to adhesives increases the mechanical properties and bond strength
[7,15,40]. Incorporation of llers into adhesive causes an
increase in its elastic modulus providing a layer with an
elastic modulus between dentin and restoration. This intermediate layer, together with the resinimpregnated dentin,
act as an elastic buffer which offers the resindentin interface
a sufcient strain capacity to accommodate both the composite and dentin [41]. Fig. 10 shows that the bond strength
increases to optimum maximum level (0.5 wt.% of PMAA-gnanoclay) and then gradually decreases with the increase of
ller content. The decrease in the shear bond strength is due
to the agglomeration of the nano-particles, and accumulation
of ller agglomerates on the top of the etched dentin substrate. The higher viscosity of the adhesive at higher ller
contents could also reduce the penetration of adhesive and
consequently decrease the micromechanical retention and
microshear bond strength. Fig. 11a shows that the adhesive
containing 0.5 wt.% is able to penetrate into the dentin tubules
providing micromechanical interlocking through the formation of reinforced resin tags while in the case of adhesive with
high ller content (5 wt.%) the penetration is not complete
leaving most of the tubules empty (Fig. 11b). The lack of resin
tags and/or insufcient penetration of the resin into the intertubular dentin matrix result in a poor retention and drop in the
microshear strength. SEM of the shear bond test area revealed
that the failure mode at different percents of ller content is
mostly adhesive (Fig. 12 is a typical micrograph).

5.

Conclusion

A pristine Na-MMT nanoclay was modied by graft polymerization of PMAA onto the surface of the clay platelets. The
grafted nanoclay was characterized using FTIR, TGA, GPC and
XRD techniques. The modied nanoclay was dispersed in an
experimental dental adhesive and the dispersion stability was

1049

shown to be improved 45 times with respect to the pristine


nanoclay. XRD and TEM studies on the adhesive containing
modied nanoclay showed a partially exfoliated morphology
for the PMAA-g-nanoclay in the adhesive matrix. Incorporation of 0.5 wt.% PMAA-g-nanoclay to the experimental dental
adhesive resulted in a signicant increase in microshear bond
strength, DTS and FS comparing to other concentrations of
ller. HalpinTsai predictions for Youngs modulus of the
adhesives with different contents showed a good agreement
with experimental results.

references

[1] Krifka S, Brzsnyi A, Koch A, Hiller KA, Schmalz G, Friedl


KH. Bond strength of adhesive systems to dentin and
enamel human vs. bovine primary teeth in vitro. Dental
Materials 2008;24(7):88894.
[2] Nakabayashi N, Kojima K, Masuhara E. The promotion of
adhesion by the inltration of monomers into tooth
substrates. Journal of Biomedical Materials Research
1982;16(3):26573.
[3] Marshall GW, Marshall SJ, Kinney JH, Balooch M. The dentin
substrate: structure and properties related to bonding.
Journal of Dentistry 1997;25(6):44158.
[4] Mannocci F, Pilecki P, Bertelli E, Watson TF. Density of
dentinal tubules affects the tensile strength of root dentin.
Dental Materials 2004;20(3):2936.
[5] Wang R, Weiner S. Strainstructure relations in human teeth
using Moire fringes. Journal of Biomechanics
1997;31(2):13541.
[6] Poitevin A, De Munck J, Van Landuyt K, Coutinho E, Peumans
M, Lambrechts P, et al. Critical analysis of the inuence of
different parameters on the microtensile bond strength of
adhesives to dentin. The Journal of Adhesive Dentistry
2008;10(1):7.
[7] Kim JS, Cho BH, Lee IB, Um CM, Lim BS, Oh MH, et al. Effect
of the hydrophilic nanoller loading on the mechanical
properties and the microtensile bond strength of an ethanol
based one bottle dentin adhesive. Journal of Biomedical
Materials Research Part B: Applied Biomaterials
2005;72(2):28491.
[8] Nakabayashi N, Ashizawa M, Nakamura M. Identication of
a resindentin hybrid layer in vital human dentin created in
vivo: durable bonding to vital dentin. Quintessence
International (Berlin, Germany: 1985) 1992;23(2):135.
[9] Van Meerbeek B, Yoshida Y, Snauwaert J, Hellemans L,
Lambrechts P, Vanherle G, et al. Hybridization effectiveness
of a two-step versus a three-step smear layer removing
adhesive system examined correlatively by TEM and AFM.
The Journal of Adhesive Dentistry 1999;1(1):7.
[10] Kanca 3rd J. Resin bonding to wet substrate. 1. Bonding to
dentin. Quintessence International (Berlin, Germany: 1985)
1992;23(1):39.
[11] Dickens SH, Cho BH. Interpretation of bond failure through
conversion and residual solvent measurements and Weibull
analyses of exural and microtensile bond strengths of
bonding agents. Dental Materials 2005;21(4):35464.
[12] Schmidt G, Malwitz MM. Properties of polymernanoparticle
composites. Current Opinion in Colloid & Interface Science
2003;8(1):1038.
[13] Johnsen BB, Kinloch AJ, Mohammed RD, Taylor AC, Sprenger
S. Toughening mechanisms of nanoparticle-modied epoxy
polymers. Polymer 2007;48(2):53041.
[14] Subramaniyan AK, Sun C. Toughening polymeric
composites using nanoclay: crack tip scale effects on

1050

[15]

[16]

[17]

[18]

[19]

[20]
[21]
[22]

[23]
[24]

[25]

[26]

[27]
[28]

d e n t a l m a t e r i a l s 2 8 ( 2 0 1 2 ) 10411050

fracture toughness. Composites Part A: Applied Science and


Manufacturing 2007;38(1):3443.
Atai M, Solhi L, Nodehi A, Mirabedini SM, Kasraei S, Akbari
K, et al. PMMA-grafted nanoclay as novel ller for dental
adhesives. Dental Materials 2009;25(3):33947.
Ueda J, Gang W, Shirai K, Yamauchi T, Tsubokawa N. Cationic
graft polymerization onto silica nanoparticle surface in a
solvent-free dry-system. Polymer Bulletin 2008;60(5):61724.
Solhi L, Atai M, Nodehi A, Imani M, Ghaemi A, Khosravi K.
Poly(acrylic acid) grafted montmorillonite as novel llers for
dental adhesives: synthesis, characterization and properties
of the adhesive. Dental Materials 2012;28(4):36977.
Atai M, Nekoomanesh M, Hashemi SA, Yeganeh H. Synthesis
and characterization of BTDA-based dimethacrylate dental
adhesive monomer and its interaction with Ca2+ ions.
Journal of Applied Polymer Science 2002;86(13):32469.
Messersmith PB, Giannelis EP. Synthesis and barrier
properties of poly(caprolactone) layered silicate
nanocomposites. Journal of Polymer Science Part A: Polymer
Chemistry 1995;33(7):104757.
ANSI/ADA. Specication no. 27. Direct lling resins 1993.
ISO 4049. Dentistry-polymer-based restorative materials
2009.
Wu YP, Jia QX, Yu DS, Zhang LQ. Modeling Youngs modulus
of rubberclay nanocomposites using composite theories.
Polymer Testing 2004;23(8):9039.
Hbaieb K, Wang QX, Chia YHJ, Cotterell B. Modelling stiffness
of polymer/clay nanocomposites. Polymer 2007;48(3):9019.
Sadr A, Shimada Y, Tagami J. Effects of solvent drying time
on micro-shear bond strength and mechanical properties of
two self-etching adhesive systems. Dental Materials
2007;23(9):11149.
Choi YS, Choi MH, Wang KH, Kim SO, Kim YK, Chung IJ.
Synthesis of exfoliated PMMA/Na-MMT nanocomposites via
soap-free emulsion polymerization. Macromolecules
2001;34(26):897885.
Greesh N, Hartmann PC, Cloete V, Sanderson RD. Adsorption
of 2-acrylamido-2-methyl-1-propanesulfonic acid (AMPS)
and related compounds onto montmorillonite clay. Journal
of Colloid and Interface Science 2008;319(1):211.
Pavia DL, Lampman GM. Introduction to spectroscopy.
Brooks/Cole Pub Co.; 2009.
Golebiewski J, Galeski A. Thermal stability of nanoclay
polypropylene composites by simultaneous DSC and TGA.

[29]

[30]
[31]

[32]

[33]

[34]

[35]

[36]

[37]

[38]

[39]

[40]

[41]

Composites Science and Technology 2007;67(1516):


34427.
McNeill IC, Sadeghi SMT. Thermal stability and degradation
mechanisms of poly(acrylic acid) and its salts: part 1
poly(acrylic acid). Polymer Degradation and Stability
1990;29(2):23346.
Mark JE. Physical properties of polymers handbook.
Springer-Verlag; 2007.
Penn RW, Craig RG, Tesk JA. Diametral tensile strength and
dental composites. Dental Materials 1987;3(1):
468.
Cao T, Fasulo PD, Rodgers WR. Investigation of the shear
stress effect on montmorillonite platelet aspect ratio by
atomic force microscopy. Applied Clay Science
2010;49(12):218.
Wu Y-P, Jia Q-X, Yu D-S, Zhang L-Q. Modeling Youngs
modulus of rubberclay nanocomposites using composite
theories. Polymer Testing 2004;23(8):9039.
Cascone MG. Dynamicmechanical properties of bioarticial
polymeric materials. Polymer International 1997;43(1):
5569.
Ahmed S, Jones F. A review of particulate reinforcement
theories for polymer composites. Journal of Materials
Science 1990;25(12):493342.
Halpin JC, Pagano NJ. The laminate approximation for
randomly oriented brous composites. Journal of Composite
Materials 1969;3(4):7204.
Affdl JCH, Kardos JL. The HalpinTsai equations: a review.
Polymer Engineering & Science 1976;16(5):
34452.
Tucker Iii CL, Liang E. Stiffness predictions for unidirectional
short-ber composites: review and evaluation. Composites
Science and Technology 1999;59(5):65571.
Fornes TD, Paul DR. Modeling properties of nylon 6/clay
nanocomposites using composite theories. Polymer
2003;44(17):49935013.
Sadat-Shojai M, Atai M, Nodehi A, Khanlar LN.
Hydroxyapatite nanorods as novel llers for improving the
properties of dental adhesives: synthesis and application.
Dental Materials 2010;26(5):47182.
Van Meerbeek B, Willems G, Celis JP, Roos J, Braem M,
Lambrechts P, et al. Assessment by nano-indentation of the
hardness and elasticity of the resindentin bonding area.
Journal of Dental Research 1993;72(10):1434.

Potrebbero piacerti anche