Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

)1818-1008-0000-0000P:ADI dPOFP( LAD NNALPR SORPPNOROSSFFORP

Accepted Article
Article Type : Article

Reinforcement of polymethyl methacrylate denture base resin with ZnO


nanostructures

Nehal A. Salahuddina*, Maged El-Kemaryb, Ebtisam M. Ibrahimb


a
Department of Chemistry, Faculty of Science, Tanta University , 31527, Tanta, Egypt
email: salahuddin.nehal@yahoo.com; nehal.attaf@science.tanta.edu.eg; Fax: +20-40-3350804
b
Nanotechnology center, Department of Chemistry, Faculty of Science, Kafrelsheikh University, 33516,
KafrElSheikh, Egypt

Abstract
Polymethyl methacrylate - zinc oxide (PMMA/ZnO) nanocomposites with different ZnO

morphology and quantities were prepared by in situ curing of MMA/PMMA used in denture

base material in the presence of ZnO nanostructured materials. The composition, structure,

and morphology of the nanocomposites were characterized by Fourier-Transform infrared

spectra (FT-IR), X-ray diffraction (XRD), transmission electron microscopy (TEM), scanning

electron microscopy (SEM) and thermogravimetric analysis (TGA). The effect of ZnO

morphology and contents on the mechanical properties including impact, flexural, and

hardness was studied. The results demonstrated that the mechanical and thermal properties of

denture base materials were improved by incorporation of nano-ZnO into the polymethyl

methacrylate matrix. The impact strength was highly improved upon using ZnO nanotubes.

However, the flexural strength was improved by using ZnO nanospheres.

Keywords: Nanocomposites, Mechanical properties, Thermal properties, Zinc oxide.

This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1111/ijac.12802
This article is protected by copyright. All rights reserved.
1. Introduction
Acrylic resins are used in dentistry for different purposes such as removable base
Accepted Article
plates, functional appliances and denture bases, due to its ease of processing, accurate fit,

stability in the oral environment, low cost, nontoxic, odorless, tasteless, nonirritant,

lightweight, color-matching ability, easy for patient to keep clean, high polish attainable, and

easy to repair after fracture.1-3 Due to its superior physical, biological, and esthetic properties,

polymethyl methacrylate (PMMA) has replaced previous denture materials such as vulcanite,

nitrocellulose, phenol formaldehyde, vinyl plastics, and porcelain. Pure PMMA resin is a

colorless, transparent solid that may be tinted to provide any shade and degree of
4
translucency for proper esthetics for dental applications, individual impression trays and

temporary crowns. However, PMMA possesses insufficient surface hardness and mechanical

strength. A common clinical problem is the fracture of dentures which usually occurs due to

accidents or when the patient applies heavy occlusal forces to dentures. Fracture of dentures

is caused due to fatigue and impact failure and involves expensive repair costs. Removable

prostheses are exposed into two types of stress; intra-orally when the patient applies a high

repeated mastication force that occurs between maxilla and mandible Jaws leading to fatigue

phenomena and extra-orally stress where it is easily broken as a result of dropping. A high

performance of denture based resin requires good impact strength, fracture toughness and

hardness. Adequate hardness of the denture surface possesses good resistance to scratching

during cleaning process and facilitates easy polishing of prosthodontics.5 In addition;

Microorganism accumulation on acrylic appliances is one of the most discussable questions

in using them. Indeed the growth of some microorganisms such as candida albicans and

bacteria like streptococcus mutans on these appliances has been demonstrated.6 All these

factors increase the need to improve the antimicrobial capability and the strength of denture

base materials.

This article is protected by copyright. All rights reserved.


There are two approaches, which can be used to improve the strength of PMMA. The

first approach is to increase the strength of denture base polymer by adding a cross-linking
Accepted Article
agent of poly-functional monomer such as polyethylene glycol dimethacrylate. The second

approach is to devise a reinforcement of denture base polymer with rods or fibers such as

rubber phase, metal oxides, metal wires or metal nets, or clays.4, 7-13 There are many factors

such as; matrix structure, dispersed phase content, shape, size and distribution having a

leading effect on the mechanical properties of reinforced resin.14,15

In recent years, ZnO nanoparticles have attracted increasing attention as one of the

multifunctional inorganic nanoparticles due to the special optical, chemical, electrical,

biological, biocompatibility, low cost, non-toxic properties and long-term environmental

stability. ZnO nanoparticles can potentially apply to catalysts, photocatalyst for degradation

of waste water pollutants, semiconductors, varistors, piezoelectric devices, gas sensors, UV-

shielding materials, field-emission displays, rubber, dental and medical, coatings and

pigments, ceramic, concrete, antibacterial and bactericide, and in chemical synthesis.16-21

ZnO nanoparticles have been added in various compositions to polymers, to enhance their

special characteristic and mechanical properties.22-24

Recently a number of studies have been reported on the preparation of PMMA/ZnO


25
nanocomposites by physical and chemical methods. Khan et al demonstrated a facile

solution-based method to synthesize PMMA/ZnO nanocomposite via a solution mixing and

casting method. Transparent PMMA/ZnO nanocomposites with a high concentration of ZnO

nanoparticles (up to 10%) have also been prepared by in situ bulk polymerization in the

presence of the prepared nanoparticles, which gives improved dispersion of nanoparticles

compared to simple mixing.26, 27 Hung and Whang 28 have reported on the luminescence of

This article is protected by copyright. All rights reserved.


PMMA/ZnO films where the ZnO particles had been formed by a sol-gel method and treated

with a silane coupling agent before the polymerization. Pre et al 29 have prepared transparent
Accepted Article
luminescent nanocomposites powders using solution mixing and solvent casting of ZnO
30
nanoparticles in PMMA. Zhuang et al presented fabrication of the visible light (VL) –

traversing and ultraviolet (UV)-shielding zinc oxide quantum dots (ZnO QDs)-poly (methyl

methacrylate) (PMMA) nanocomposite films by incorporating suitable UV-absorbing ZnO

QDs into a transparent PMMA matrix. The ZnO QDs had been formed by a sol-gel route and

treated with 3-(trimethoxysilyl)propylmethacrylate (TPM) as a coupling agent. Zhang et al 31

have prepared transparent PMMA/ZnO nanocomposite films through free-radical

copolymerization of asymmetric zinc methacrylate acetate and in-situ thermal decomposition.


32
Agrawal et al reported solution mixing of ZnO nanoparticles into PMMA and showed a

better dispersion and good thermal stability of PMMA/ZnO nanocomposites thin films. Up to

dates no literature exists regarding the ability of ZnO nanoparticles to reinforce denture base

material, though the reports claim that addition ZnO nanoparticles to PMMA has resulted in

improved physical properties.

In view of the fact that the physical and mechanical properties arise from the

morphology of the material, this work pursues the effect of different ZnO morphology and

content on the mechanical properties of PMMA. The reinforcement of denture base PMMA

with ZnO nanotubes and nanoparticles was studied. First, ZnO nanoparticles (ZnO NPs) and

nanotubes (ZnO NTs) were prepared followed, by impregnation of the nanostructured ZnO in

MMA monomer and mixing with PMMA then, cured as recommended by the manufacture.

Furthermore, flexural strength, impact and hardness were measured for the nanocomposites

and compared with pure PMMA. The hypothesis was that ZnO nanostructures increase the

mechanical properties of denture base polymer.

This article is protected by copyright. All rights reserved.


2. Materials and Methods

2.1. Materials
Accepted Article
One commercial type of heat-cured acrylic resin PMMA used in dentistry was

selected for this work (super acryl, clear type Spofa Dental, Praha), Zinc nitrate hexahydrate

(Zn(NO3)2.6H2O, 99% Sigma-Aldrich), sodium carbonate (Na2CO3, 99% Sigma-Aldrich),

sodium hydroxide pellets (NaOH, 99% Merck), Ethylene Diamine (EDA C2H8N2; 99%,

Merck), Ethanol ( C2H5OH, 95%, Pio chem) were used without further purification.

2.2. Methods

2.2.1. Preparation of ZnO nanoparticles

ZnO NPs were prepared by a precipitation method according to the previous report.33

Briefly, two solutions were prepared as follows: Solution A (0.1 mol Zn(NO3)2.6H2O) was

prepared by dissolving 29.747 g of Zinc nitrate hexahydrate in 200 ml distilled water; and

solution B (0.12 mol Na2CO3) was prepared by dissolving 12.7188 g of sodium carbonate in

240 ml distilled water. After that, the precursor was prepared by adding solution A to solution

B drop wise under vigorous stirring. The white precipitate was collected by filtration and

rinsed with distilled water three times. The solid was then washed with ethanol and dried at

100°C for 6 h. Finally, ZnO NPs were obtained after calcinations of the precursor in air at

350 °C for 2 h.

2.2.2. Preparation of ZnO Nanotubes

ZnO NTs were synthesized according to the previous report. 34Firstly, 80 g of NaOH

pellets and 29.748 g of Zn(NO3)2.6H2O were dissolved in deionized water (200 ml) to

This article is protected by copyright. All rights reserved.


prepare a precursor clear solution with a molar ratio of Zn2+:OH- of 1:20. Then 3 ml of this

solution (Zn2+: OH- =1:20) was mixed with 5 ml of distilled water. After that, 25 ml of pure
Accepted Article
ethanol was added, followed by 5 ml of ethylene diamine. The mixture was then kept in an

ultrasonic water bath for 40 min and subsequently transferred to a Teflon-lined stainless steel

autoclave. The autoclave was kept inside an electric oven set at 180 °C for 2 h and cooled.

The precipitate (ZnO NTs) was separated by centrifugation and washed several times with

deionized water and pure ethanol. The collected product was dried in a vacuum desiccator at

room temperature for 24 h.

2.2.3. Preparation of Polymethyl methacrylate (PMMA)

PMMA powder and MMA monomer liquid were mixed as recommended by the

manufacture and left to stand for 10-15 minutes. In this period, the resin changes its

consistency. The dough material was kneaded in clean fingers into a stick roll and forced into

the stone mold ready for each experiment, packed to a slight excess and the flask was closed.

Trial closure of the flask was performed to remove the excess dough, followed by final

closure under pressure. The curing cycle followed the manufacturer's recommendations (75

°C for 1.5 h followed by 100 °C for 1 h). Then they were removed from the water bath and

allowed to cool before deflasking. Deflasking, finishing and polishing of resin samples were

done by using carborundum paper and stones.

2.2.4. Preparation of PMMA/ZnO nanocomposites

MMA monomer liquid was stirred with different ratios of ZnO nanostructures (0.2

wt%, 0.4 wt%, 0.8 wt% and 1 wt%) then the suspension was mixed with PMMA powder and

This article is protected by copyright. All rights reserved.


the same procedure for preparation of PMMA was followed to prepare the composite. ZnO

ratio was calculated based on the total weight of PMMA& MMA.


Accepted Article
2.3. Characterization

The crystallinity of the ZnO nanostructures and PMMA/ZnO nanocomposites were

characterized by an X- ray diffractometer, XRD-6000, Shimadzu, Japan with Cu-kα radiation

(λ =1.5412 Å), 40 Kv, 30 mA in the 2θ range of 10-80° with 2°/min scanning rate. The

crystallite sizes (D) of selected samples were estimated using the Scherer's equation, D=Kλ/

(Bcosθ) where: k = constant (0.89 < k < 1), λ= wavelength of the X-ray, B = FWHM (Full

Width at Half Maximum) width of the diffraction peak, θ=diffraction angle.

Thermogravimetric analysis (TGA) was carried out on TGA-50, Shimadzu at a


heating rate of 10 °C min-1 under a continuous nitrogen flow (15 ml/ min).
Differential scanning calorimetry (DSC) was recorded using a DSC-60A instrument,

Shimadzu, from 30 °C to 200°C with a heating rate of 10°C/ min under nitrogen

atmospheres.

The morphological feature of the nanostructures and nanocomposites was observed by

a scanning electron microscope SEM-EDS, JSM-6510 LV, JEOL, Japan and a high-

resolution transmission electron microscope (HRTEM), JEM-2100, JEOL, Japan. The

samples for SEM were coated with gold and the samples for TEM analysis were prepared by

dropping dilute suspension of ZnO nanostructures in H2O onto copper meshes. Transmission

electron microscopy (TEM) specimens were cut from nanocomposite block using an ultra-

microtome, power tome XL with a glass knife. The ultra-thin films were cut by moving the

This article is protected by copyright. All rights reserved.


sample across a knife edge of glass. The ultra-thin flakes floats onto a trough filled with

water from where they collected on 200 mesh copper grids and dried at room temperature.
Accepted Article
Flexural testing was carried out on (Instron 5500R) Universal Testing Machine at

room temperature. The test specimens were prepared in the rectangular-shaped specimens (65

x 10 x 2.5 mm), five samples were tested and the average value was taken. The flexural test

performed at a crosshead speed of 5 mm/min. The impact strength test was carried out on

(Izod-Impact tester CS-137) test machine at room temperature. The test specimens were

prepared in the rectangular-shaped specimens (75 x 10 x 10 mm), each having a 2 mm deep

standard notch with 60° angle in the middle of the bar. Five samples were tested and the

average value was taken. Nanoindentation tests were carried out using a Nano Test Vantage

from Micro Materials, Inc., Wrexham, UK. Indentation data were obtained for the samples

using maximum loads of 10 mN. The loading rate was set to 0.1 mN/S. A Berkovich

diamond tip (three-sided pyramidal) diamond indenter was used for all indentations.

Statistical analyses were performed using SPSS statistical software program SPSS

Version 10 (SPSS, Inc, Chicago, IL, USA). The quantitative data were presented as means

and standard deviations. One-way analysis of variance (ANOV) was used to examine

variable effects. The test was considered significant when P˂0.05 and highly significant

when P˂0.001.

3. Results and discussion

Acrylic resin is the most commonly used material for the construction of denture due

to esthetics, ease of manipulation and low cost. However, this material is not ideal because of

This article is protected by copyright. All rights reserved.


its relatively low mechanical strength, which can cause the fracture of denture. Polymers

have been reinforced by adding materials such as metal oxides, clay minerals and carbon-
Accepted Article
graphite fibers. The size, shape and distribution of these particles in the polymer matrix play

a major role on the mechanical properties of particulate-filled polymer composites.35 In

addition, microbial plaque adheres to acrylic resin appliances with a wider adhesion area than
36
to natural teeth and mechanical methods have been shown ineffective in eradicating

microorganisms completely.37 The present study investigated the effect of ZnO

nanostructures with different morphology on flexural and impact strength. Accordingly,

nanocomposites have been developed that has self- sterilizing agent by incorporation ZnO

into PMMA.

3.1. X-ray diffraction (XRD) analysis

The XRD patterns of the ZnO NPs, ZnO NTs, pristine PMMA and of the

nanocomposites are shown in Figure 1 (a, b). The X-ray diffraction pattern of ZnO

nanoparticles is shown in Figure 1a. It was observed that all the peaks of ZnO nanoparticles

were indexed as hexagonal wurtzite structure of ZnO (lattice parameter, a=b= 0.3249 nm, c=

0.5206 nm) and the diffraction data are in good agreement with those taken from the Joint

committee of powder diffraction standards (JCPDS) card No. 36-1451). Diffraction peaks

related to impurity are not observed in the XRD pattern, confirming the high purity of the

synthesized product. According to the XRD analysis the average crystalline size of ZnO NPs

was calculated according to Scherer's equation. The highest three peaks were selected for this

calculation and the average diameter was found to be 28.33 nm. Figure 1b. Shows the XRD

pattern of ZnO nanotubes (powders). It is apparent that, the diffraction providing clear

This article is protected by copyright. All rights reserved.


evidence for the formation of hexagonal wurtzite structure, and no diffraction peaks of any

other phases were detected. 38


Accepted Article
Figure 1 (a, b) also, shows The XRD patterns of the pristine PMMA and PMMA/ZnO

nanocomposites loaded with different contents of ZnO NPs and ZnO NTs, respectively, after

curing. The diffraction pattern of PMMA shows a broad diffraction peak at 2θ= 14°, typical

of an amorphous material, together with two bands of lower intensities centered at 31° and

43°. The XRD patterns of PMMA/ ZnO nanocomposites show the broad, non-crystalline

peak (~14°) of PMMA together with the sharp diffraction peaks of ZnO nanostructures.

These peaks were assigned to be (100), (002) and (101) planes of the crystalline phase of

ZnO in which the intensity increase with increasing ZnO quantity. This confirms that the

ZnO maintained its structure in the composite and that the orientation of the PMMA chains

was also not influenced during the nanocomposite preparation process.39

3.2. Thermal analysis

The thermal behavior of PMMA and PMMA/ZnO nanocomposites was investigated

by TGA and DSC. The TGA curves obtained for the ZnO NPs, ZnO NTs, pristine PMMA,

PMMA/ZnO NPs and PMMA/ZnO NTs nanocomposites are shown in Figure 2 (a, b). The

results reveal that the trend of PMMA/ZnO nanocomposites degradation is similar to pure

PMMA; all nanocomposites degrade in a single step similar to pristine PMMA, suggesting
40
that degradation is due to random chain scission. The thermal stability factors, including

the initial decomposition temperatures where 5% weight loss occurred (TIDT, °C), the

temperatures with 10 wt% weight loss (T10, °C) and the mid-point of the degradation process

at which 50 wt% degradation (T50, °C), could be determined from the TGA thermograms.

This article is protected by copyright. All rights reserved.


The results obtained for the thermal stability factors of the nanocomposites are summarized

in Table 1. In case of PMMA/ZnO NPs all nanocomposites at all composition show higher
Accepted Article
TIDT, T10 and T50 than the pure PMMA resin, which implies an improvement in the thermal

stability. The decomposition temperatures of nanocomposites at 5, 10, and 50 wt% loss were

about~ 12, 14, and 22 °C higher than those of pure PMMA, respectively. On the other hand,

the TGA curves obtained for some PMMA/ZnO NTs composites exhibit initial degradation at

lower temperatures (100-200 °C) than the pristine PMMA resins; these results suggest that

the weight loss is due to the evaporation of absorbed water in the nanocomposites.40

The DSC results of PMMA and the PMMA/ZnO nanocomposites are also shown in

Table 1. The glass transition temperature (Tg) of PMMA (119 °C) is improved up to 126 at

0.4 wt% of ZnO NTs. This small value indicates that there is no interaction between particles

and the polymer.

27
Demir et al reported that the glass transition temperature of PMMA/ZnO

nanocomposites measured by DSC is found to be positioned between 102 to 109 °C, very

close to Tg of atactic PMMA (105 °C) at range of loading up to 11% due to absence of any

specific interaction between the particles and the polymer matrix. In another publication

DMA analysis of PMMA/ZnO composites shows that ZnO nanoparticles and nanorods

increase the storage modulus of nanocomposites and shift the Tg towards higher temperature

at low concentration (0.01-0.1 weight %) while at higher concentration (1 %) the

reinforcement effect is deteriorated which was ascribed to the aggregation of ZnO

nanorods.41

This article is protected by copyright. All rights reserved.


3.3. Mechanical properties

Dentures fractures usually result from two types of forces; flexural fatigue and
Accepted Article
impact. When the dentures are dropped the impact causes the fracture. Also repeated flexing

from chowing causes denture fatigue in the mouth. Impact tests measure the energy required

to break the specimen by applying a dynamic load. The flexural strength is a measure of

stiffness and resistance to fracture in which the specimen is under astatic load, which was

considered relevant to the loading characteristics of denture base in a clinical situation. The

hardness measure the resistance to scratching during the cleaning process and polishing.

Zappini G et al.,42 reported that evaluating the strength of acrylic resins depend on impact

strength which not simulate the clinical situation and to predict the clinical function, flexural

strength tests are preferred.35Previous study done by Arnold A M et al., have demonstrated

that ultimate flexural strength of denture base resin indicates the resistance of failure under a

flexural load.43There are several studies showed that addition of metal oxides decreased the

tensile strength35,44 and others demonstrated improvement in flexural strength.45,46Improving

mechanical properties of acrylic denture bases by addition particulate fillers has been studied

by many researches. Zuccari et al.,47 evaluated the effect of Al2O3, MgO and ZrO2 to

reinforce acrylic resins. Asor et al.,35 found that 2% ZrO2 exhibited the greatest improvement

in modules of elasticity, transverse strength, hardness and toughness. Schulze et al.,48

explained that increase in filler fraction is not necessary to increase the strength since high

content lead to more defects that weaken the material.

3.3.1 The impact strength

The mean, standard deviation, minimum and maximum values for impact strength and

ANOVA results of impact test are presented in Table 2. It was observed that when ZnO

This article is protected by copyright. All rights reserved.


nanoparticles and nanotubes were used, especially up to 0.8 wt% all the nanocomposites

improved impact strength up to 104 % higher than pristine PMMA. This improvement was
Accepted Article
related to the small particle dimension and proper distribution which lead to an increase in

crack length during the process of fracture and increase in energy absorption before

fracture.35 The maximum reinforcement effect was obtained, with the addition of ZnO

nanotubes (Table 2). The higher reinforcement effect in PMMA/ZnO NTs may be attributed

to the high aspect ratio of nanotube (15:1). However, nanoparticles aspect ratio can be

regarded as (1:1). To investigate the reasons for these changes, the fractured surface of each

specimen was characterized by SEM (Figure 3, 4). The fractured surface of PMMA (Figure

3b) is smooth, showing the characteristics of brittle fracture. When ZnO structures (Figure 3a,

4a) are used, the surface was rougher, the roughness increases with increasing the wt% of

ZnO. This was attributed to ZnO nanotubes had unique long tubular structures and the

polymer chains wound around the added nanotubes and formed a three dimensional network

structure without any bonding effect. As a result, the polymer chain could slip along the

nanotube axis under an external force so, bending stress and bending displacement were

improved. However, nanoparticles were just mechanically mixed in the composite and did

not combine with the surrounding polymer chains so, bending strength was improved but

bending displacement was not.49 Incorporation multiwalled carbon nanotubes (2%) into denture

base material showed poor fatigue resistance due to agglomeration and poor interfacial bonding

between multiwalled carbon nanotube and PMMA.50 The impact strength and surface hardness of

zirconia-reinforced high impact resin were not significantly different from that of zirconia –free high

impact resin.51

3.3.2. Flexural strength

The mean, standard deviation, minimum and maximum values for flexural strength

and ANOVA results of Flexural test are presented in Table 3. The results indicate that the

This article is protected by copyright. All rights reserved.


incorporation of ZnO could enhance denture base and the flexural strength increased at

loading (0.2-0.4 wt%) and then decreased with constant addition of ZnO (0.8-1 wt%). This
Accepted Article
decrease of flexural strength may be due to irregular distribution of ZnO into the PMMA

which causes stress concentration due to agglomeration based on TEM observations. The best

flexural strength was reached at 95.780 MPa and 94.196 MPa, for ZnO NPs and ZnO NTs

respectively, which are 14 %, 12 % higher than pristine PMMA, with the optimal content of

0.4-0.8 wt%. Schulze et al 48concluded that an increase in filler fractions creates more defects

that weaken the materials.

It is worth noting that the nanotubes endowed the PMMA with the best reinforcement

effect. When the content of ZnO nanotubes further increased the flexural performance

decreased. This implied that the addition of 0.8 weight% nanotubes provided the

nanocomposites with its maximum flexural performance. Because the ZnO nanotubes possess

the unique pore structures and large specific surface area, the addition of ZnO nanotubes to

the mixture of PMMA and MMA would influence the viscosity of PMMA-MMA dough and

the ratio of PMMA-MMA, thus produced an impact on the mechanical performance of

composite.49 Yadav et al 52
have compared the strength of acrylic resin reinforced with

aluminum oxide (5 wt%) and when processed by conventional water bath technique and

microwave energy. The results showed that flexural strength had decreased with addition of

Al2O3. This decrease of flexural strength of modified acrylic resin was explained by the

irregular distribution of Al2O3 filler particles into the acrylic resin which causes stress

concentration due to filler agglomeration. The second reason was due to untreated filler

particles, which decrease the bonding filler particles and the acrylic resin. This suggests that

to get a better bonding between the filler particles and acrylic resin the Al2O3 should be

treated with silane coupling agent. Addition of 2.5 wt% Al2O3 was responsible for a 6.36%

This article is protected by copyright. All rights reserved.


8
increase in flexural strength. However, the addition of 5 wt% Al2O3 powder caused a

5.82% decrease in flexural strength. This was explained by a decrease in the cross-section of
Accepted Article
the load-bearing polymer matrix; stress concentration because of too many filler particles;

changes in the modulus of elasticity of the resin and mode of crack propagation through the

specimen due to an increased amount of fillers; void formation from entrapped air and

moisture; incomplete wetting of the fillers by the resin; and the fact that the Al2O3 acts as an

interfering factor in the integrity of the polymer matrix. In another publication, the

incorporation of nano-sized TiO2 and SiO2 at concentrations of 0.5% and 1% by weight into

acrylic resin adversely affected the mechanical properties of polymerized material and

flexural strength values decreased with increase in concentration of TiO2. These findings

were attributed to the effect of nano-sized oxides on the internal structure of polymerized

PMMA. Dispersion of TiO2 NP in PMMA matrix adversely affects degree of conversion

which in turn leads to increase in the level of residual unreacted monomer that acts as

plasticizer.53

3.3.3. Nanoindentation test

Hardness is the material resistance to plastic deformation measured under an

indentation load.54 There have been studies which found a correlation between hardness and

wear resistance of artificial denture teeth.55 The effect of ZnO morphology and contents on

the hardness of the samples is summarized in Table 4. As shown in Table 4, all

nanocomposites have higher hardness values than that of pristine PMMA. This is attributed to

the dispersion of the nanoparticles in the nanocomposite on account of the imposed

restriction on matrix deformation by uniform distribution of the nanoparticles. It was reported

that PMMA unreinforced demonstrated higher hardness than reinforced PMMA due to the

presence of unreached monomr that act as plasticizer.56

This article is protected by copyright. All rights reserved.


3.4. TEM micrographs

Figure 5a displays TEM micrographs of ZnO NPs. TEM images show that the ZnO
Accepted Article
NPs are spherical shape and the average diameter is 25.4 nm. On the basis of high-resolution

TEM the formation of ZnO NTs was further confirmed (Figure 5b), and the analysis shows

that the mean length and wall thickness of the ZnO NTs were about 2.4 μm and 200 nm,

respectively.

In order to substantiate the dispersion of ZnO nanostructures, the PMMA/ZnO

nanocomposites samples were further characterized by TEM. Figures 6 and 7 show the TEM

images of PMMA/ZnO NPs and PMMA/ZnO NTs, respectively. As shown in (Figure 6)

microscopic observations indicated that the ZnO NPs were well distributed and dispersed in

the PMMA matrix when 0.2 wt% and 0.4 wt% were used, but at high concentration (0.8- 1

wt%) more agglomerations were noted. These results may explain why high content of ZnO

had the lowest value in flexural strength. It is worth mentioning that TEM images clearly

show that ZnO nanoparticles are agglomerated all over the matrix of PMMA at 1 wt%

content; this agglomeration is not shown at low content.

Limitation

Further research is needed to predict the success of nanocomposites in clinical

environment by measuring the material fatigue before and after thermocycling in artificial

saliva and evaluate the antimicrobial activity after incorporation ZnO nanoparticles.

4. Conclusion
PMMA nanocomposites reinforced with ZnO nanoparticles and nanotubes were prepared. It

is demonstrated that incorporating ZnO nanotubes can distinctly improve the impact strength

up to 104 % at 0.8 wt% ZnO NTs and flexural strength up to 14 % at 0.4 wt% ZnO

nanoparticles. TEM analysis indicated the absence of particle aggregation at this content. The

decreased in mechanical properties at 0.8- 1 wt% are due to aggregation. The highest impact

This article is protected by copyright. All rights reserved.


strength can be achieved by the higher aspect ratio of ZnO NTs. As determined by ANOVA,

the addition of ZnO nanostructures resulted in a highly significant increase in mechanical


Accepted Article
properties of heat-cured PMMA denture base resin. The TGA and DSC results, demonstrated

that nanocomposites have higher thermal stability and a glass transition temperature than

pristine PMMA.

References

1. Sodagar A, Kassaee MZ, Akhavan A, Javadi N, Arab S, Kharazifard M.J. Effect of


silver nano particles on flexural strength of acrylic resins. J Prosthodont Res. 2012;
56: 120-124.
2. Hong G, Murata H, Li Y, Sadamori S, Hamada T. Influence of denture cleaners on the
color stability of three types of denture base acrylic resin. J Prosthet Dent. 2009; 101:
205-213.
3. John J, Gangadhar SA, Shah L. Flexural strength of heat-polymerized polymethyl
methacrylate denture resin reinforced with glass, aramid, or nylon fibers. J Prosthet
Dent. 2001; 86: 424-427.
4. Hamza T, Wee AG, Alapati S, et al. The Fracture Toughness of Denture Base
Material Reinforced With Different Concentrations of POSS. J Macromol Sci., Pure
Appl Chem. 2004; 41A: 897-906.
5. Rakhshan V. Marginal integrity of provisional resin restoration materials a review of
the literatures. Saudi J Dent Res. 2015; 6: 33-40.
6. Radford DR, Sweet SP, Challacombe SJ, et al. Adherence of Candida albicans to
denture-base materials with different surface finishes. J Dent. 1998; 26: 577–583.
7. Kanie T, Fujii K, Arikawa H, et al. Flexural properties and impact strength of denture
base polymer reinforced with woven glass fibers. Dent Mater. 2000; 16:150-158.
8. Zhang C, Bai X, Lian X, et al. Study on morphology and mechanical properties of
PMMA-based nanocomposites containing POSS molecules or functionalized SiO2
particles. High Perform Polym. 2011; 23: 468-476.

This article is protected by copyright. All rights reserved.


9. Vojdani M, Bagheri R, Khaledi AA R. Effects of aluminum oxide addition on the
flexural strength, surface hardness, and roughness of heat- polymerized acrylic resin.
J Dent Sci. 2012; 7: 238-244.
Accepted Article
10. Xu J, Li Y, Yu T, et al. Reinforcement of denture base resin with short vegetable
fiber. Dent Mater 2013; 29: 1273-1279.
11. Tu M-G, Liang W-M, Wu T-C, et al. Improving the mechanical properties of fiber-
reinforced acrylic denture-base resin. Mater Des. 2009; 30: 2468-2472.
12. Salahuddin N, Shehata M. polymethyl methacrylate-montmorillonite composites:
preparation, characterization and properties. Polymer. 2001; 42: 8379-8385.
13. Salahuddin N. The effect of polyoxypropylene-montmorillonite intercalates on
polymethyl methacrylate. Polym Compos. 2009; 30: 13-21.
14. Hari PA, Kalavathy, Mohamed HS. Effect of glass fiber and silane treated glass fiber
reinforcement on impact strength of maxillary complete denture. Ann Essences Dent.
2011; 3: 7-12.
15. Vuorinen A-M, Dyer SR, lassila LVJ, Vallittu PK. Effect of rigid rod polymer filler
on the mechanical properties of poly-methyl methacrylate denture base material. Dent
Mater. 2008; 24: 708-713.
16. Chase DW, Kim BC. Characterization on Polystyrene/ Zinc Oxide nanocomposites
prepared from solution mixing. Poly Adv Technol. 2005; 16: 846-850.
17. Xiong M, Gu G, You B, et al. Preparation and characterization of poly (Styrene
butylacrylate) latex/ Nano-ZnO Nanocomposites. J Appl Polym Sci 2003; 90: 1923-
1931.
18. El-Kemary M, El-Shamy H, El-Mehasseb I. Photocatalytic degradation of
ciprofloxacin drug in water using ZnO nanoparticles. J. Lumin 2010; 130: 2327-2331.
19. Moezzi A, McDonagh A-M, Cortie M B. Zinc oxide particles: synthesis, properties
and applications. Chem Eng J 2012; 185-186: 1-22.
20. Schmidt-Mende L, MacManus-Driscoll JL. ZnO-nanostructures, defects, and devices.
Mater Today 2007; 10:40–48.
21. Ma X-Y, Zhang W-D. Effects of flower-like ZnO nanowhiskers on the mechanical,
thermal and antibacterial properties of waterborne polyurethane. Polym Degrad Stab
2009; 94: 1103-1109.
22. Ding KH, Wang GL, Zhang M. Characterization of mechanical properties of epoxy
resin reinforced with submicron-sized ZnO prepared via in situ synthesis method.
Mater Des 2011; 32: 3986-3991.

This article is protected by copyright. All rights reserved.


23. Harishanand KS, Nagabhushana H, Nagabhushana BM, et al. Comparative Study on
Mechanical Properties of ZnO, ZrO2 and CeO2 Nanometal Oxides Reinforced
Epoxy Composites. Adv Polym Sci Technol 2013; 3: 7-13.
Accepted Article
24. Ramezanzadeh B, Attar MM, Farzam M. Effect of ZnO nanoparticles on the thermal
and mechanical properties of epoxy-based nanocomposite. J Therm Anal Calorim
2011; 103: 731-739.
25. Khan M, Chen M, Wei C, et al. Synthesis at the nanoscale of ZnO into poly(methyl
methacrylate) and its characterization. Appl Phys A Mater Sci Process 2014;
117: 1085-1093.
26. Demir MM, Koynoy K, Akbey U, et al. Optical Properties of Composites of PMMA
and Surface-Modified Zincite Nanoparticles. Macromolecules 2007; 40:1089-1100.
27. Demir MM, Memesa M, Castignolles P, et al. PMMA/Zinc Oxide Nanocomposites
Prepared by In-Situ Bulk Polymerization. Macromol Rapid Commun 2006; 27: 763–
770.
28. Hung C-H, Whang W-T. Effect of surface stabilization of nanoparticles on
luminescent characteristics in ZnO/ poly(hydroxyethyl methacrylate) nanohybrid
films. J Mater Chem 2005; 15: 267–274.
29. Pre MD, Martucci A, Martin DJ, et al. Structural features, properties, and relaxations
of PMMA-ZnO nanocomposite. J Mater Sci 2015; 50: 2218-2228.
30. Zhang Y, Zhuang S, Xu X, et al. Transparent and UV-Shielding ZnO@PMMA
nanocomposite films. Opt Mater 2013; 36:169-172.
31. Zhang L, Li F, Chen Y, et al. Synthesis of transparent ZnO/PMMA nanocomposite
films through free-radical copolymerization of asymmetric zinc methacrylate acetate
and in-situ thermal decomposition. J Lumin 2011; 131:1701-1706.
32. Agrawal M, Gupta S, Zafeiropoulos NE, et al. Nano-Level Mixing of ZnO into
poly(methylmethacrylate). Macromol Chem. Phys 2010; 211: 1925-1932.
33. Salahuddin NA, El-Kemary M, Ibrahim EM. Synthesis and Characterization of ZnO
Nanoparticles via Precipitation Method: Effect of Annealing Temperature on Particle
Size. Nanosci. Nanotechnol 2015; 5: 82-88.
34. Salahuddin NA, El-Kemary M, Ibrahim EM. Synthesis and Characterization of ZnO
Nanotubes by Hydrothermal Method. Int. J. Sci. Res. Publ 2015; 5: 1-4.
35. Asar NV, Albayrak H, Korkmaz T, et al. Influence of various metal oxides on
mechanical and physical properties of heat-cured polymethyl methacrylate denture
base resins. J Adv Prosthodont 2013; 5: 241-247.

This article is protected by copyright. All rights reserved.


36. Radford D R, Challacombe SJ, Walter JD. Denture plaque and adherence of Candida
albicans to denture-base materials in vivo and in vitro. Crit Rev Oral Biol Med 1999;
10: 99-116.
Accepted Article
37. Dills SS, Olshan AM, Goldner S, et al. Comparison of the antimicrobial capability of
an abrasive paste and chemical-soak denture cleaners. J Prosthet Dent 1988; 60: 67-
70.
38. Li C, Du X, Lu W, et al. Luminescent single-crystal ZnO nanorods: controlled
synthesis through altering the solvent composition. Mater Lett 2012; 81: 229-231.
39. Motaung TE, Luyt AS, Bondioli F, et al. PMMA-titania nanocomposites: Properties
and thermal degradation behaviour. Polym Degrad Stab 2012; 97: 1325-1333.
40. Acevedo B, Martinez F, Olea AF. Synthesis, Thermal behavior, and aggregation in
aqueous solution of poly(methyl methacrylate)-B-Poly (2-hydroxyethyl
methacrylate). J Chil Chem Soc 2013; 58: 2038-2042.
41. Anzlovar A, Orel ZC, Zigon M. Mechanical properties of PMMA/ZnO
nanocomposites. European conference on composite Materials, Venice, Italy, 2012;
24-28 June.
42. Zappini G, Kammann A, Wachter W. Comparison of fracture test of denture base
materials. J Prosthet Dent. 2003: 90: 578-585.
43. Arnold AM, Vargas MA, Shaull KL, Laffoon JE, Qain F. Flexural and fatigue
strengths of denture base resin. J Prosthet Dent. 2008; 100: 47-51.
44. Abdulhamed AN, Mohamed AM. Evaluation of thermal conductivity of alumina
reinforced heat cure acrylic resin and some other properties. J Bagh Coll Dent. 2010;
22: 1-7.
45. Dhole RI, Srivatsa G, Shetty R, et al. Reinforcement of aluminum oxide filler on the
flexural strength of different types of denture base resins: an in vitro study. J Clin
Diagn Res. 2017; 11:ZC101-ZC104.
46. Arora N, Jain V, Chawla A, Mathur VP. Effect of addition of sapphire (aluminium
oxide) or silver fillers on the flexural strength thermal diffusivity and water sorption
of heat polymerized acrylic resins. Int J Prosthodont and Resto Dent. 2011; 1: 21-27.
47. Zuccari AG, Oshida Y, Moore BK. Reinforcement of acrylic resins for provisional
fixed restorations. Part 1: mechanical properties. Biomed Mater Eng. 1997; 7: 327-
343.
48. Schulze KA, Zaman AA, Soderholm KJ. Effect of filler fraction on strength, viscosity
and porosity of experimental compomer materials. J Dent. 2003; 31: 373-382.

This article is protected by copyright. All rights reserved.


49. Yu W, Wang X, Tang Q, et al. Reinforcement of denture base PMMA with ZrO2
nanotubes. J Mech Behav of Biomed Mater 2014; 32: 192-197.
50. Wang R, Tao J, Yu B, Dai L. Characterization of multiwalled carbon nanotube-
Accepted Article
polymethyl methacrylate composite resins as denture base materials. J Prosthet Dent.
2014; 111: 318-326.
51. Ayad NM, Badawi MF, Fatah AA. Effect of reinforcement of high-impact acrylic
resin with zirconia on some physical and mechanical properties. Rev Clin Pesq
Odontol Curitiba. 2008; 4: 145-151.
52. Yadav NS, Elkawash H. Flexural strength of denture base resin reinforced with
aluminum oxide and processed by different processing techniques. J Adv Dent Res
2011; 2(1): 33-36.
53. Sodagar A, Bahador A, Khalil S, et al. The effect of TiO2 and SiO2 nanoparticles on
flexural strength of poly (methyl methacrylate) acrylic resins. J Prosthodont Res
2013; 57:15-19.
54. Powers JM, Sakaguchi RL. Craig's restorative dental materials. 12th ed .St Louis:
Mosby Inc: 2006. P.60.
55. Suzuki S. In vitro wear of nano-composite denture teeth. J Prosthodont. 2004; 13:
238-243.
56. Alhareb AO, Akil HM, Ahmad ZA. Impact strength, fracture toughness and hardness
improvement of PMMA denture base through addition of nitrile rubber/ceramic
fillers. Saudi J Dent Res. 2017; 8: 26-34.

Figure and Table Captions

Figure 1: XRD patterns of (a) ZnO NPs, PMMA/ZnO nanocomposites loaded with different
amounts of ZnO NPs (PMMA/ZnO NPs) and (b) ZnO NTs, PMMA/ZnO
nanocomposites loaded with different amounts of ZnO NTs (PMMA/ ZnO NTs).

Figure 2: TGA thermograms of (a) PMMA, PMMA/ZnO nanocomposites loaded with


different amounts of ZnO NPs and (b) PMMA, PMMA/ZnO nanocomposites
loaded with different amounts of ZnO NTs.

This article is protected by copyright. All rights reserved.


Figure 3: SEM images of (a) ZnO NPs, (b-f) Low magnification SEM images of the fractured
surface of (b) pristine PMMA, (c) PMMA-0.2wt% ZnO NPs, (d) PMMA-0.4wt%
ZnO NPs, (e) PMMA-0.8wt% ZnO NPs and (f) PMMA-1wt% ZnO NPs.
Accepted Article
Figure 4: SEM images of (a) ZnO NTs, (b-f) Low magnification SEM images of the fractured
surface of (b) pristine PMMA, (c) PMMA-0.2wt% ZnO NTs, (d) PMMA-0.4wt%
ZnO NTs, (e) PMMA-0.8wt% ZnO NTs and (f) PMMA-1wt% ZnO NTs.

Figure 5: TEM images of (a) ZnO NPs and (b) ZnO NTs.

Figure 6: TEM images of PMMA/ZnO nanocomposites loaded with different amounts of


ZnO NPs (a) PMMA-0.2wt% ZnO NPs, (b) PMMA-0.4wt% ZnO NPs, (c)
PMMA-0.8wt% ZnO NPs and (d) PMMA-1wt% ZnO NPs.

Figure 7: TEM images of PMMA/ZnO nanocomposites loaded with different amounts of


ZnO NTs, (a) PMMA-0.2wt% ZnO NTs, (b) PMMA-0.4wt% ZnO NTs, (c)
PMMA-0.8wt% ZnO NTs and (d) PMMA-1wt% ZnO NTs.

Table 1: Thermal stability factors for PMMA/ZnO nanocomposites calculated from TGA and
DSC thermograms.

Table 2: Means, standard deviation and ANOVA results of Impact test.

Table 3: Means, standard deviation and ANOVA results of Flexural test.

Table 4: Hardness of PMMA/ZnO nanocomposites.

This article is protected by copyright. All rights reserved.


Table 1. Thermal stability factors for PMMA/ZnO nanocomposites calculated from TGA and
DSC thermograms.
Accepted Article
Sample code TIDT °C T10 °C T50 °C Tg °C

Pristine PMMA 241.0 256.5 303.0 119.4

PMMA – 0.2 wt% ZnO NPs 247.0 257.9 303.6 119.7

PMMA -0.4 wt% ZnO NPs 250.8 270.5 319.9 123.3

PMMA -0.8 wt% ZnO NPs 243.7 260.8 312.9 124.3

PMMA -1.0 wt% ZnO NPs 253.0 265.0 304.0 124.2

PMMA -0.2 wt% ZnO NTs 261.4 273.8 308.8 126.0

PMMA -0.4 wt% ZnO NTs 202.7 277.9 320.0 126.1

PMMA -0.8 wt% ZnO NTs 190.1 268.8 309.4 124.5

PMMA -1.0 wt% ZnO NTs 171.6 242.0 313.6 124.2

Table 2. Means, standard deviation and ANOVA results of Impact test

Impact strength kj/m2 ANOVA


Sample code
Range Mean ± SD f P-value
Pristine ( PMMA) 0.81 - 1.49 1.30 ± 0.28
PMMA – 0.2 wt% ZnO NPs 1.41 - 1.66 1.54 ± 0.10
PMMA – 0.4 wt% ZnO NPs 1.43 - 1.81 1.56 ± 0.15
PMMA – 0.8 wt% ZnO NPs 1.42 - 1.66 1.57 ± 0.11
PMMA – 1.0 wt% ZnO NPs 1.29 - 1.79 1.49 ± 0.21 9.878 <0.001*
PMMA – 0.2 wt% ZnO NTs 1.43 - 3.18 2.42* ± 0.84
PMMA – 0.4 wt% ZnO NTs 1.86 - 3.43 2.61* ± 0.57
PMMA – 0.8 wt% ZnO NTs 1.99 - 3.06 2.70* ± 0.42
PMMA –1.0 wt% ZnO NTs 0.90 - 1.57 1.37 ± 0.28

* Significant P˂0.05, High significant P˂0.001*

This article is protected by copyright. All rights reserved.


Table 3. Means, standard deviation and ANOVA results of Flexural test
Accepted Article
Flexural Strength (MPa) ANOVA
Sample code
Range Mean ± SD f P-value
Pristine ( PMMA) 82.0 - 85.3 84.000 ± 1.764
PMMA – 0.2 wt% ZnO NPs 92.5 - 97.5 94.372* ± 2.726
PMMA – 0.4 wt% ZnO NPs 92.7 - 98.9 95.780* ± 3.074
PMMA – 0.8 wt% ZnO NPs 85.8 - 87.7 86.786 ± 0.954
PMMA – 1.0 wt% ZnO NPs 84.8 - 86.0 85.267 ± 0.643 18.956 <0.001*
PMMA – 0.2 wt% ZnO NTs 92.3 - 94.5 93.473* ± 1.099
PMMA – 0.4 wt% ZnO NTs 92.0 - 94.9 93.772* ± 1.528
PMMA – 0.8 wt% ZnO NTs 93.0 - 96.0 94.196* ± 1.611
PMMA –1.0 wt% ZnO NTs 89.0 - 90.8 89.908* ± 0.908

* Significant P˂0.05, High significant P˂0.001*

Table 4. Hardness of PMMA/ZnO nanocomposites

Sample code Average hardness (GPa)

Pristine ( PMMA) 0.1448 ± 0.062

PMMA – 0.2 wt% ZnO NPs 0.2126 ± 0.650

PMMA – 0.4 wt% ZnO NPs 0.2346 ± 0.110

PMMA – 0.8 wt% ZnO NPs 0.2245 ± 0.087

PMMA –1 wt% ZnO NPs 0.2515 ± 0.070

PMMA – 0.2 wt% ZnO NTs 0.2642 ± 0.092

PMMA – 0.4 wt% ZnO NTs 0.2660 ± 0.109

PMMA – 0.8 wt% ZnO NTs 0.2796 ± 0.062

PMMA –1 wt% ZnO NTs 0.2823 ±0.125

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.


Accepted Article

This article is protected by copyright. All rights reserved.

You might also like