Next Article in Journal
Study of the Electrooxidation of a Zinc Concentrate
Next Article in Special Issue
On the Multi-Functional Behavior of Graphene-Based Nano-Reinforced Polymers
Previous Article in Journal
Nonlocality-Enabled Magnetic Free Optical Isolation in Hyperbolic Metamaterials
Previous Article in Special Issue
The Effect of Surfactant-Modified Montmorillonite on the Cross-Linking Efficiency of Polysiloxanes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Particle Size on the Mechanical Properties of TiO2–Epoxy Nanocomposites

1
Department of Thermal Systems, Korea Institute of Machinery and Materials, Daejeon 34103, Korea
2
Nanosafety Team, Safety Measurement Institute, Korea Research Institute of Standards and Science, Daejeon 34113, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to the work.
Materials 2021, 14(11), 2866; https://doi.org/10.3390/ma14112866
Submission received: 31 March 2021 / Revised: 30 April 2021 / Accepted: 24 May 2021 / Published: 27 May 2021
(This article belongs to the Special Issue Advanced Polymer Matrix Nanocomposite Materials)

Abstract

:
This study investigated the effects of the packing density and particle size distribution of TiO2 nanoparticles on the mechanical properties of TiO2–epoxy nanocomposites (NCs). The uniform dispersion and good interfacial bonding of TiO2 in the epoxy resin resulted in improved mechanical properties with the addition of nanoparticles. Reinforcement nano-TiO2 particles dispersed in deionized water produced by three different ultrasonic dispersion methods were used; the ultrasonication effects were then compared. The nano-TiO2 suspension was added at 0.5–5.0 wt.%, and the mechanical and thermal properties of TiO2–epoxy NCs were compared using a universal testing machine, scanning electron microscopy (SEM), Fourier-transform infrared spectroscopy (FT-IR), and differential scanning calorimetry (DSC). The tensile strength of the NCs was improved by the dispersion strengthening effect of the TiO2 nanoparticles, and focused sonication improved the tensile strength the most when nano-TiO2 suspensions with a particle size of 100 nm or smaller were used. Thus, the reinforcing effect of TiO2 nanoparticles on the epoxy resin was observed, and the nano-TiO2 suspension produced by focused sonication showed a more distinct reinforcing effect.

1. Introduction

Epoxy resin (EP) is an important industrial material that has many applications in the electronics, automobiles, and aerospace fields, owing to its high strength and stiffness, resistance to chemicals, low contraction during curing, excellent corrosion resistance, and thermal characteristics [1,2,3]. However, its other characteristics such as brittleness, poor resistance to crack propagation, and low wear resistance limit its applications [4,5,6]. To this end, several studies have investigated various fillers that can be added to improve the properties of the matrix.
Composite technology and nanotechnology are being extensively researched to solve these chronic problems that cannot be solved using single-layer materials. Nanocomposite technology has gained considerable research interest in all disciplines including engineering, physics, chemistry, and medicine; furthermore, it is believed to promote technical development in various fields such as information, electronics, materials, and energy [7]. Nanocomposites (NCs) reinforced with nanosized materials (1–100 nm size distribution) can be applied to a variety of materials including polymer-based NCs. Polymer nanocomposites (PNCs) are produced by combining lightweight, flexible polymers with low production costs as a matrix with inorganic nanoparticles that exhibits excellent mechanical and electromagnetic properties [8,9,10]. A PNC is a polymer matrix modified with reinforcement nanoparticles at a low ratio (<5.0 wt.%). Various nanoparticles, such as carbon nanotubes, titania (TiO2), silica, and alumina are used as the reinforcement particles [11,12,13,14]. These PNCs exhibit excellent mechanical, physical, and chemical properties; however, these properties strongly depend on the interfacial bond strength between the polymer matrix and reinforcement particles. The cohesion of nanoparticles and the high viscosity of polymers makes it considerably difficult to uniformly disperse nanoparticles of 5 wt.% or lower. Therefore, a mechanism that can uniformly disperse nanoparticles in a highly viscous liquid is highly desirable in academic circles [15,16,17].
The performance of nanoparticle-reinforced PNCs as a function of dispersion and interfacial interaction has been investigated in many studies; the results have suggested that good dispersion and strong interfacial interaction can effectively transfer the stress from the polymer matrix to the nanomaterials [18,19]. Various dispersion methods such as extrusion, ball milling, three-roll milling, solvent casting, and functionalization of nanoparticles have been employed. Among these methods, the mechanical mixing method is most commonly applied to create violent mixing, thereby generating turbulence on a large scale to nanoparticles. However, this method cannot prevent nanoparticle aggregation into the polymer matrix [20,21,22]. Therefore, ultrasonic dispersion methods are used instead of mechanical mixing methods to achieve high homogeneity [23,24]. Ultrasonic waves are classified into low- and high-frequency waves; the latter can decompose aggregates and separate particles accumulated on the microchannel surface by generating cavitation microbubbles. In contrast, the former form an ultrasonic sound field and lead to acoustic radiation and streaming effects, which facilitates biological and chemical treatments such as cell separation and fluid mixing [25,26,27,28,29]. Ultrasonic cavitation is an effective method for dispersing nanoparticles. As compression and rarefaction cycles are switched sequentially, high-intensity ultrasonic waves spread to the liquid polymer matrix; simultaneously, microbubbles of vacuum are created and collapsed in the matrix. When these microbubbles collapse, the local temperature changes to 105 Ks−1 because of the shockwave, and a pressure of several megapascals is generated. This energy can be used to disperse nanoparticles [30,31].
This study compared the dispersion of nano-TiO2 suspensions using three types of ultrasonic systems (bath sonication, probe sonication, and the novel focused sonication method we developed and reported in a previous paper [32]) to verify the nanoparticle dispersion efficiency of a proprietary focused sonication system. In detail, focused ultrasound uses cylindrical piezoelectric ceramics to dissolve agglomeration through the energy collected in the center. Generally, the surface charge of particles that prevent agglomeration is improved, to have a zeta potential value of about 40–50 mV in the dispersion process, which ensures long-term dispersion stability that does not re-aggregate for more than 1 year. We had confirmed that even when lyophilized particles were re-dispersed in water, they were dispersed to less than 100 nm without aggregation. Thus, the NCs based on EP were produced using the nano-TiO2 suspensions obtained by each ultrasonication method in this study; then, the physical properties of the TiO2–epoxy NCs were compared based on the added amount.

2. Experiment

2.1. Materials

TiO2 nanoparticles used in this study were type P25 (CAS: 13463-67-7) purchased from EVONIK Co. Ltd. (Essen, Germany); they had an average particle size of 25 nm and a density of 3.78 g/cm3. Figure 1 shows the shape of the TiO2 nanoparticles observed using transmission electron microscopy (TEM) (JEM-ARM200F, JEOL, Tokyo, Japan).
The polymer matrix was produced by mixing an EP and hardener supplied by Kukdo Chemical Co. Ltd. (Seoul, Korea). YD-128 (CAS: 25068-38-6) of the bisphenol-A type was used as the EP, and KBH-1085 (CAS: 25134-21-8) of nadic methyl anhydride (NMA) was used as the hardener. An accelerator (CAS: 103-83-3) of benzyl dimethyl amine (BDMA) was used as the hardening accelerator and was supplied by NEW SEOUL CHEMICAL (Seoul, Korea). YD-128, KBH-1085, and the accelerator were mixed at a manufacturer-recommended weight ratio of 100:60:2.5. Figure 2 shows the molecular structures of the EP, hardener, and hardening accelerator.

2.2. Sample Preparation

In this study, the PNC was produced by dispersing TiO2 nanoparticles (1.0 wt.%) in 100 mL of the deionized water and by adding it into a matrix. The nano-TiO2 suspension was dispersed using bath-type, probe-type, or focused-type sonication. SD-300H (Sungdong Ultrasonic Co., Seoul, Korea) and Branson SFX 550 (Emerson Co. Ltd., Saint Louis, MO, USA) were used for bath- and probe-type sonication, respectively. The results were compared with those of a proprietary focused sonication system [32].
Delivered sonic energy (DSE) (J/mL) is an important parameter in ultrasonication; it is the actual ultrasonic energy delivered to the sample. The mean particle size of TiO2 when DSE was applied to the same volume is compared in this study [33]. The DSE was calculated using
P = (dT/dt)MCp,
DSE = P × (t/V),
where P, dT/dt, M, Cp, t, and V denote the ultrasonic energy obtained by the calorimetric method [J/s], variation in temperature at time t [K/s], mass of the solution [g], heat capacity of water, time [s], and volume of the solution (mL), respectively.
The DSEs obtained using the three types of sonication methods are summarized in Table 1. To verify the particle size distribution of the TiO2 suspension dispersed at different ultrasonic energies, the mean particle size was measured after ultrasonication using dynamic light scattering (Zetasizer Nano ZS, Malvern Panalytical Ltd., Malvern, UK).
The particle size of TiO2 was 25 nm; however, it was agglomerated to approximately 500 nm [34]. After dispersion with the three sonication methods, the mean sizes of the nanoparticles were 96.4 nm, 155.3 nm, and 206.8 nm, as shown in Figure 3. Nano-TiO2 suspensions with a diameter of 100 nm were produced when the particles were dispersed by focused sonication. Furthermore, the intensity of the size distribution was higher than that of the others, which indicated that the particle sizes were more uniform because of the focused sonication.
Figure 4 shows an overview of the PNC manufacturing process. In this study, NCs with four different compositions were manufactured by adding nano-TiO2 suspensions (0.5 wt.%, 1.0 wt.%, 3.0 wt.%, and 5.0 wt.%) to an EP. It was difficult to disperse the EP with nanoparticles because of its high viscosity; therefore, the EP and TiO2 suspensions were dispersed using a stirring device and probe-type sonication. After mixing for 30 min at 700 rpm at 293.15 K using a mechanical agitator OS20-Pro (Korea Process Technology Co., Ltd., Seoul, Korea), they were dispersed for 180 min at 293.15 K using an ultrasonic dispersion system VC-505 (Sonics & Materials Inc. (Newtown, CT, USA). The residual gas was removed from the EP with a TiO2 suspension dispersed through a degassing process for 180 min in a 333.15 K vacuum oven. Then, the hardener and hardening accelerator were added and mixed for 5 min by hand stirring to blend and evenly distribute it with minimal bubbles; an additional degassing process was applied for 30 min in a 333.15 K vacuum oven. The TiO2–epoxy NCs were fabricated as tensile specimens (shape and dimensions as shown in Figure 5) hardened in two steps of 398.15 K for 60 min and 423.15 K for 180 min (Figure 4).

2.3. Characteristics of the Nanocomposites

The mechanical properties of 12 different batches of TiO2–epoxy NCs (namely, four different compositions and three different sonication methods) were measured using tensile tests and compared with those of pure epoxy resin samples. As shown in Figure 5, eight specimens per batch were molded into dog-bone-shaped tensile test specimens, which were tested in the tension mode of a universal testing machine M350-10CT (Testometric Co. Ltd., Rochdale, UK) at a crosshead speed of 2 mm/min. Then, the five median values were determined from each batch of eight samples, and the average of these five median values was calculated. In addition, the fracture surface was analyzed using SEM (SEM; SU5000, Hitachi, Tokyo, Japan) after the tensile test.
The thermal properties of one sample per batch were examined using DSC, which can analyze thermal properties such as glass transition temperature, crystallization temperature, and melting temperature using the calorie difference based on the temperature gradient between the reference material and the sample. Quantitative data can be obtained from the positions, shapes, and areas of the peaks in a DSC curve. In this study, the thermal properties of pure epoxy and TiO2–epoxy NCs were analyzed using the Perkin Elmer Diamond model (Perkin Elmer, Waltham, MA, USA) at a heating rate of 10 K/min from 273.15 K to 673.15 K in a nitrogen (N2) atmosphere.
Attenuated total reflectance Fourier-transform infrared (ATR-FTIR) spectroscopy was also used on one sample per batch to verify the bonding structure of the samples. ATR-FTIR spectroscopy is useful for measuring the surfaces of polymers such as sponges and adhesives, which are difficult to process. The IR beam from the interferometer is transmitted into a diamond crystal that results in an internal reflection at the boundary with the sample, because it has a greater refractive index than the sample in contact on both sides. Meanwhile, the IR spectrum that reflects the vibration and rotational motions of molecules are measured, and a qualitative analysis of the sample including molecular species is performed. The pure epoxy and TiO2–epoxy NCs were analyzed using ALPHA-P (Bruker, Billerica, MA, USA); the ATR-FTIR spectra were collected by scanning the sample surface in the range of 400–4000 cm−1.

3. Results and Discussion

3.1. Mechanical Properties of Nanocomposites

Dog-bone-shaped specimens were fabricated, and the tensile strengths were measured to examine the changes in the mechanical properties of the TiO2–epoxy NCs. Figure 6 shows the average tensile strengths of the NCs after sonication of the nano-TiO2 suspension. The average tensile strength of the composites with the nano-TiO2 suspension added was up to 15% higher than that of the pure EP. In particular, the tensile strengths of the NCs with 1.0 wt.% nano-TiO2 suspension processed by focused sonication was higher (92.9 MPa) compared to those processed using probe sonication (91.1 MPa) or bath sonication (89.7 MPa). Usually, bath sonication can handle large amounts at once, although it has been known to be less efficient at dispersion, and probe sonication has a strong irradiation energy, but it also has blind spots such as on the upper side of the probe. On the other hand, compared with the former sonication methods, it was determined that focused sonication had a positive effect on the TiO2 dispersion in the EP matrix. Due to the driving characteristics of focused sonication, it could continuously apply ultrasonic energy to the circulating object [32].
The tensile strength of the composites tended to decrease gradually when the added amount of nano-TiO2 suspension that passed through the ultrasonic dispersion system was 3.0 wt.% or higher. The decreased tensile strength of the NCs reinforced with nano-TiO2 suspension above a certain amount seems to be attributed to the following two problems: (1) The deionized water among the component materials of the nano-TiO2 suspension fabricated in this study had an adverse effect on the polymerization process of the EPs with hydroxyl and carboxyl groups. Therefore, the residual water in the TiO2–epoxy NCs needed to be checked via thermal analysis and the improvements should be identified [35,36]; (2) The interference of the epoxy polymerization and the reduced effect of particle reinforcement was attributed to the agglomeration of TiO2 nanoparticles. Adding a large quantity of nanoparticles into a highly viscous EP causes agglomeration because of the attraction between particles. Accordingly, the agglomerated nanoparticles serve as impurities in the resin; therefore, they decrease the mechanical properties [37,38,39]. In this study, when PNCs were produced by adding a nano-TiO2 suspension produced by the same sonication method in the range of 0.5–5.0 wt.%, the addition of 1.0 wt.% was found to be the optimal amount.

3.2. Microstructural and Morphological Analysis

Microstructures at 200× magnification were observed using SEM to examine the failure behavior of the TiO2–epoxy NCs. Figure 7 shows the fracture surface images after the tensile testing of pure EP and epoxy NCs reinforced with TiO2 using different sonication methods. Pure EP was very clear: from Figure 7a, the fracture surface of neat epoxy showed a very smooth pattern where cracks propagated arbitrarily without restrain, indicating the brittle fracture of neat epoxy without resistance to crack initiation and transmission. Compared to the fracture surface of the pure EP, the fracture surface of NCs with TiO2 nanoparticles added in Figure 7b–d showed a rough cliff-like pattern (blue arrow: flat cross-section, and red arrow: rough cross-section). In general, more fracture energy is dissipated for greater roughness. These observations reveal the generation of a rough surface by the deflection of propagating crack fronts on the loading of TiO2 in the EP matrix. In other words, the cracks progressed relatively rapidly without high resistance at the fracture surface of the pure epoxy specimen; furthermore, the crack resistance of the TiO2–epoxy NCs improved with the addition of nanoparticles, which acted as a factor that increased the tensile strength [40,41].
Figure 8 shows the SEM images with 500× magnification for the fracture surfaces of the nanocomposite reinforced with nano-TiO2 suspension dispersed by focused sonication. The characteristic brittleness and low fracture toughness of the pure epoxy are a product of their high cross-link densities, which resulted in the poor absorption of energy during the fracture. These factors frequently lead to mirror-like fracture surfaces, as shown in Figure 8a. Changes in the fracture surface caused by the added nanoparticles were examined. Compared to that for the pure epoxy, the fracture surface of NCs contained greater roughness, as shown in Figure 8b–e. The coarse multiplane features on the TiO2–epoxy NCs fracture surface suggest that the TiO2 nanoparticles induced the deflection of propagating crack fronts. Furthermore, NCs with 3.0 wt.% and 5.0 wt.% TiO2 showed a decreased roughness compared to the NCs with 1.0 wt.% nano-TiO2 suspension. The PNCs are effective only if the nanoparticles are well dispersed in the surrounding polymer matrix. In addition, the agglomeration of nanoparticles should not be overlooked, because nanoparticles provide a very specific surface area in the polymer matrix [10]. In other words, over-addition of nanoparticles causes the agglomeration of particles, which reduces the particle reinforcement effect. Consequently, the tensile strength could be decreased, as shown in Figure 6 above.

3.3. FT-IR Spectroscopy

In this study, PNCs were produced using nano-TiO2 suspensions processed by different sonication methods but with the same content, and their physical properties were compared. As listed in Table 1, the particle size varied according to the sonication method used. The focused sonication showed an average particle size of 96.4 nm, which is smaller than those of other sonication methods. In general, crosslinking density and hydrogen bonding density are significantly affected by the content and size of the nanoparticles and the packing density in the polymer matrix. Therefore, singularities based on the bonding type and added amount of TiO2 in the polymer nanocomposite network were examined by FT-IR analysis.
Figure 9 shows the FT-IR spectra of pure epoxy and TiO2–epoxy NCs in the range of 400–4000 cm−1. The O–H bond could appear at 3500–3000 cm−1, and H–O–H bonds are normally detected at 1700–1200 cm−1. These bonds could occur in the presence of water, and the results of FT-IR analysis are shown in this study. Therefore, it is assumed that there was no residual water. The peaks at 2850–2950 cm−1 indicated the C–H group, whereas those at 1730–1740 cm−1 and 1510 cm−1 indicated the C=O and aromatic groups, respectively. The TiO2–epoxy NCs processed by focused sonication showed peak intensities similar to those of pure epoxy. Overall, the FT-IR patterns of all samples including the pure EP were almost the same, which suggests that there was no chemical bonding between TiO2 and the EP.

3.4. Thermal Property of Nanocomposites

Nanoparticles play a key role in improving the thermodynamic properties of EPs. Among them, TiO2 is widely used in industries because of its diverse functions and synthesis methods. The state of materials can be analyzed using DSC based on temperature changes with high speed and precision, and it is used in research investigating the physical properties of polymers. In particular, DSC is widely used for examining polymeric materials to determine their thermal transitions. Important thermal transitions include the glass transition temperature (Tg), crystallization temperature (Tc), and melting temperature (Tm). The epoxy resin of bisphenol-A-type, an amorphous polymer, was used in this study; thus, we expected to obtain acceptable Tg and Tm values. The Tg is a very important thermal property in polymeric systems. The unit of polymers is a certain segment in the molecular chain, and the thermal motion by the rotation of this segment is called micro-Brownian motion. The temperature at which micro-Brownian motion starts is called the glass transition temperature, and it is also referred to as the second transition temperature, which is a unique property of polymers.
Figure 10 shows the DSC curves of the pure epoxy and TiO2–epoxy NCs. A glass transition point was observed at 350–450 K, wherein the endothermic reaction of pure EP appeared distinctly. The glass transition temperature of the pure EP is 355 K. Furthermore, the glass transition temperatures of the NCs processed by focused sonication, probe sonication, and bath sonication were 400.0 K, 375.7 K, and 360.0 K, respectively. This confirmed that the thermal properties of the TiO2–epoxy NCs were improved compared to those of the pure EP [42,43]. The NCs in which the nano-TiO2 suspensions were dispersed by focused sonication showed the best thermal properties, and therefore, thermal stability could be expected.
On the other hand, there were no other reactions than phase transitions of the nanocomposites. We fabricated the nanocomposites using an aqueous solution-based TiO2 suspension in this study, and we have confirmed that the residual moisture could have an adverse effect on the nanocomposites. However, compared to pure epoxy resin, the specificity peak of nanocomposites was not shown in the DSC analysis.

4. Conclusions

In this study, the effect of TiO2–epoxy size on the mechanical properties of particle NCs was investigated. In addition, nano-TiO2 suspensions were dispersed separately using the existing bath sonication and probe sonication methods to examine the dispersion efficiency of a proprietary focused sonication dispersion system. A nano-TiO2 suspension with a mean particle size of 96.4 nm was obtained using the focused sonication dispersion system. TiO2 particle-reinforced epoxy NCs were fabricated, and their mechanical properties were compared. The following conclusions could be drawn based on the results.
  • The tensile strength of the TiO2–epoxy NCs was improved compared to that of the pure EP. The highest tensile strength was observed when 1.0 wt.% of TiO2 suspension (mean particle size 100 nm or lower) dispersed by focused sonication was added. However, the tensile strength of more than 3.0 wt.% TiO2-reinforced epoxy NCs decreased regardless of particle size;
  • The SEM analysis of the fracture surface indicated that the crack resistance of the TiO2–epoxy NCs was greater than that of the pure EP; this influences the tensile strength;
  • The FT-IR spectra of the pure EP and TiO2–epoxy NCs were almost the same, which suggests that chemical bonding with TiO2 did not occur. The TiO2–epoxy NCs showed improved thermal properties compared to pure EP;
  • The glass transition temperature of the TiO2-reinforced epoxy NCs with a particle size of 100 nm or less (dispersed by focused sonication) was the highest at 400 K; this implies that the performance improvement of the nanocomposite was affected by the size of the nanoparticles;
  • Both the tensile strength and thermal properties improved when a suspension with a particle size of 100 nm or less was used; this was confirmed to be greatly influenced by the nanoparticle dispersion method.
Continuous research is needed to further optimize and standardize the focused ultrasound method used in this study so that it can be applied to various NCs.

Author Contributions

Conceptualization, Y.-M.C. and S.-A.H.; Methodology, Y.-M.C.; Software, Y.-M.C. and S.-A.H.; Validation, Y.-M.C. and S.-A.H.; Formal analysis, Y.-M.C. and S.-A.H.; Investigation, Y.-M.C. and S.-A.H.; Resources, S.-A.H.; Data curation, Y.-M.C. and S.-A.H.; Writing—original draft preparation, Y.-M.C., S.-A.H. and Y.-B.H.; Writing—review and editing, Y.-M.C., S.-A.H., T.G.L. and Y.-B.H.; Visualization, Y.-M.C. and S.-A.H.; Supervision, S.-A.H., T.G.L. and Y.-B.H.; Project administration, S.-A.H., T.G.L. and Y.-B.H.; Funding acquisition, S.-A.H., T.G.L. and Y.-B.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Nano-Material Technology Development Program of the National Research Foundation (NRF) funded by the Ministry of Science and ICT (grant number 2016M3A7B6908929), and the Development of Measurement Standards and Technology for Biomaterials and Medical Convergence funded by the Korea Research Institute of Standards and Science (KRISS–2021–GP2021-0004).

Data Availability Statement

Data available in a publicly accessible repository.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ashrafi, M.; Hamadanian, M.; Ghasemi, A.R.; Kashi, F.J. Improvement mechanical and antibacterial properties of epoxy by polyethylene glycol and Ag/CuO nanoparticles. Polym. Compos. 2019, 40, 3393–3401. [Google Scholar] [CrossRef]
  2. Heng, Z.; Chen, Y.; Zou, H.; Liang, M. Simultaneously enhanced tensile strength and fracture toughness of epoxy resins by a poly (ethylene oxide)-block-carboxyl terminated butadiene-acrylonitrile rubber dilock copolymer. RSC Adv. 2015, 5, 42362–42368. [Google Scholar] [CrossRef]
  3. Flick, E.W. Epoxy Resins, Curing Agents, Compounds, and Modifiers: An Industrial Guide; William Andrew: Norwich, NY, USA, 2012. [Google Scholar]
  4. Shen, X.J.; Pei, X.Q.; Liu, Y.; Fu, S.Y. Tribological performance of carbon nanotube–graphene oxide hybrid/epoxy composites. Compos. Part B Eng. 2014, 57, 120–125. [Google Scholar] [CrossRef]
  5. Jiang, T.; Kuila, T.; Kim, N.H.; Ku, B.C.; Lee, J.H. Enhanced mechanical properties of silanized silica nanoparticle attached graphene oxide/epoxy composites. Compos. Sci. Technol. 2013, 79, 115–125. [Google Scholar] [CrossRef]
  6. Deng, S.H.; Zhou, X.D.; Zhu, M.Q.; Fan, C.J.; Lin, Q.F. Interfacial toughening and consequent improvement in fracture toughness of carbon fiber reinforced epoxy resin composites: Induced by diblock copolymers. Exp. Polym. Lett. 2013, 7, 925–935. [Google Scholar] [CrossRef]
  7. Rajangam, K.; Amuthameena, S.; Thangavel, S.; Sanjanadevi, V.S.; Balraj, B. Synthesis and characterisation of Ag incorporated TiO2 nanomaterials for supercapacitor applications. J. Mol. Struct. 2020, 1219, 128661. [Google Scholar] [CrossRef]
  8. Green, K.J.; Dean, D.R.; Vaidya, U.K.; Nyairo, E. Multiscale fiber reinforced composites based on a carbon nanofiber/epoxy nanophased polymer matrix: Synthesis, mechanical, and thermomechanical behavior. Compos. Part A Appl. Sci. Manuf. 2009, 40, 1470–1475. [Google Scholar] [CrossRef]
  9. Pinto, D.; Bernardo, L.; Amaro, A.; Lopes, S. Mechanical properties of epoxy nanocomposites using titanium dioxide as reinforcement–a review. Const Build. Mater. 2015, 95, 506–524. [Google Scholar] [CrossRef]
  10. Chatterjee, A.; Islam, M.S. Fabrication and characterization of TiO2–epoxy nanocomposite. Mater. Sci. Eng. A 2008, 487, 574–585. [Google Scholar] [CrossRef]
  11. Raja, A.; Rajasekaran, P.; Selvakumar, K.; Raman, R.G.; Swaminathan, M. Visible active TiO2-CdS-rGO ternary nanocomposite for enhanced photodecomposition of methylene blue. Mater. Today Proc. 2020, 29, 1125–1128. [Google Scholar] [CrossRef]
  12. Bayan, E.M.; Lupeiko, T.G.; Pustovaya, L.E.; Volkova, M.G.; Butova, V.V.; Guda, A.A. Zn–F co-doped TiO2 nanomaterials: Synthesis, structure and photocatalytic activity. J. Alloy Compd. 2020, 822, 153662. [Google Scholar] [CrossRef]
  13. Takari, A.; Ghasemi, A.R.; Hamadanian, M.; Sarafrazi, M.; Najafidoust, A. Molecular dynamics simulation and thermo-mechanical characterization for optimization of three-phase epoxy/TiO2/SiO2 nano-composites. Polym. Test. 2021, 93, 106890. [Google Scholar] [CrossRef]
  14. Kumar, K.; Ghosh, P.K.; Kumar, A. Improving mechanical and thermal properties of TiO2-epoxy nanocomposite. Compos. Part B Eng. 2016, 97, 353–360. [Google Scholar] [CrossRef]
  15. Barczewski, M.; Sałasińska, K.; Szulc, J. Application of sunflower husk, hazelnut shell and walnut shell as waste agricultural fillers for epoxy-based composites: A study into mechanical behavior related to structural and rheological properties. Polym. Test. 2019, 75, 1–11. [Google Scholar] [CrossRef]
  16. Alian, A.R.; Kundalwal, S.I.; Meguid, S.A. Multiscale modeling of carbon nanotube epoxy composites. Polymer 2015, 70, 149–160. [Google Scholar] [CrossRef]
  17. Abdalla, M.; Dean, D.; Robinson, P.; Nyairo, E. Cure behavior of epoxy/MWCNT nanocomposites: The effect of nanotube surface modification. Polymer 2008, 49, 3310–3317. [Google Scholar] [CrossRef]
  18. Kumar, A.; Kumar, K.; Ghosh, P.K.; Yadav, K.L. MWCNT/TiO2 hybrid nano filler toward high-performance epoxy composite. Ultrason. Sonochem. 2018, 41, 37–46. [Google Scholar] [CrossRef]
  19. Kumar, A.; Ghosh, P.K.; Yadav, K.L.; Kumar, K. Thermo-mechanical and anti-corrosive properties of MWCNT/epoxy nanocomposite fabricated by innovative dispersion technique. Compos. Part B Eng. 2017, 113, 291–299. [Google Scholar] [CrossRef]
  20. Mallakpour, S.; Abdolmaleki, A.; Azimi, F. Ultrasonic-assisted biosurface modification of multi-walled carbon nanotubes with Thiamine and its influence on the properties of PVC/Tm-MWCNTs nanocomposite films. Ultrason. Sonochem. 2017, 39, 589–596. [Google Scholar] [CrossRef]
  21. Kudo, T.; Sekiguchi, K.; Sankoda, K.; Namiki, N.; Nii, S. Effect of ultrasonic frequency on size distributions of nanosized mist generated by ultrasonic atomization. Ultrason. Sonochem. 2017, 37, 16–22. [Google Scholar] [CrossRef]
  22. Maktedar, S.S.; Mehetre, S.S.; Avashthi, G.; Singh, M. In situ sonochemical reduction and direct functionalization of graphene oxide: A robust approach with thermal and biomedical applications. Ultrason. Sonochem. 2017, 34, 67–77. [Google Scholar] [CrossRef] [PubMed]
  23. Mallakpour, S. Ultrasonic-assisted fabrication of starch/MWCNT-glucose nanocomposites for drug delivery. Ultrason. Sonochem. 2018, 40, 402–409. [Google Scholar] [CrossRef]
  24. Mallakpour, S.; Rashidimoghadam, S. Application of ultrasonic irradiation as a benign method for production of glycerol plasticized-starch/ascorbic acid functionalized MWCNTs nanocomposites: Investigation of methylene blue adsorption and electrical properties. Ultrason. Sonochem. 2018, 40, 419–432. [Google Scholar] [CrossRef]
  25. Dong, Z.; Delacour, C.; Mc Carogher, K.; Udepurkar, A.P.; Kuhn, S. Continuous ultrasonic reactors: Design, mechanism and application. Materials 2020, 13, 344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Trujillo, F.J.; Juliano, P.; Barbosa-Cánovas, G.; Knoerzer, K. Separation of suspensions and emulsions via ultrasonic standing waves–A review. Ultrason. Sonochem. 2014, 21, 2151–2164. [Google Scholar] [CrossRef]
  27. Shields IV, C.W.; Reyes, C.D.; López, G.P. Microfluidic cell sorting: A review of the advances in the separation of cells from debulking to rare cell isolation. Lab Chip 2015, 15, 1230–1249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Evander, M.; Nilsson, J. Acoustofluidics 20: Applications in acoustic trapping. Lab. Chip 2012, 12, 4667–4676. [Google Scholar] [CrossRef]
  29. Gelin, P.; Van Lindt, J.; Bratek-Skicki, A.; Stroobants, S.; Krzek, M.; Ziemecka, I.; Tompa, P.; De Malsche, W.; Maes, D. Focusing of microcrystals and liquid condensates in acoustofluidics. Crystals 2019, 9, 120. [Google Scholar] [CrossRef] [Green Version]
  30. Muthoosamy, K.; Manickam, S. State of the art and recent advances in the ultrasound-assisted synthesis, exfoliation and functionalization of graphene derivatives. Ultrason. Sonochem. 2017, 39, 478–493. [Google Scholar] [CrossRef] [PubMed]
  31. Ghosh, P.K.; Patel, A.; Kumar, K. Adhesive joining of copper using nano-filler composite adhesive. Polymer 2016, 87, 159–169. [Google Scholar] [CrossRef]
  32. Hwangbo, S.; Kwak, M.; Kim, J.; Lee, T.G. Novel surfactant-free water dispersion technique of TiO2 NPs using focused ultrasound system. Nanomaterials 2021, 11, 427. [Google Scholar] [CrossRef]
  33. DeLoid, G.M.; Cohen, J.M.; Pyrgiotakis, G.; Demokritou, P. Preparation, characterization, and in vitro dosimetry of dispersed, engineered nanomaterials. Nat. Protoc. 2017, 12, 355–371. [Google Scholar] [CrossRef]
  34. Sato, K.; Li, J.G.; Kamiya, H.; Ishigaki, T. Ultrasonic dispersion of TiO2 nanoparticles in aqueous suspension. J. Am. Ceram. Soc. 2008, 91, 2481–2487. [Google Scholar] [CrossRef]
  35. Ahmad, Z.; Ansell, M.P.; Smedley, D. Moisture absorption characteristics of epoxy based adhesive reinforced with CTBN and ceramic particles for bonded-in timber connection: Fickian or non-Fickian behaviour. IOP Conf Ser. Mater. Sci. Eng. 2011, 17, 012011. [Google Scholar] [CrossRef]
  36. Antoon, M.K.; Koenig, J.L. Irreversible effects of moisture on the epoxy matrix in glass-reinforced composites. J. Polym. Sci. Polym. Phys. 1981, 19, 197–212. [Google Scholar] [CrossRef]
  37. Hu, H.; Onyebueke, L.; Abatan, A. Characterizing and modeling mechanical properties of nanocomposites-review and evaluation. J. Miner. Mater. Charact. Eng. 2010, 9, 275. [Google Scholar] [CrossRef]
  38. Fisher, F.T.; Brinson, L.C. Nanomechanics of nanoreinforced polymers. In Handbook of Theoretical and Computational Nanoscience; Volume 9: Nanocomposites, Nano-Assemblies, and Nanosurfaces; American Scientific Publishers: Karlsruhe, Germany, 2006; pp. 253–360. [Google Scholar]
  39. Moniruzzaman, M.; Winey, K.I. Polymer nanocomposites containing carbon nanotubes. Macromolecules 2006, 39, 5194–5205. [Google Scholar] [CrossRef]
  40. Bortz, D.R.; Heras, E.G.; Martin-Gullon, I. Impressive fatigue life and fracture toughness improvements in graphene oxide/epoxy composites. Macromolecules 2012, 45, 238–245. [Google Scholar] [CrossRef]
  41. Rafiee, M.A.; Rafiee, J.; Wang, Z.; Song, H.; Yu, Z.Z.; Koratkar, N. Enhanced mechanical properties of nanocomposites at low graphene content. ACS Nano 2009, 3, 3884–3890. [Google Scholar] [CrossRef]
  42. Kumar, R.V.; Koltypin, Y.; Palchik, O.; Gedanken, A. Preparation and characterization of nickel–polystyrene nanocomposite by ultrasound irradiation. J. Appl. Polym. Sci. 2002, 86, 160–165. [Google Scholar] [CrossRef]
  43. Wu, C.L.; Zhang, M.Q.; Rong, M.Z.; Friedrich, K. Silica nanoparticles filled polypropylene: Effects of particle surface treatment, matrix ductility and particle species on mechanical performance of the composites. Compos. Sci. Technol. 2005, 65, 635–645. [Google Scholar] [CrossRef]
Figure 1. TEM image of the P25 purchased from EVONIK Co., Ltd.
Figure 1. TEM image of the P25 purchased from EVONIK Co., Ltd.
Materials 14 02866 g001
Figure 2. Chemical structures of (a) epoxy resin, (b) hardener, and (c) hardening accelerator.
Figure 2. Chemical structures of (a) epoxy resin, (b) hardener, and (c) hardening accelerator.
Materials 14 02866 g002
Figure 3. Size distribution of TiO2 nanoparticles in deionized water based on sonication type.
Figure 3. Size distribution of TiO2 nanoparticles in deionized water based on sonication type.
Materials 14 02866 g003
Figure 4. Flowchart of TiO2–epoxy nanocomposite preparation.
Figure 4. Flowchart of TiO2–epoxy nanocomposite preparation.
Materials 14 02866 g004
Figure 5. Dimensions for the tensile test specimen.
Figure 5. Dimensions for the tensile test specimen.
Materials 14 02866 g005
Figure 6. Tensile strength of the TiO2–epoxy nanocomposites in accordance with the sonication method.
Figure 6. Tensile strength of the TiO2–epoxy nanocomposites in accordance with the sonication method.
Materials 14 02866 g006
Figure 7. SEM images of (a) the pure epoxy resin and the 1.0 wt.% TiO2–epoxy nanocomposites with (b) bath sonication, (c) probe sonication, and (d) focused sonication.
Figure 7. SEM images of (a) the pure epoxy resin and the 1.0 wt.% TiO2–epoxy nanocomposites with (b) bath sonication, (c) probe sonication, and (d) focused sonication.
Materials 14 02866 g007
Figure 8. SEM images of (a) pure epoxy resin, (b) 0.5 wt.%, (c) 1.0 wt.%, (d) 3.0 wt.%, and (e) 5.0 wt.% TiO2–epoxy nanocomposites with focused sonication.
Figure 8. SEM images of (a) pure epoxy resin, (b) 0.5 wt.%, (c) 1.0 wt.%, (d) 3.0 wt.%, and (e) 5.0 wt.% TiO2–epoxy nanocomposites with focused sonication.
Materials 14 02866 g008
Figure 9. FT-IR spectra of the TiO2 powder, pure epoxy resin, and the 1.0 wt.% TiO2–epoxy nanocomposites.
Figure 9. FT-IR spectra of the TiO2 powder, pure epoxy resin, and the 1.0 wt.% TiO2–epoxy nanocomposites.
Materials 14 02866 g009
Figure 10. DSC curve of the 1.0 wt.% TiO2–epoxy nanocomposites in accordance with the sonication method conducted within the range of 200–700 K.
Figure 10. DSC curve of the 1.0 wt.% TiO2–epoxy nanocomposites in accordance with the sonication method conducted within the range of 200–700 K.
Materials 14 02866 g010
Table 1. Specifications of the three types of sonication methods.
Table 1. Specifications of the three types of sonication methods.
SpecificationBath SonicationProbe SonicationFocused Sonication
Frequency40 kHz20 kHz400 kHz
Solution volume100 mL100 mL100 mL
Irradiation time32 h 12 min1 h 47 min2 h
Ultrasonic energy (P)1.7 J/s29.6 J/s27.0 J/s
Delivered sonic energy (DSE)194.7 J/mL194.7 J/mL194.7 J/mL
Median particle size206.8 nm155.3 nm96.4 nm
Range of particle size distribution122–5560 nm60–531 nm37.8–255 nm
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Choi, Y.-M.; Hwangbo, S.-A.; Lee, T.G.; Ham, Y.-B. Effect of Particle Size on the Mechanical Properties of TiO2–Epoxy Nanocomposites. Materials 2021, 14, 2866. https://doi.org/10.3390/ma14112866

AMA Style

Choi Y-M, Hwangbo S-A, Lee TG, Ham Y-B. Effect of Particle Size on the Mechanical Properties of TiO2–Epoxy Nanocomposites. Materials. 2021; 14(11):2866. https://doi.org/10.3390/ma14112866

Chicago/Turabian Style

Choi, Young-Min, Seon-Ae Hwangbo, Tae Geol Lee, and Young-Bog Ham. 2021. "Effect of Particle Size on the Mechanical Properties of TiO2–Epoxy Nanocomposites" Materials 14, no. 11: 2866. https://doi.org/10.3390/ma14112866

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop