You are on page 1of 12

journal of the mechanical behavior of biomedical materials 119 (2021) 104498

Contents lists available at ScienceDirect

Journal of the Mechanical Behavior of Biomedical Materials


journal homepage: http://www.elsevier.com/locate/jmbbm

On the properties of nanosilicate-based filled dental adhesives: Synthesis,


characterization, and optimized formulation
Laleh Solhi a, b, *, Mohammad Atai a, Azizollah Nodehi a, Mohammad Imani a
a
Iran Polymer and Petrochemical Institute (IPPI), P.O. Box 14965/115, Tehran, Iran
b
Department of Bioproducts and Biosystems, School of Chemical Engineering, Aalto University, P.O. Box 16300, FI-00076, Aalto, Finland

A R T I C L E I N F O A B S T R A C T

Keywords: Objective: In this study, we incorporated hybrid nanoparticles (poly (acrylic acid)-grafted nanoclay/nanosilica,
Dental adhesives respectively, with platelet and spherical morphologies, abbreviated as PAA-g–NC–Sil) in different concentrations
Nanoparticles (0, 0.2, 0.5, 1, 2 and 5 wt%) to an experimental dentin bonding system and investigated the physical properties
Poly (acrylic acid)
of the filled adhesive and its shear bond strength (μ-SBS) to dentin. We subsequently compared the properties of
Bond strength
Nanoclay
the adhesives containing PAA-g–NC–Sil with previously studied adhesives containing poly (methacrylic acid)-g-
Nanosilica nanoclay (PMA-g-NC) (Solhi et al., 2012a), poly (acrylic acid)-g-nanoclay (PAA-g-NC) (Solhi et al., 2012b), and
the hybrid poly (methacrylic acid)-grafted-nanoclay-nanosilica (PMA-g–NC–Sil) (Solhi et al., 2020).
Materials and methods: In a set of previous publications and the present paper, we grafted poly (acrylic acid)
(PAA) or poly (methacrylic acid) (PMA) onto the surface of pristine Na-MMT nanoclay (Cloisite® Na+) through
free radical polymerization of monomer in an aqueous media in the presence or absence of nanosilica particles.
We characterized the resulting modified nanoparticles (PMA-g-NC, PAA-g-NC, PMA-g–NC–Sil and PAA-g–NC–Sil)
using GPC, FTIR, TGA, and XRD. We then incorporated the modified particles as functionalized fillers to
experimental dentin adhesives in different concentrations and studied the stability of modified fillers dispersion
by separation analysis. We also studied the properties of the photo-cured adhesive matrices using FTIR, TEM,
SEM, EDXA, and XRD. We examined the shear bond strength of the adhesives (containing different contents of
each modified filler, separately) to human premolar teeth. The results were analysed and compared statistically.
Results: The results confirmed that the polymers have been grafted onto the surface of nanoclay. An exfoliated
structure for the nanoclay platelets in the photo-cured adhesive containing PAA-g–NC–Sil was observed. Addition
of 0.5 wt% of PAA-g–NC–Sil to the experimental adhesive increased the shear bond strength and the dispersion
stability in comparison to unfilled adhesive. The same trend was also observed for adhesives containing PMA-g-
NC, PAA-g-NC, and PMA-g–NC–Sil. The adhesive containing PAA-g–NC–Sil showed the best dispersion stability
and subsequently the highest shear bond strength in the optimal concentration among adhesives containing the
four available fillers (PMA-g-NC, PAA-g-NC, PMA-g–NC–Sil and PAA-g-NC-Sil).
Significance: Addition of poly (acrylic acid) modified nanoparticles to the experimental dentin adhesives resulted
in higher shear bond strength due to the potential interactions between the carboxylic acid functional groups on
the surface of the modified particles and the dentin structure. Between the poly (acrylic acid) and poly (meth­
acrylic acid), the former acid with higher PKa performed better. Addition of the spherical nanosilica particles to
the adhesives containing platelet nanoclay helped to better exfoliate the platelets resulting in improved μ-SBS
and dispersion stability.

1. Introduction adhesive layer. Incorporation of microfillers in dental adhesives could


adversely affect their polymerization reaction and subsequently their
Addition of fillers to dental adhesives is usually a strategy to improve clinical performance (Nassif and El Askary, 2019; Wang et al., 2014).
the strength (Miyazaki et al., 1995), modify the viscosity (Pashley et al., However, this is not the situation with nanofiller-containing adhesives
2002), and induce antibacterial properties (Melo et al., 2013) in the (Nassif and El Askary, 2019). Metal oxide nanoparticles (Saffarpour

* Corresponding author. Iran Polymer and Petrochemical Institute (IPPI), P.O. Box 14965/115, Tehran, Iran.
E-mail address: Laleh.solhi@aalto.fi (L. Solhi).

https://doi.org/10.1016/j.jmbbm.2021.104498
Received 29 January 2021; Received in revised form 21 March 2021; Accepted 23 March 2021
Available online 31 March 2021
1751-6161/© 2021 Elsevier Ltd. All rights reserved.
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 1. FT-IR spectra of nanosilica, Na-MMT and PAA-g–NC–Sil. Characteristic peaks of PAA confirm the grafting reaction.

inorganic silicate nanoparticles with polymers, which was proved to be


Table-1
an effective way to overcome the problems (Mazloom Jalali et al., 2016,
Mw and PDI of de-bonded polymers by reverse ion exchange, and homo-
2017). Another strategy concerns with the incorporation of silicate
polymers produced during the reaction and isolated by Soxhlet extraction sys­
tem for different fillers. nanofillers with different morphologies such as spherical, nanorod, and
platelet like or a mixture of these morphologies, into the dentin bonding
Polymer Filler Mw PDI
formulations, which have been shown to improve the properties of the
De-bonded PMA PMA-g-NC 2300 1.4 adhesive layer, effectively (Solhi et al., 2020; Sadat-Shojai et al., 2010;
Homo-polymer PMA extracted by PMA-g-NC 7500 1.8
Kim et al., 2005).
Soxhlet system
De-bonded PAA PAA-g-NC 1300, 8500 1.1,1.3
In accordance with the first strategy, different polymer modified
Homo-polymer PAA extracted by PAA-g-NC 7500 1.8 layered nanoclays were incorporated to experimental dental adhesives
Soxhlet system to improve their properties (Solhi et al., 2012a, 2012b; Atai et al., 2009).
De-bonded PMA PMA- 1000 and 1.3 and In these studies, pristine Na-MMT nanoclay (a platelet like nanosilicate),
g–NC–Sil 48,700 1.4
was modified through graft polymerization of methyl methacrylate,
Homo-polymer PMA extracted by PMA- 35,200 2.0
Soxhlet system g–NC–Sil methacrylic acid, and acrylic acid, which resulted in significant im­
De-bonded PAA PAA- 1000 and 1.2 and provements in the shear bond strength of the experimental dental ad­
g–NC–Sil 20,700 1.4 hesives (Solhi et al., 2012a, 2012b, 2020). In a more recent study and in
Homo-polymer PAA extracted by PAA- 19,000 2.8
accordance with the second strategy we utilized spherical nanosilica
Soxhlet system g–NC–Sil
particles along with the layered montmorillonite fillers to improve their
exfoliation and consequently the properties of the bonding agent in
et al., 2016; Lohbauer et al., 2010; Sun et al., 2011, 2017), metal which the particles were incorporated (Solhi et al., 2020). We discov­
nanoparticles (Melo et al., 2013; Li et al., 2014), hydroxyapatite nano­ ered that the small spheres of nanosilica can potentially lay between the
rods (Sadat-Shojai et al., 2010), and silicate nanoparticles (Solhi et al., nanoclay platelets, prevent the nanoclay platelets from aggregation in
2012a, 2012b, 2020; Schulz et al., 2008; Atai et al., 2009; Mazloom-­ the adhesive matrix resulting in improved exfoliation and dispersion
Jalali et al., 2020; Kim et al., 2005) have so far been added to dental stability. However, the shear bond strength of the adhesive containing
adhesives and have improved at least some properties of the adhesive 0.5 wt% PMA-g–NC–Sil did not increase significantly in comparison to
layer. Among dental adhesives, dentin bonding agents have attracted the adhesive containing optimal amount of PMA-g-NC (Solhi et al.,
special attention, since adhesion to dentin is more challenging. The 2012a, 2020). Another important observation in our previous studies
strength of the restoration depends on the strength of the hybrid layer was the fact that adhesives containing PAA-g-NC showed a better
(Nakabayashi et al., 1982; Kanca, 1992). Therefore, many of these dispersion in the adhesive and a higher shear bond strength in com­
studies have been focused on the improvement of the properties of parison to those containing PMA-g-NC (Solhi et al., 2012a, 2012b). The
dentin bonding systems. One of the most common strategies of me­ Pka values of methacrylic acid (MA) and acrylic acid (AA) are 4.65 and
chanically reinforcing polymeric materials is to add fillers to the matrix. 4.26, respectively (Ibarra-Montaño et al., 2015). Therefore, we expect
The common problems of addition of the fillers into the solvent diluted PAA to be a stronger acid and be more reactive than PMA. Acrylic acid
resin matrices, such as dentin adhesives, are incompatibility, agglom­ can interact more readily with Ca2+ ions of hydroxyapatite (Atai et al.,
eration, and sedimentation (Solhi et al., 2012a, 2012b; Atai et al., 2009; 2002) providing chemical interactions between the adhesive and dentin
Kim et al., 2005). Our group has followed two main strategies to over­ structure. Considering all the previous results, we expected the optimal
come these problems. One strategy was the surface modification of the nanosilicate filled dentin adhesive comprises dual morphology of

2
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 2. TGA curves of Na-MMT, neat PAA, nanosilica, PAA-g–NC–Sil and deb–NC–Sil-PAA. The grafting percentage of PAA on fillers surface in PAA-g–NC–Sil is 18.5
wt%, while after the reverse ion exchange process, polymer grafting percentage declined to about 9.2 wt%.

Fig. 3. XRD pattern of a) Na-MMT, b) nanosilica, c) PAA-g–NC–Sil, d) deb–NC–Sil-PAA, and e) Adh-PAA-g–NC–Sil. The interlayer spacing, d001, for Na-MMT and deb-
nanoclay-nanosilica-PAA is 9.6 Å, while PAA-g–NC–Sil, Adh-PAA-g–NC–Sil, and nanosilica show no peak in the shown range, implying full dispersion and exfoliation.

spherical nanosilica and layered montmorillonite fillers and PAA surface adhesive and μ-SBS of the adhesive containing different amounts of
modification. PAA-g–NC–Sil were investigated. Furthermore, we compared the prop­
In this study, we pursued the synthesis and characterization of poly erties of the adhesives containing PAA-g–NC–Sil with the properties of
(acrylic acid)-grafted-nanoclay/nanosilica (PAA-g–NC–Sil) and the adhesives containing PMA-g-NC (Solhi et al., 2012a), PAA-g-NC (Solhi
dispersion stability of this nanofiller in an experimental dilute dental et al., 2012b), and PMA-g–NC–Sil (Solhi et al., 2020), which original

3
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 4. Separation analysis of the adhesives con­


taining 1 wt% of a) PMA-g-NC (Solhi et al.,
2012a), b) PAA-g-NC (Solhi et al., 2012b), c)
PMA-g–NC–Sil (Solhi et al., 2020) and d)
PAA-g–NC–Sil during the total time span of 20 h.
The grafted polymer on the nanoparticles lowers
the overall density of the particles and increases
their sedimentation time in the adhesive. a is
reprinted from “Dental Materials, Solhi, L.; Atai,
M.; Nodehi, A.; Imani, M.; 28, 10, A novel dentin
bonding system containing poly (methacrylic
acid) grafted nanoclay: Synthesis, characteriza­
tion and properties, 1041–1050, copyright
(2012)” (Solhi et al., 2012a) with permission from
Elsevier. b is reprinted from “Dental Materials,
Solhi, L.; Atai, M.; Nodehi, A.; Imani, M., Ghaemi,
A.; Khosravi, K.; 28, 4, Poly (acrylic acid) grafted
montmorillonite as novel fillers for dental adhe­
sives: Synthesis, characterization and properties
of the adhesive, 369–377, copyright (2012)”
(Solhi et al., 2012b) with permission from Elsev­
ier. c is reprinted from “Journal of the Mechanical
Behavior of Biomedical Materials, Solhi, L.; Atai,
M.; Nodehi, A.; Imani, M., 109, Poly (methacrylic
acid) modified spherical and platelet hybrid
nanoparticles as reinforcing fillers for dentin
bonding systems: Synthesis and properties, 103,
840, copyright (2020)” (Solhi et al., 2020) with
permission from Elsevier.

data were published previously by our group. The adhesive samples (Solhi et al., 2012a, 2012b, 2020; Atai et al., 2009). High values of
containing 0.5 wt% PAA-g–NC–Sil showed the maximum dispersion microshear bond strength translate to better adhesion to dentin, while
stability and shear bond strength among all other tested formulations high dispersion stability is important from practical point of view, since
with different silicate morphologies (spherical, layered, and hybrid of PAA-g–NC–Sil gives the dentist a more homogenous composition of the
spherical and layered) and different polymer modifications (poly filled adhesive throughout the time of application.
(methyl methacrylate), poly (methacrylic acid), poly (acrylic acid))

4
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 5. Sedimentation behavior of adhesives con­


taining 1 wt% of Na-MMT, PMA-g-NC (Solhi et al.,
2012a), PAA-g-NC (Solhi et al., 2012b),
PMA-g–NC–Sil (Solhi et al., 2020), and
PAA-g–NC–Sil (transmittance versus time) at the
middle point of the test tube.
The PMA-g–NC–Sil curve is reprinted from “Jour­
nal of the Mechanical Behavior of Biomedical
Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani, M.,
109, Poly (methacrylic acid) modified spherical
and platelet hybrid nanoparticles as reinforcing
fillers for dentin bonding systems: Synthesis and
properties, 103,840, copyright (2020)” (Solhi
et al., 2020) with permission from Elsevier.

2.2. Methods
Table 2
DC% results for Adper™ Single Bond 2 and the adhesives con­
2.2.1. Graft polymerization of AA onto nanoclay platelets in the presence of
taining 0, 0.2, 0.5, 1, 2 and 5 wt% of PAA-g–NC–Sil. The mea­
surements are carried out in triplicates. There is no significant
nanosilica
difference between the DC% of all adhesive groups (p > 0.05). We graft polymerized AA onto the surface of Na-MMT in the presence
Groups that do not share a letter are statistically different. of nanosilica according to a previous method (Solhi et al., 2012a, 2012b,
2020). Briefly, we heated a dispersion of Na-MMT and nanosilica in
Adhesive DC%
water in the presence of AMPS and ammonium persulfate, monomer
Adper™ Single Bond 2 83.3 ± 3.9 (A)
(AA) and TDM. The polymerization reaction took place at 70 ◦ C in 1 h.
Blank 81.1 ± 4.3 (A)
0.2 wt% PAA-g–NC–Sil 80.2 ± 3.0 (A)
The product was precipitated in acetone, vacuum-filtered, washed
0.5 wt% PAA-g–NC–Sil 81.2 ± 3.5 (A) several times with distilled water until the conductivity of the liquid,
1 wt% PAA-g–NC–Sil 79.0 ± 3.3 (A) measured by a conductivity meter (JENWAY 4010, Staffordshire, UK)
2 wt% PAA-g–NC–Sil 82.8 ± 2.9 (A) showed a constant value. Finally, we washed the precipitate with
5 wt% PAA-g–NC–Sil 79.1 ± 3.5 (A)
ethanol and vacuum-dried the product at room temperature. We
Soxhlet-extracted the product in water to remove the homopolymer
2. Experimental (PAA) produced during the polymerization reaction. The purified
PAA-g–NC–Sil was vacuum-dried and powdered using a ball mill
2.1. Materials (PM100, Retch, Germany), passed through a 400 mesh (ASTM) sieve
and used as filler. A similar procedure was also followed for synthesis
We purchased 2-Hydroxyethyl methacrylate (HEMA), camphorqui­ and purification of PMA-g-NC (Solhi et al., 2012a), PAA-g-NC (Solhi
none (CQ), 2-ethyl-2-hydroxymethyl-1,3-propandiol trimethacrylate et al., 2012b) and PMA-g–NC–Sil (Solhi et al., 2020).
(TMPTMA), acrylic acid (AA), acetone, ethanol and methanol from
Merck (Darmstadt, Germany). We obtained N,N-dimethyl aminoethyl 2.2.2. Characterization of modified filler
methacrylate (DMAEMA) and tert-dodecyl mercaptan (TDM) from Fluka We analysed the pristine Na-MMT, nanosilica, and PAA-g–NC–Sil by
(Buchs, Germany). Evonik (Essen, Germany) kindly provided 2,2- Bis [4- FTIR spectroscopy (EQUINOX 55, Bruker, Hamburg, Germany) at a
(2-hydroxy-3-methacryloxypropoxy) phenyl] propane (Bis-GMA). We resolution of 4 cm− 1 and 32 scans in the range of 4000–400 cm− 1 using
obtained Cloisite® Na+ from Southern Clays Product, Inc. (Austin, TX, KBr disc technique. We performed thermo-gravimetric analysis of the
USA). Amorphous fumed silica with the primary particle size of 12 nm in pristine Na-MMT, nanosilica, PAA-g–NC–Sil, pure PAA, and debonded
diameter and surface area of 200 m2 g-1 (Aerosil® 200) was obtained PAA (TGA-1500, Polymer Laboratories, UK) in a range from room
from Evonik (Essen, Germany). We purchased 2-acrylamido-2-methyl-1- temperature to 600 ◦ C at a heating rate of 10 ◦ C.min− 1 and under N2
propane-sulfonic acid (AMPS), ammonium persulfate and lithium atmosphere. X-ray diffraction patterns of pristine Na-MMT, PAA-
chloride (LiCl) from Sigma–Aldrich (Munich, Germany). We obtained g–NC–Sil, the hybrid nanoclay/nanosilica after debonding the grafted
Adper™ Single Bond 2, a commercially available nanoparticle con­ PAA, and the adhesive containing the hybrid fillers were collected in the
taining dentin bonding, from 3 M ESPE (St. Paul, MN, USA). The 37.5% range of 2θ = 1.5–10◦ and step size of 0.02◦ , using a Philips X-ray
phosphoric acid gel (Kerr Gel Etchant) was obtained from SDS Kerr diffractometer (Philips, X’pert, Amsterdam, the Netherlands) with cop­
(Orange, CA, USA). We used deionized water throughout all the per target, ʎ = 1.54056 Å operating at a voltage of 40 kV and a current of
experiments. 40 mA at the rate of 2◦ min− 1. We applied the Bragg’s equation to
calculate the d001 spacing values.
To de-bond the polymer chains from the clay surface we used a

5
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 6. TEM images of the photo-cured


adhesive samples containing 0.5 wt%
a) Na-MMT (Solhi et al., 2020), b)
nanosilica, c) PMA-g-NC, d) PAA-g-NC,
e) PMA-g–NC–Sil (Solhi et al., 2020), f)
PAA-g–NC–Sil. The PMA and PAA
chains swell and allow the diffusion of
the adhesive matrix in the space be­
tween the nanocaly layers. In the
clay-silica containing formulations,
nanosilica particles are positioned be­
tween the nanoclay platelets and pre­
serve the exfoliated structure of
nanoclay platelets during the process of
solvent evaporation and curing. Nano­
clay platelets and nanosilica particles
are marked with orange and blue ar­
rows respectively. a, b, and e are
reprinted from “Journal of the Me­
chanical Behavior of Biomedical Mate­
rials, Solhi, L.; Atai, M.; Nodehi, A.;
Imani, M., 109, Poly (methacrylic acid)
modified spherical and platelet hybrid
nanoparticles as reinforcing fillers for
dentin bonding systems: Synthesis and
properties, 103,840, copyright (2020)”
(Solhi et al., 2020) with permission
from Elsevier.

reverse ion exchange process (Messersmith and Giannelis, 1995). We Note: The photoinitiator was added after mixing of all components and
filtered the de-bonded fillers (deb–NC–Sil-PAA), dried and characterized sonication to prevent un-wanted polymerization before applying the
them by the XRD and TGA techniques. We precipitated the dissolved adhesive). Pristine NA-MMT and PAA-g–NC–Sil were added to the ad­
de-bonded PAA, in acetone. The molecular weight of the de-bonded hesive in 0.2, 0.5, 1, 2 and 5 wt%. We dispersed the fillers in the ad­
polymer and the molecular weight of the un-bonded homopolymer hesive solution by ultra sonication using a probe sonication apparatus
which was extracted during Soxhlet-extraction were determined by GPC (Sonoplus UW2200, Bandelin, Berlin, Germany) for 3 min in an ice bath.
(GPC, Agilent 1100 series, Santa Clara, CA, USA) using Aquagel column We investigated the stability of the dispersions by a separation analyzer
of 7.5 × 300 mm (ID × L) with the flow rate of 1 ml min− 1 at 23 ◦ C. (LUMi-Reader® 416.1, LUM, Berlin, Germany) working with visible
Similar procedures were followed for characterization of PMA-g-NC light and intensity of 25% and tilt of 0, for 20 h (PAA-g–NC–Sil) and 30
(Solhi et al., 2012a), PAA-g-NC (Solhi et al., 2012b) and min (pristine nanoclay), including 255 intervals. The dispersion char­
PMA-g–NC–Sil (Solhi et al., 2020). acteristics of PMA-g-NC (Solhi et al., 2012a), PAA-g-NC (Solhi et al.,
2012b) and PMA-g–NC–Sil (Solhi et al., 2020), which had been
2.2.3. Characterization of adhesive before photo-curing measured in a similar procedure, were compared with the present study.
The adhesive was prepared according to a previously used experi­
mental dentin adhesive (14 wt% Bis-GMA, 26 wt% HEMA, 8 wt% 2.2.4. Characterization of adhesive after photo-curing
TMPTMA, 12 wt% UDMA, 40 wt% Ethanol, 1 phr butanedione (BD), Degree of conversion of adhesives containing 0, 0.2, 0.5, 1, 2 and 5

6
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

of caries-free extracted human premolars was exposed and ground by


silicon carbide paper with 600 grit (Soft Flex, Germany). The 72
extracted teeth were etched, washed and blot dried. The bonding sys­
tems containing 0, 0.2, 0.5, 1, 2 and 5 wt% PAA-g–NC–Sil were applied
on the dentin surface in the first six groups. The commercially available
dentin bonding (Adper™ Single Bond 2) was used for the last group as
control and applied according to the manufacturer. We filled silicone
tubes (internal diameter: 0.8 mm, height: 1.0 mm) with the resin com­
posite (Premise, Kerr, Orange, CA, USA) and light cured for 40s after
placing the tubes on the prepared dentin. Although different size of
adherent for μ-SBS test has been reported (Shimada et al., 2002;
McDonough et al., 2002), the size of adherent in this study was chosen as
in our previous studies to have comparable results. We removed the
tubes around the composite and performed micro-shear bond test after
having stored the specimens for 24 h in distilled water at 37 ◦ C. The
samples were observed by microscope to make sure there is no visible
defect in the sample. A thin wire (diameter 0.2 mm) was looped around
each composite cylinder. Pulling up the wire, we applied a shear force to
each specimen at a crosshead speed of 1 mm min− 1 using a universal
testing machine (STM 20, Santam, Tehran, Iran) until failure occurred.
We calculated the micro-shear bond strength by dividing the force at
break by the composite-dentin interface area, which leads to equation
(1) (Sadr et al., 2007).
4p
Fig. 7. Si map of the photo-cured adhesive samples containing 0.5 wt% PAA- μSBS = (1)
π D2
g–NC–Sil. A uniform distribution of the particles in the adhesive is observed.
where P is the force at break (N) and D (mm) is the diameter of the
wt% PAA-g–NC–Sil was measured using FT-IR spectroscopy and the composite cylinder, which was loaded on dentin surface. The μ-SBS of
method is explained in details in our previous publications (Solhi et al., the adhesives containing PMA-g-NC (Solhi et al., 2012a), PAA-g-NC
2012a, 2012b, 2020). Briefly, a thin layer of adhesive was placed be­ (Solhi et al., 2012b) and PMA-g–NC–Sil (Solhi et al., 2020) were
tween two polyethylene films and the absorbance peaks were obtained measured applying the same procedure.
by transmission mode of FT-IR before and after 40 s of photo-curing by a
dental light source with an irradiance of circa 600 mWcm− 2 (Optilux 2.2.6. Statistical analyses
501, SDS Kerr, Orange, CA, USA). We determined the degree of con­ We analysed and compared all the mechanical tests using one-way
version (DC%) from the ratio of absorbance intensities of aliphatic C– –C ANOVA and Tukey test at the significance level of 0.05. The reported
(peak at 1638 cm− 1) against internal reference of aromatic C–– C (peak at values are the average of at least 10 measurements for μ-SBS and 3
1608 cm− 1) before and after curing the specimens (Atai et al., 2004). measurements for degree of conversion.
The degree of conversion of adhesives containing PMA-g-NC (Solhi et al.,
2012a), PAA-g-NC (Solhi et al., 2012b) and PMA-g–NC–Sil (Solhi et al., 3. Results
2020) were measured using the same procedure.
To prepare samples for TEM imaging, adhesive specimens containing Fig. 1 shows the FTIR spectra of the nanosilica, pristine Na-MMT, and
0.5 wt% PAA-g–NC–Sil were solvent evaporated and light cured in a PAA-g–NC–Sil illustrating the characteristic peaks of PAA, which con­
mold. We used an ultra-microtome (DMU3, Reichert, Austria) to prepare firms the grafting reaction.
approximately 70 nm thick TEM specimens. We used a Philips TEM Table 1 tabulates the molecular weight (Mw) and poly-dispersity
(CM200, FEG, Amsterdam, Netherlands) and a Hitachi TEM (Hitachi, index (PDI) of extracted homopolymer by Soxhlet system and de-
H7600 TEM, Tokyo, Japan) for TEM observations. To prepare samples bonded polymer (PMA or PAA) from the surface of fillers (PMA-g-NC,
for SEM imaging, we gold coated the composite-dentin interfaces, after PAA-g-NC, PMA-g–NC–Sil, PAA-g–NC–Sil) after the reverse ion exchange
the micro-shear bond test. The samples were observed using SEM process.
(TESCAN, VEGAII, XMU, Brno, Czech Republic) with an accelerating Fig. 2 shows the TGA curves for Na-MMT, nanosilica, PAA-g–NC–Sil,
voltage of 15 kV. neat PAA, and finally, nanoclay and nanosilica mixture after performing
We used EDXA elemental composition analyzer to map the distri­ reverse ion exchange procedure on PAA-g–NC–Sil (deb–NC–Sil-PAA).
bution of inorganic fillers in the photo-cured dentin bonding systems. A Considering the ash content of neat PAA and PAA-g–NC–Sil in ther­
probe current of 2.0 × 10− 9 A, an accelerating voltage of 30 kV, and spot mogravimetric analysis (≈14 and 81 wt% at 550 ◦ C), the grafting per­
size of 500 nm with a collection time of 100 s, were used during the centage of PAA on fillers surface in PAA-g–NC–Sil was calculated as 18.5
mapping (EDXA, QX2, RONTEC Co.). We performed the XRD measure­ wt%. The grafted polymer percentage after the reverse ion exchange
ments on the photo-cured adhesive samples containing 1 wt% of PAA- process dramatically declined for deb–NC–Sil-PAA; however, the pro­
g–NC–Sil to verify the potential exfoliation of nanoclay platelets in the cess does not detach all the polymers on the nanoclay surface. After the
adhesive. The photo-cured adhesives containing PMA-g-NC (Solhi et al., reverse ion exchange process, polymer grafting percentage declined to
2012a), PAA-g-NC (Solhi et al., 2012b) and PMA-g–NC–Sil (Solhi et al., about 9.2 wt%.
2020) were characterized similarly. Fig. 3 presents the XRD patterns of Na-MMT, nanosilica, PAA-
g–NC–Sil, deb–NC–Sil-PAA and photo-cured specimen containing 1 wt%
2.2.5. Microshear bond strength (μ-SBS) test of PAA-g–NC–Sil (Adh-PAA-g–NC–Sil). The exact peak position in the
Micro-shear bond strength to dentin was measured according to XRD patterns and the related interlayer spacing, d001 are calculated by
previous works (Solhi et al., 2012a, 2012b, 2020; Atai et al., 2009; Bragg’s equation. According to XRD results, the interlayer spacing, d001
Schmidt and Malwitz, 2003). In brief, the dentin surface of seven groups for Na-MMT and deb-nanoclay-nanosilica-PAA is 9.6 Å, while PAA-

7
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 8. Microshear bond strength results for Adper™ Single Bond 2, blank adhesive, and adhesive samples containing different weight percentages of PMA-g-NC
(Solhi et al., 2012a), PAA-g-NC (Solhi et al., 2012b), PMA-g–NC–Sil (Solhi et al., 2020), and PAA-g–NC–Sil. The y-error bars indicate the standard deviation. The
incorporation of polyacid modified nanoclays to the experimental dentin adhesive improved microshear bond strength of the adhesive in 0.2 wt% to 0.5 wt%
concentration of the fillers. Groups that do not share a letter are statistically different. (P < 0.05).
The results of bond strength of Adh-PMA-g-NC are reproduced from “Dental Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani, M.; 28, 10, A novel dentin bonding
system containing poly (methacrylic acid) grafted nanoclay: Synthesis, characterization and properties, 1041–1050, copyright (2012)” (Solhi et al., 2012a) with
permission from Elsevier. The results of bond strength of Adh-PAA-g-NC are reproduced from “Dental Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani, M., Ghaemi, A.;
Khosravi, K.; 28, 4, Poly (acrylic acid) grafted montmorillonite as novel fillers for dental adhesives: Synthesis, characterization and properties of the adhesive,
369–377, copyright (2012)” (Solhi et al., 2012b) with permission from Elsevier. The results of bond strength of Adh-PMA-g–NC–Sil are reproducedd from “Journal of
the Mechanical Behavior of Biomedical Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani, M., 109, Poly (methacrylic acid) modified spherical and platelet hybrid
nanoparticles as reinforcing fillers for dentin bonding systems: Synthesis and properties, 103,840, copyright (2020)” (Solhi et al., 2020) with permission
from Elsevier.

g–NC–Sil and nanosilica show no peak in the shown range. et al., 2020), and PMA-g–NC–Sil.
Fig. 4 shows the sedimentation behavior of the PMA-g-NC (Solhi We investigated the distribution of modified particles in the adhesive
et al., 2012a), PAA-g-NC (Solhi et al., 2012b), PMA-g–NC–Sil (Solhi by examination of the cross section of cured adhesives using EDXA. The
et al., 2020), and PAA-g–NC–Sil in the dental adhesive. Fig. 5 shows a Si map of the cured adhesives containing 0.5 wt% PAA-g–NC–Sil is
comparison between the transmittance values of the test tubes, con­ collected and displayed in Fig. 7.
taining Na-MMT, PMA-g-NC (Solhi et al., 2012a), PAA-g-NC (Solhi et al., Fig. 8 shows the micro-shear bond strengths. The results show a
2012b), PMA-g–NC–Sil (Solhi et al., 2020), and PAA-g–NC–Sil, with significant increase in the bond strength of the adhesive containing 0.5
respect to the time at 27.7 mm below the top of the tube (the middle wt% PAA-g–NC–Sil, compared to the adhesives containing different
point of the tube). After 30 min the transmittance for the adhesive concentrations of PMA-g-NC, PAA-g-NC, PMA-g–NC–Sil, Adper™ Single
containing 1 wt% Na-MMT is 100% at the middle point of the tube. The Bond 2, and the blank sample (p < 0.05).
transmittance values at the same point is 92% for the adhesive con­ Figs. 9 and 10 illustrate the SEM micrographs of the fracture area in
taining 1 wt% PMA-g-NC (Solhi et al., 2012a), 83% for the adhesive micro-shear bond strength test for the adhesives containing PMA-g-NC
containing 1 wt% PAA-g-NC (Solhi et al., 2012b), 47% for the adhesive (Solhi et al., 2012a), PAA-g-NC (Solhi et al., 2012b), PMA-g–NC–Sil
including 1 wt% PMA-g–NC–Sil (Solhi et al., 2020), and 36% for the (Solhi et al., 2020), and PAA-g–NC–Sil, for 0.5 and 5 wt% filler content.
adhesive having 1 wt% PAA-g–NC–Sil, after 20 h.
The DC% for the adhesives containing 0, 0.2, 0.5, 1, 2 and 5 wt% of 4. Discussion
PAA-g–NC–Sil and a commercial dentin bonding system (Adper™ Single
Bond 2) are given in Table 2. The measurements are carried out in The reaction conditions, which involved heat, AA monomer, nano­
triplicates for each sample. clay platelets, nanosilica particles, AMPS, and ammonium persulfate
Fig. 6 shows the TEM images of the photo-cured specimens of the facilitated potentially both graft polymerization onto clay and silica
adhesives containing 0.5 wt% of Na-MMT, nanosilica, PMA-g-NC (Solhi surfaces and homopolymerization of AA in aqueous phase. Therefore,
et al., 2012a), PAA-g-NC (Solhi et al., 2012b), PMA-g–NC–Sil (Solhi the reaction product was expected to be a mixture of poly (acrylic acid)-

8
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

Fig. 9. SEM micrographs of the fracture area of micro-shear


bond strength test for the adhesives containing 0.5% of a1)
PMA-g-NC, b1) PAA-g-NC (Solhi et al., 2012b), c1)
PMA-g–NC–Sil (Solhi et al., 2020) and d1) PAA-g–NC–Sil,
and 5% of a2) PMA-g-NC (Solhi et al., 2012a), b2) PAA-g-NC
(Solhi et al., 2012b), c2) PMA-g–NC–Sil (Solhi et al., 2020)
and d2) PAA-g–NC–Sil. a1 and a2 are reprinted from “Dental
Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani, M.; 28, 10, A
novel dentin bonding system containing poly (methacrylic
acid) grafted nanoclay: Synthesis, characterization and
properties, 1041–1050, copyright (2012)” (Solhi et al.,
2012a) with permission from Elsevier. c1 and c2 are
reprinted from “Journal of the Mechanical Behavior of
Biomedical Materials, Solhi, L.; Atai, M.; Nodehi, A.; Imani,
M., 109, Poly (methacrylic acid) modified spherical and
platelet hybrid nanoparticles as reinforcing fillers for dentin
bonding systems: Synthesis and properties, 103,840, copy­
right (2020)” (Solhi et al., 2020) with permission from
Elsevier.

grafted-nanosilica, poly (acrylic acid)-grafted-nanoclay and poly 2008), but during the harsh conditions, applied in the process of
(acrylic acid) homopolymer chains. Poly (acrylic acid)-grafted- washing and extraction, almost all of these hydrogen bonds were
nanosilica can be formed by attachment of AMPS onto the surface of broken. In a test batch, only nanosilica was surface modified using the
nanosilica via hydrogen bonds between the amido groups of AMPS and above method and the washing process was proved to be effective in
the hydroxyl groups present on the surface of the filler (Greesh et al., removing all the polymer from nanosilica surface (confirmed by TGA

9
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

GPC chromatogram of homo-polymer PMA extracted by Soxhlet


system from PMA-g-NC, showed a peak correlating to a Mw of 7500 (PDI
= 1.8). The GPC chromatogram of homo-polymer PAA extracted by
Soxhlet system from PAA-g-NC, showed a peak indicating a Mw of 7500
(PDI = 1.8). GPC chromatogram of homo-polymer PMA extracted by
Soxhlet system from PMA-g–NC–Sil showed the relating peak to the Mw
of 35,200 (PDI = 2.0). These Mws are related to the homo-polymers,
which were produced during the grafting reaction. The relatively low
molecular weight of homopolymers is a consequence of using TDM as a
transfer agent in the polymerization reaction. These Mws are reasonably
close to the Mw of the small fractions, observed in GPC chromatogram of
de-bonded polymers detached from PAA-g-NC, PMA-g–NC–Sil and PAA-
g–NC–Sil, strengthening the hypothesis that the small fractions are the
trivial amounts of homopolymers, which were not removed during the
Soxhlet extraction probably because of the entanglements with the
grafted polymer chains. Having debonded the grafted chains by reverse
ion exchange process, the homopolymers untangle and dissolve in
water, which consequently, are observed in the GPC chromatogarms.
The presence of inorganic nanoclay and nanosilica particles in­
creases the thermal stability of poly (acrylic acid). The sharp two-stage
drop, which are seen in the thermogram of neat poly (acrylic acid) is not
observed in the TGA of the PAA-g–NC–Sil (Fig. 2). The residual weight in
the TGA curve (ash content) was used to calculate the percentage of
grafted polymer.
The peak at 2θ = 9.2◦ in XRD pattern of Na-MMT is disappeared for
Fig. 10. SEM micrographs of the fracture area of micro-shear bond strength test
PAA-g–NC–Sil, which is shown in Fig. 3. The peak at 2θ = 9.2◦ corre­
for the adhesives containing 0.5% of PAA-g–NC–Sil, showing an adhe­
sponds to an interlayer spacing of d001 = 9.6 Å for Na-MMT layers and
sive failure.
disappears in XRD pattern of PAA-g–NC–Sil, due to the exfoliation/
intercalation of the nanoclay platelets. We expected the PAA chains on
measurements). In our calculations in this paper, we assumed, that the
the surface of nanoclay platelets and the presence of the nanosilica
polymer content was solely attached to the surface of nanoclay.
spheres help the platelets to be disengaged. The reverse ion exchange
The appearance of two characteristic IR peaks at 1736 and 2955
process on PAA-g–NC–Sil, which removes the majority of the attached
cm− 1 for PAA-g–NC–Sil are assigned to carbonyl group and stretching
polymer chains from the surface of nanoclay platelets, leads to the re-
vibration of C–H in PAA structures, respectively (Pavia et al., 2009).
appearance of the peak close to 2θ = 9.2◦ in XRD. This observation
Since the extraction method is expected to remove the homopolymers,
shows that interlayer spacing of d001 = 9.6 Å is back for PAA-g–NC–Sil
the peaks are corresponded to the PAA chains attached to nanocaly in
after reverse ion exchange process and the remained polymer (around 9
PAA-g–NC–Sil, which confirms the grafting.
wt% calculated from TGA curve of de-bonded polymer in Fig. 2) is not
The GPC chromatograms of de-bonded PAA from PAA-g–NC–Sil
enough to increase the basal spacing or is attached on the edges of the
shows a large and a small peak, which after deconvolution present a Mw
nanoclay platelets.
of around 1000 (PDI = 1.2) for the major peak and a Mw of around
According to literature, the interaction of AMPS with clay mainly
20,700 (PDI = 1.4) for the small peak. GPC chromatograms of homo-
occurs by proton transfer from sulfonic acid to amide group and a sub­
polymer PAA extracted by Soxhlet system from PAA-g–NC–Sil, show
sequent ion exchange with sodium ion on silicate (Solhi et al., 2012b,
the relating peaks to the Mw of 19,000 (PDI = 2.8). This Mw is related to
2020; Greesh et al., 2008; Yeong Suk et al., 2001). There are two other
the homo-polymers, which were produced during the grafting reaction.
possibilities for attachment of AMPS to the surface of montmorillonite,
A similar trend was observed across our studies for polymerization of
which has been explained in details elsewhere (Solhi et al., 2020).
PAA and PMA on nanoclay surface in absence and presence of
Briefly, one is by formation of hydrogen bonds between the amide
nanosilica.
groups of AMPS and water molecules surrounding the sodium cation on
In Table 1, the GPC chromatogram of de-bonded PMA from PMA-g-
the silicate surface and the other by formation of ion-dipole interactions
NC showed a peak relating to Mw of around 2300 (poly dispersity index
between sulfonate groups and the sodium ion on the silicate surface.
(PDI) = 1.4). The GPC chromatogram of de-bonded PAA from PAA-g-NC
Based on the TGA results, from 18.5 wt% polymer on the surface of
showed two peaks, which after deconvolution reveal a Mw of about 1300
nanoclay in PAA-g–NC–Sil, 9.3% (almost half) were expected to be
(PDI = 1.1) and a Mw of about 8500 (PDI = 1.3). The GPC chromatogram
attached to the surface via the ion exchange of sodium ions by amide
of de-bonded PMA from PMA-g–NC–Sil showed a major and a smaller
groups of AMPS, which were detached during reverse ion exchange
peak, which after deconvolution showed a Mw of around 1000 (PDI =
process. The other half of the polymer (9.2%), which remained attached
1.3) for the larger peak and a Mw of around 48,700 (PDI = 1.4) for the
on the nanoclay after reverse ion exchange process, were expected to be
small peak.
attached by the other two mechanisms (Solhi et al., 2020). The fact that
The large peak in these GPC chromatograms (PAA-g-NC, PMA-
the remained polymers after reverse ion exchange process, which are
g–NC–Sil and PAA-g–NC–Sil) is linked to the de-bonded polymer from
expected to be attached by the two side mechanisms do not play any role
polymer-grafted-nanoclay, while the small peak might be related to the
in increasing the basal spacing, suggests that they are mainly located at
minor amount of homopolymers not being washed away properly during
the edges of the platelets rather than the surface of the nanoclays. We
the extraction, which were washed away during the long process of
think the nanosilica particles are withdrawn from the nanoclay galleries
reverse ion exchange in contact with an aqueous solution with a high
during the long process of reverse ion exchange, due to their hydro­
ionic strength. The higher Mw of homopolymers in comparison to de-
philicity and small size and therefore, do not have an effect on increasing
bonded polymer chains, can be explained due to less steric hindrance
the basal spacing after reverse ion exchange process.
in homopolymerization in comparison to the grafting in the small space
There are 255 curves in each graph in Fig. 4, which contribute to
between the clay platelets.
different time spans and visualize precipitation behavior. Each curve

10
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

represents the transmission versus the position of the test tube, in a Si-mapping of the adhesive containing 0.5 wt% PAA-g–NC–Sil shows
specific time after starting the test. Ethanol is a good solvent for PMA a uniform distribution of the particles in Fig. 7.
and PAA (solubility parameters of ethanol, PAA and PMA are 11.7, 12.7 Within the scope of our studies on nanofilled dentin adhesives, Fig. 8
and 12.48 cal1/2.cm− 3/2, respectively) (Mark, 2007). Polymer chains shows that the bond strength increases to a maximum level in a specific
grafted onto the surface of the nanoparticles, swell in the adhesive and concentration of the modified nanoclay (0.2 wt% for PAA-g-NC (Solhi
force the nanoclay platelets to separate from each other, which results in et al., 2012b) and 0.5 wt% for PMA-g-NC (Solhi et al., 2012a),
decreasing the overall density of the filler particles and increasing the PMA-g–NC–Sil (Solhi et al., 2020) and PAA-g–NC–Sil) and then gradu­
sedimentation time. Fumed silica particles are smaller with a lower ally decreases with the increase of the filler content. Fillers reinforce the
density (d-nanoclay = 2.86 g mL− 3, d-nanosilica = 2.2 gr.ml− 3) and elastic modulus of the adhesive (Van Meerbeek et al., 1993) and prevent
have many hydroxyl groups on the surface; therefore, they float easier in the cohesive failure of the adhesive layer. To reinforce the cohesive
the matrix. strength of the adhesive layer and improving the mechanical properties,
Different patterns of sedimentation in the figure, imply the different layered nanoclay or a mixture of layered and spherical fillers were added
interactions of PMA and PAA with the adhesive matrix. PAA seems to to the adhesive formulation. To improve the adhesive strength of this
have a better interaction with the matrix ingredients and exhibits longer layer, the silicate fillers were surface modified with carboxylic acid
sedimentation times. Additionally, inclusion of nanosilica in the fillers bearing polymers such as PMA and PAA. The carboxylic acid groups on
shows better dispersion stability, which can have two possible reasons. polymer chains grafted to the surface of nanoclay in PMA-g-NC (Solhi
First, the nanoclay platelets in PMA-g–NC–Sil and PAA-g–NC–Sil are et al., 2012a), PAA-g-NC (Solhi et al., 2012b), PMA-g–NC–Sil, (Solhi
better separated (the XRD patterns and TEM images confirmed the et al., 2020) and PAA-g–NC–Sil reacted with Ca2+ ions of the hydroxy­
separated nanoclay platelets in PMA-g–NC–Sil and PAA-g–NC–Sil), apatite in dentin and improved the bond strength of the adhesive.
hence the possibility of coagulation of the particles in the adhesives, The adhesive samples with low filler contents (0.5 wt%) allow
which may lead to faster sedimentation, is less than the PMA-g-NC and penetration into the dentin tubules and provide stronger micro­
PAA-g-NC cases. Second, the lower transmission by the end of the test mechanical interlocking through the formation of reinforced resin tags,
can be a result of the better interaction of nanosilica surface and the which is demonstrated in Fig. 9-a1 for PMA-g-NC (Solhi et al., 2012a),
matrix and consequent longer sedimentation time in the adhesives Fig. 9-b1 for PAA-g-NC (Solhi et al., 2012b), Fig. 9-c1 for PMA-g–NC–Sil
comparing to the nanoclay particles. In practice, both of these reasons (Solhi et al., 2020) and Fig. 9-d1 for PAA-g–NC–Sil. When the filler
are effective. content increased to 5 wt% most tubules in Fig. 9-a2 for PMA-g-NC
In Fig. 5, after 30 min the transmission for the adhesive containing 1 (Solhi et al., 2012a), Fig. 9-b2 for PAA-g-NC (Solhi et al., 2012b),
wt% Na-MMT is 100% at the middle point of the tube (height of 27.7 Fig. 9-c2 for PMA-g–NC–Sil, (Solhi et al., 2020) and Fig. 9-d2 for
mm from the top of the tube). The transmission values at the same point PAA-g–NC–Sil are empty. In high concentrations of all four types of
for modified fillers do not reach 100% even after 20 h. The results fillers, insufficient resin penetration into the inter-tubular dentin matrix
represented in Figs. 4 and 5, confirm that the dispersion stability of the caused insufficient resin tag and poor retention. Therefore, micro-shear
adhesives containing 1 wt% of different fillers follows the sequence bond strength in the adhesive samples containing 5 wt% of all four types
shown in equation (2). of fillers decreased in comparison to the samples containing optimal
amount of the fillers. An adhesive failure pattern for the adhesive con­
Dispersion stabilityadhesive-Na-MMT ≪ Dispersion stabilityadhesive-PMA-g-NC <
taining 0.5 wt% PAA-g–NC–Sil is observed in Fig. 10.
Dispersion stabilityadhesive-PAA-g-NC < Dispersion stabilityadhesive-PMA-g–NC–Sil Addition of nanosilica to the reaction mixture in case of PMA-
< Dispersion stabilityadhesive-PAA-g–NC–Sil 2
g–NC–Sil, increased the dispersion stability of the adhesive containing
There is no significant difference between the DC% of the adhesives 0.5 wt% filler compared to PMA-g-NC and PAA-g-NC as it is shown in
with different filler contents (in the range of 79–83%) and the Adper™ Figs. 4 and 5 and our previous studies (Solhi et al., 2012a, 2012b, 2020).
Single Bond 2 (p > 0.05). Therefore, the properties of the adhesives are In the case of PAA-g–NC–Sil, due to higher acidity of PAA in comparison
comparable. to PMA and its higher tendency to react with hydroxyapatite, we ex­
As Fig. 6-a shows, in the photo-cured dentin bonding system con­ pected higher shear bond strength. This effect is shown in Fig. 8.
taining 0.5 wt% Na-MMT, the multilayer stacks of the nanoclay platelets Grafting PAA instead of PMA on the surface of the hybrid fillers
are observed. Fumed silica particles are shown in Fig. 6-b. Fig. 6-c and improved the μ-SBS and the dispersion stability of the filled adhesive,
Fig. 6-d show the photo-cured adhesive containing 0.5 wt% PMA-g-NC significantly. Surface grafting of the hybrid nanofillers with PAA instead
and PAA-g-NC, respectively. The number of nanoclay platelets in an of PMA was a successful strategy that increased the μ-SBS at the optimal
array decreases by surface modification of nanoclay platelets. The filler content by 80%. As it is shown in Fig. 5, the % transmission for the
polymer chains prevent the nanoclay platelets from clinging together adhesive containing 1 wt% PMA-g–NC–Sil at the middle point of the
after water was removed from the product in the drying process of the tube is 44.6% after 400 min, while this number is about 10.3% for
PMA-g-NC and PAA-g-NC. When the dried fillers are added to ethanol- PAA-g–NC–Sil. This shows the better compatibility of PAA with the
based dentin bonding system, the PMA and PAA chains swell and adhesive matrix environment, which enables the PAA-g–NC–Sil particles
allow the diffusion of the adhesive matrix in the space between the to remain floated for a longer time.
nanocaly layers. After removing the solvent, during the process of photo- From practical point of view, among all the polymer modified
curing, the adhesive monomers polymerize in this space and result in montmorillonites we prepared for application in dentin adhesives, the
increased basal spacing in comparison to adhesives containing Na-MMT. PAA-g–NC–Sil showed the best dispersion stability and bond strength
The XRD patterns confirm this (at least) partially exfoliated structure when added to the adhesive. Based on this information, PAA-g–NC–Sil is
(Fig. 3). In Fig. 6-e and Fig. 6-f, the TEM of adhesives containing 0.5 wt% an optimal nanoclay-based filler to be added to dentin adhesives for
PMA-g–NC–Sil and PAA-g–NC–Sil are shown. In these samples, in addi­ improving the immediate performance and dispersion stability of the
tion to the effect of polymer chains grafted onto the surface of nanoclay adhesives.
platelets, nanosilica particles are also left between the nanoclay platelets
after solvent evaporation and help to observe exfoliation in the structure 5. Conclusion
of nanoclay platelets in the photo-cured dentin bonding system. This
effect can be especially better observed in Fig. 6-f for the adhesive In this paper, we reported the original results regarding synthesis,
containing PAA-g–NC–Sil. In this Figure, nanoclay platelets are no characterization, and addition of PAA-g–NC–Sil to an experimental
longer exist in groups and the layers are completely exfoliated. dentin adhesive and investigated the effect on the properties of the ad­
hesive. In addition, we compared the results of a series of related studies

11
L. Solhi et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104498

in which, polymer surface modified nanosilicates (PMA-g-NC, PAA-g- McDonough, W.G., Antonucci, J.M., He, J., Shimada, Y., Chiang, M.Y., Schumacher, G.E.,
Schultheisz, C.R., 2002. A microshear test to measure bond strengths of
NC, PMA-g–NC–Sil, PAA-g–NC–Sil) were prepared, characterized, and
dentin–polymer interfaces. Biomaterials 23 (17), 3603–3608.
added to dentin adhesives. Incorporation of an optimal amount of Melo, M.A.S., Cheng, L., Zhang, K., Weir, M.D., Rodrigues, L.K., Xu, H.H., 2013. Novel
modified fillers to the experimental dentin adhesive resulted in a sig­ dental adhesives containing nanoparticles of silver and amorphous calcium
nificant increase in μ-SBS, in comparison to the blank adhesive in case of phosphate. Dent. Mater. 29 (2), 199–210.
Messersmith, P.B., Giannelis, E.P., 1995. Synthesis and barrier properties of poly
all four modified fillers (PMA-g-NC, PAA-g-NC, PMA-g–NC–Sil, PAA- (ϵ-caprolactone)-layered silicate nanocomposites. J. Polym. Sci. Polym. Chem. 33
g–NC–Sil). Incorporation of 0.5 wt% PAA-g–NC–Sil in the experimental (7), 1047–1057.
dentin bonding system resulted in the highest shear bond strength and Miyazaki, M., Ando, S., Hinoura, K., Onose, H., Moore, B.K., 1995. Influence of filler
addition to bonding agents on shear bond strength to bovine dentin. Dent. Mater. 11
most stable dispersion in the adhesive among all the modified fillers, (4), 234–238.
making it the optimum montmorillonite-based filler for reinforcing Nakabayashi, N., Kojima, K., Masuhara, E., 1982. The promotion of adhesion by the
dentin bonding systems, which has been so far introduced. infiltration of monomers into tooth substrates. J. Biomed. Mater. Res. 16 (3),
265–273.
Nassif, M., El Askary, F., 2019. Nanotechnology and nanoparticles in contemporary
Declaration of competing interest dental adhesives. In: Nanobiomaterials in Clinical Dentistry. Elsevier, pp. 163–198.
Pashley, E.L., Agee, K.A., Pashley, D.H., Tay, F.R., 2002. Effects of one versus two
applications of an unfilled, all-in-one adhesive on dentine bonding. J. Dent. 30 (2–3),
The authors declare that they have no known competing financial 83–90.
interests or personal relationships that could have appeared to influence Pavia, D.L., Lampman, G.M., Kriz, G.S., Vyvyan, J.R., 2009. Introduction to
the work reported in this paper. Spectroscopy. Brooks/Cole Pub. Co., California.
Sadat-Shojai, M., Atai, M., Nodehi, A., Khanlar, L.N., 2010. Hydroxyapatite nanorods as
novel fillers for improving the properties of dental adhesives: synthesis and
References application. Dent. Mater. 26 (5), 471–482.
Sadr, A., Shimada, Y., Tagami, J., 2007. Effects of solvent drying time on micro-shear
Atai, M., Nekoomanesh, M., Hashemi, S., Yeganeh, H., 2002. Synthesis and bond strength and mechanical properties of two self-etching adhesive systems. Dent.
characterization of BTDA-based dimethacrylate dental adhesive monomer and its Mater. 23 (9), 1114–1119.
interaction with Ca2+ ions. J. Appl. Polym. Sci. 86 (13), 3246–3249. Saffarpour, M., Rahmani, M., Tahriri, M., Peymani, A., 2016. Antimicrobial and bond
Atai, M., Nekoomanesh, M., Hashemi, S., Amani, S., 2004. Physical and mechanical strength properties of a dental adhesive containing zinc oxide nanoparticles. Braz. J.
properties of an experimental dental composite based on a new monomer. Dent. Oral Sci. 15 (1), 66–69.
Mater. 20 (7), 663–668. Schmidt, G., Malwitz, M.M., 2003. Properties of polymer–nanoparticle composites. Curr.
Atai, M., Solhi, L., Nodehi, A., Mirabedini, S.M., Kasraei, S., Akbari, K., Babanzadeh, S., Opin. Colloid Interface Sci. 8 (1), 103–108.
2009. PMMA-grafted nanoclay as novel filler for dental adhesives. Dent. Mater. 25 Schulz, H., Schimmoeller, B., Pratsinis, S.E., Salz, U., Bock, T., 2008. Radiopaque dental
(3), 339–347. adhesives: dispersion of flame-made Ta2O5/SiO2 nanoparticles in methacrylic
Greesh, N., Hartmann, P.C., Cloete, V., Sanderson, R.D., 2008. Adsorption of 2-acryla­ matrices. J. Dent. 36 (8), 579–587.
mido-2-methyl-1-propanesulfonic acid (AMPS) and related compounds onto Shimada, Y., Yamaguchi, S., Tagami, J., 2002. Micro-shear bond strength of dual-cured
montmorillonite clay. J. Colloid Interface Sci. 319 (1), 2–11. resin cement to glass ceramics. Dent. Mater. 18 (5), 380–388.
Ibarra-Montaño, E.L., Rodríguez-Laguna, N., Sánchez-Hernández, A., Rojas- Solhi, L., Atai, M., Nodehi, A., Imani, M., 2012a. A novel dentin bonding system
Hernández, A., 2015. Determination of pKa values for acrylic, methacrylic and containing poly (methacrylic acid) grafted nanoclay: synthesis, characterization and
itaconic acids by 1H and 13C NMR in deuterated water. J. Appl. Solut. Chem. Model. properties. Dent. Mater. 28 (10), 1041–1050.
4 (1), 7–18. Solhi, L., Atai, M., Nodehi, A., Imani, M., Ghaemi, A., Khosravi, K., 2012b. Poly (acrylic
Kanca III, J., 1992. Resin bonding to wet substrate. I. Bonding to dentin. Quintessence acid) grafted montmorillonite as novel fillers for dental adhesives: synthesis,
Int. 23 (1). characterization and properties of the adhesive. Dent. Mater. 28 (4), 369–377.
Kim, J.S., Cho, B.H., Lee, I.B., Um, C.M., Lim, B.S., Oh, M.H., Chang, C.G., Son, H.H., Solhi, L., Atai, M., Nodehi, A., Imani, M., 2020. Poly (methacrylic acid) modified
2005. Effect of the hydrophilic nanofiller loading on the mechanical properties and spherical and platelet hybrid nanoparticles as reinforcing fillers for dentin bonding
the microtensile bond strength of an ethanol-based one-bottle dentin adhesive. systems: synthesis and properties. Journal of the Mechanical Behavior of Biomedical
J. Biomed. Mater. Res. Part B: Applied Biomaterials: An Official Journal of The Materials 103840.
Society for Biomaterials, The Japanese Society for Biomaterials, and The Australian Sun, J., Forster, A.M., Johnson, P.M., Eidelman, N., Quinn, G., Schumacher, G.,
Society for Biomaterials and the Korean Society for Biomaterials 72 (2), 284–291. Zhang, X., Wu, W.-l., 2011. Improving performance of dental resins by adding
Li, F., Weir, M.D., Fouad, A.F., Xu, H.H., 2014. Effect of salivary pellicle on antibacterial titanium dioxide nanoparticles. Dent. Mater. 27 (10), 972–982.
activity of novel antibacterial dental adhesives using a dental plaque microcosm Sun, J., Petersen, E.J., Watson, S.S., Sims, C.M., Kassman, A., Frukhtbeyn, S., Skrtic, D.,
biofilm model. Dent. Mater. 30 (2), 182–191. Ok, M.T., Jacobs, D.S., Reipa, V., 2017. Biophysical characterization of
Lohbauer, U., Wagner, A., Belli, R., Stoetzel, C., Hilpert, A., Kurland, H.-D., Grabow, J., functionalized titania nanoparticles and their application in dental adhesives. Acta
Müller, F.A., 2010. Zirconia nanoparticles prepared by laser vaporization as fillers Biomater. 53, 585–597.
for dental adhesives. Acta Biomater. 6 (12), 4539–4546. Van Meerbeek, B., Willems, G., Celis, J.-P., Roos, J., Braem, M., Lambrechts, P.,
Mark, J.E., 2007. Physical Properties of Polymers Handbook, vol. 1076. Springer. Vanherle, G., 1993. Assessment by nano-indentation of the hardness and elasticity of
Mazloom Jalali, A., Afshar Taromi, F., Atai, M., Solhi, L., 2016. Effect of reaction the resin-dentin bonding area. J. Dent. Res. 72 (10), 1434–1442.
conditions on silanisation of sepiolite nanoparticles. J. Exp. Nanosci. 11 (15), Wang, Z., Shen, Y., Haapasalo, M., Wang, J., Jiang, T., Wang, Y., Watson, T.F., Sauro, S.,
1171–1183. 2014. Polycarboxylated microfillers incorporated into light-curable resin-based
Mazloom Jalali, A., Afshar Taromi, F., Atai, M., Solhi, L., 2017. An insight into the dental adhesives evoke remineralization at the mineral-depleted dentin. J. Biomater.
silanization of montmorillonite nanoparticles. Chem. Eng. Commun. 204 (2), Sci. Polym. Ed. 25 (7), 679–697.
176–181. Yeong Suk, C., Min Ho, C., Ki Hyun, W., Sang Ouk, K., Yoon Kyung Kim, a., Chung*, I.J.,
Mazloom-Jalali, A., Taromi, F.A., Atai, M., Solhi, L., 2020. Dual modified nanosilica 2001. Synthesis of Exfoliated PMMA/Na-MMT Nanocomposites via Soap-free
particles as reinforcing fillers for dental adhesives: synthesis, characterization, and Emulsion Polymerization.
properties. Journal of the Mechanical Behavior of Biomedical Materials 103904.

12

You might also like