Next Article in Journal
The Effect of Stereocomplex Polylactide Particles on the Stereocomplexation of High Molecular Weight Polylactide Blends
Next Article in Special Issue
Poly(Alkylene 2,5-Thiophenedicarboxylate) Polyesters: A New Class of Bio-Based High-Performance Polymers for Sustainable Packaging
Previous Article in Journal
Thermal and Chemical Characterization of Kenaf Fiber (Hibiscus cannabinus) Reinforced Epoxy Matrix Composites
Previous Article in Special Issue
A Review of Bioplastics and Their Adoption in the Circular Economy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synergic Effect of TiO2 Filler on the Mechanical Properties of Polymer Nanocomposites

1
Renewable Energy Systems and Recycling Research Center, Transilvania University of Brasov, 500036 Brasov, Romania
2
Product Design, Mechatronics and Environment Department, Transilvania University of Brasov, 500036 Brasov, Romania
*
Author to whom correspondence should be addressed.
Polymers 2021, 13(12), 2017; https://doi.org/10.3390/polym13122017
Submission received: 9 May 2021 / Revised: 11 June 2021 / Accepted: 18 June 2021 / Published: 20 June 2021
(This article belongs to the Special Issue Advances in Sustainable Polymeric Materials)

Abstract

:
Nanocomposites with polymer matrix offer excellent opportunities to explore new functionalities beyond those of conventional materials. TiO2, as a reinforcement agent in polymeric nanocomposites, is a viable strategy that significantly enhanced their mechanical properties. The size of the filler plays an essential role in determining the mechanical properties of the nanocomposite. A defining feature of polymer nanocomposites is that the small size of the fillers leads to an increase in the interfacial area compared to traditional composites. The interfacial area generates a significant volume fraction of interfacial polymer, with properties different from the bulk polymer even at low loadings of the nanofiller. This review aims to provide specific guidelines on the correlations between the structures of TiO2 nanocomposites with polymeric matrix and their mechanical properties. The correlations will be established and explained based on interfaces realized between the polymer matrix and inorganic filler. The paper focuses on the influence of the composition parameters (type of polymeric matrix, TiO2 filler with surface modified/unmodified, additives) and technological parameters (processing methods, temperature, time, pressure) on the mechanical strength of TiO2 nanocomposites with the polymeric matrix.

1. Introduction

Polymer nanocomposites represent a new class of composite materials that generally exhibit better properties than traditional microcomposites, in terms of mechanical properties, thermal and dimensional stability, fire and chemical resistance, optical and electrical properties, etc. Polymer nanocomposites with inorganic fillers attracted significant an attention due to their unique properties and their numerous applications in modern technology. The properties of polymer nanocomposites are mostly a simple combination of incorporated inorganic nanoparticles and polymeric matrix.
Polymeric materials can be used as matrices in nanocomposites due to their good thermal stability, environmental resistance (durability), and electrical, chemical and mechanical properties [1]. However, it is well known that some polymers (e.g., epoxy resin) are highly brittle. This disadvantage limits the application of these polymers in products that require high impact and fracture strength. Inorganic filler added into polymer matrix improved the mechanical performance of the polymeric nanocomposites. Nanofillers have large surface areas, making them chemically active, and making them interact more easily with the matrix [2]. There are many methods to reinforce polymers with rigid fillers to reduce the cost of production, alleviate some of the polymers limitations and expand their applications [3]. How fillers influence the characteristics of these polymers depends on the polymer nature and the proportion of the filler. Fillers are used to modify many properties of polymers, such as mechanical [4] (flexural strength, tensile modulus, tensile strength, fracture toughness and impact energy), thermal, electrical, and magnetic properties [5,6].
The polymer mass, chemical structure, semi-crystallinity, chemical solubility, and thermal stability, and the nanoparticle surface area, chemical structure, and dispersion are essential for obtaining polymer nanocomposites and understanding their behavior. There are several methods to obtain polymeric nanocomposites, such as modified emulsion polymerization [7], in situ polymerizations [8,9], via direct blending (mechanical mixing) [10,11], solution dispersion [12,13,14], the sol-gel method and melt compounding [15,16], selective laser sintering process [17], and melt extrusion and injection molding [18,19]. Each process is specific, but the final morphology of the nanocomposites plays an important role. The morphology depends not only on the method of obtaining the nanocomposites, but also on the polymer–nanoparticle interactions that promote good dispersion and distribution of the nanoparticles in the polymer matrix [20,21].
Polymer nanocomposites have superior mechanical and physical properties over host polymers, due to the large interfacial area between the polymer matrix and nano-fillers.
Among the different fillers used, such as clays, silicas, nanotubes, inorganics, etc., titanium dioxide (TiO2) play a special role in polymeric matrices, to synthesize high-performance and malleable polymer networks (e.g., improving viscosity, obtaining filaments for 3D printing) [22,23,24]. TiO2 is found in many applications due to its good photocatalytic properties, hence it is used in antiseptic and antibacterial compositions, degrading organic contaminants and germs, as a UV-resistant material; this is due to its chemical inertness properties, non-toxicity, low cost, high refractive index, and other advantageous surface properties. In these applications, TiO2 is used as a component of various types of nanocomposite materials with special properties, which open up opportunities in the following various fields of applicability: in the production of pharmaceuticals, cosmetics or paints [25], drug delivery systems with controlled release [26], solar cell [27], chemical sensing, luminescent material, and photocatalyst [28]. For example, as materials for obtaining membranes for integration in environmental applications, including water treatment or reducing humidification [29,30]. Polymer nanocomposites find applications in the development of optical and electronic devices, sensors, and bio-sensors [31,32].
The incorporation of TiO2 nanoparticles into different types of the polymeric matrix could produce synergistic effects. Studies have been performed on the TiO2 nanoparticle effect on several properties of polymeric composite, mainly to figure out whether the application of nanoparticles can enhance the mechanical performance of polymeric composites for applications in various fields.
This paper comprehensively reviews some essential aspects, such as the processing, characterization, and mechanical properties of various nanocomposites with a polymeric matrix and TiO2 fillers.

2. Polymeric Matrix

2.1. Matrix

The main component in the nanocomposite of the polymer matrix is the polymer itself. There are many varieties of polymers used in the preparation of polymeric matrix nanocomposites. These polymers can be thermoplastics, thermosetting, elastomers, natural, and biodegradable polymers. The choice of filler depends on the nature of the polymer, thus obtaining materials with the following specific properties: mechanical, electrical, magnetic, optical biocompatibility, chemical stability, and functionalization. Thermosetting polymer nanocomposites are usually the most common nanocomposites. They are used in many applications, but recently thermoplastic polymer nanocomposites have attracted much of the research interest in industry and academia. The properties of polymers depend mainly on the polymer structure, which in turn depends on the chemical composition, surface morphology, and processing parameters. Polymers are a source of a wide variety of low-priced raw materials, which offer many advantages, such as the following [10]: low specific weight, high material stability against corrosion, good electrical and thermal insulation, ease of shaping and economical mass production, and attractive optical properties. However, they have some deficiencies in strength and stiffness. Fillers are integrated into polymer materials to make up for those deficiencies. These polymers can be epoxy resins, polyester fibreglass resins systems, PURs, PIs, urea, etc.
Theoretically, the associations that can be made between different polymers and the wide range of fillers are infinite. In practice, however, although numerous, the polymer–filler associations are limited. Among the thermoplastic polymers for the processing of which fillers can be introduced, the most important are as follows: polyolefin, polyamides, ABS polymers, polyesters, polycarbonates, and PVC. Elastomers are flexible polymers that comprise a low crosslink density and generally have low Young’s modulus, and by incorporating the fillers, these matrices can be more resistant [11].

2.2. Matrix–Filler Interface

The nature of the interface between the matrix and filler is an essential factor influencing the nanocomposite properties. According to Sharpe [33], the interface is defined as an intermediate region of two phases in contact, whose composition, structure, and properties vary throughout the area and are generally different from the two phases. Such phases are rarely devoid of chemical interaction. The volume of material affected by the interface interaction forms a a three-dimensional zone, called the interphase. The term interphase is widely used in the adhesion community to indicate the presence of a chemically or mechanically altered zone between adjacent phases [34,35]. The interphase concept, according to Drzal [36], is schematically represented in Figure 1.
Knowledge of the relationship between microstructure and properties in the interface region is essential for the correct use of composite materials. There is no simple quantitative relationship for interface optimization that combines polymeric matrix and fillers [16]; the physicochemical variation and the thermodynamic–mechanical principles are sources of information for the qualitative assessment of the interface phenomena. Numerous researches have been carried out to improve the properties of the composites, particularly the interface when the filling is inorganic, for example, TiO2 [17].
Studies have shown [37,38] that the interface has different properties from both the matrix and the filler material. This consists of several layers that can each affect the adhesion of the components. The filler–matrix bonding depends on the following physicochemical aspects of the interfaces of the composite: atomic arrangement, molecular conformation, the chemical constitution of the fillers, matrix and fillers morphological properties, and the diffusivity of the elements in each constituent. The adhesion between the polymeric matrix and the dispersed phase particles was explained in mechanical and thermodynamic adhesion, chemical compatibility, chemical reactions with new bonds, electrostatic attraction forces, and macromolecular interdiffusion, adsorption and watering, as shown in Figure 2 [39]. The mechanical coupling or interlocking adhesion mechanism is based on the adhesive keying into the surface of the substrate [40] and locking the rough irregularities on the surface of the nanocomposites. In many studies, it was shown that the adhesion mechanism was due to interchain entanglement and not chemical bonding between the components of the composites. The mechanical adhesion primarily depends on the forces in the transition region between the non-contacting areas [41,42]. The thermodynamic mechanism assumes that it does not require a molecular interaction for good adhesion, only an equilibrium process at the interface [43]. In neutral environments, such as air, the thermodynamics of the polymer system will attempt to minimize the surface free energy by orientating the surface into the non-polar region of the polymer. When the polymer surface is in contact with a polar substance, such as water, good adhesion requires that the interfacial tension be minimized [44]. The other theories mentioned above are explained based on the physico–chemical interactions between the components of the composites.

2.3. Fillers and Surface Modifications

Composite materials with optimal performances are obtained if an optimal adhesion between the matrix and the filler is achieved. Optimal adhesion is realized between materials of close chemical nature. It is weak between very different materials, from a chemical point of view, such as between the polymer matrix and inorganic fillers. Thus, the adhesion can be improved by treating the surface of the dispersed particles with coupling agents. Surface modification of fillers is becoming more important because of its adhesion improvement on the stress transfer between polymer and filler, which leads to an increase in the dispersion degree [45]. The coupling agent diffusion and adsorption processes at the surface of the filler particles occur at the interface. The properties of the interface and the adhesion of the components can be modified by treating the surface of the fillers before introduction into the polymer matrix. These treatments either remove the weaker layers related to the filler surface of the material or introduce new functional groups capable of influencing the adhesion between the materials.
Surface treatment of the fillers can be achieved by [46,47] the following:
The chemical interaction of the fillers with compounds that possess functional groups;
Chemical absorption on the surface of the particles of the filling material of some modifying agents;
Coating the filler particles with a suitable coupling agent.
These processes are generally laborious and increase the cost of the fillers, but they offer the possibility of considerably increasing the fillers content in mixtures without worsening their characteristics.
Modification of the surface of fillers is becoming more important because of its improvement in adhesion [48]. Hence, it is on the stress transfer between the polymer and filler, leading to an increase in the dispersion degree.

3. Titanium Dioxide Nanoparticles

3.1. Size, Shape and Specific Surface Area of the Nanoparticles

Titanium dioxide (TiO2) is the natural oxide of the element titanium. Titanium dioxide adopts four structures polymorphs found in nature rutile, anatase, brookite, and TiO2 (B). An additional four high-pressure forms have been synthesized, as follows: TiO2 (II) with the α-PbO2 structure, TiO2 (H) with hollandite, baddeleyite with ZrO2, and cotunnite with PdCl2 [49]. Among the eight structures, rutile and anatase are mostly manufactured in the chemical industry as microcrystalline materials. Thermodynamically, rutile is the most stable phase at all temperatures and pressures below 60 kbar, when TiO2 (II) becomes the favourable phase. Particle size influences surface energy and phase stability. Thus, anatase is most stable at sizes less than 11 nm, brookite at sizes between 11 and 35 nm, and rutile at sizes greater than 35 nm. Anatase and brookite are more stable than rutile at nano-size, due to the differences in surface energy. Anatase is more stable than brookite at even smaller sizes [50]. From a commercial point of view, titanium dioxide can be found in the following two common forms that differ in crystal structure: anatase and rutile [51,52,53].
Titanium dioxide can be prepared in the following various morphologies: nanoparticles, nanowires, nanotubes, and mesoporous structures. There are physical and chemical methods for synthesizing TiO2 nanoparticles in the liquid phase, as follows: hydrothermal/solvothermal method, sonochemical method, electrochemical synthesis, sol-gel method, microwave field synthesis, and vapor phase, which includes spraying, atomic deposition of layers, pulsed laser deposition, chemical vapor deposition, physical vapor deposition, and pyrolysis spray [54,55]. The controllable synthesis of TiO2 with unusual morphologies and dimensions can give the polymeric matrices with particular features and qualities.
The specific surface area of TiO2 increases as the particle size decreases, meaning nanoparticles are attracted due to van der Waal electrostatic forces. With the decreasing particle size, the ratio of surface/volume increases. Therefore, the smaller the particles are, the more important the surface properties will be, influencing agglomeration behavior and interfacial properties as a result of interaction with the polymer matrix [56,57]. The formation of particle agglomerates and non-uniform dispersion has motivated research to better process polymer–TiO2 nanocomposites. Several methods have been approached to minimize agglomeration and ensure better distribution. Such methods may be as follows: melt mixing, solution mixing in aqueous media or polymer matrices, particle surface modification involving polymer surfactant molecules or other modifiers, which must generate a strong repulsion between nanoparticles, mechanical stirring, and ultrasonic irradiation.

3.2. Surface Modification of TiO2 Nanoparticles

TiO2 nanoparticles can be directly added to the organic matrix, but due to the high surface area and high polarity, there is a strong tendency for them to aggregate. TiO2 nanoparticles form agglomerates at higher concentrations due to their high surface energy. Surface modification of TiO2 nanoparticles effectively reduces their surface energy and improves their dispersion properties in the organic matrix. Therefore, to improve the homogeneous dispersion of nanoparticles, many researchers have focused on the surface modification of nanoparticles and a new method for incorporating inorganic nanofiller into an organic matrix [58,59,60]. Several ways have been employed to modify the surface of nanoparticles [61,62].
The surface modification of TiO2 nanoparticles is often conducted by either a physical or chemical method. The chemical method has attracted the attention of many researchers because the interactions between inorganic nanoparticles and the matrix are much stronger [63]. The surface modification of nanoparticles by chemical treatments is a useful method to improve the dispersion stability of TiO2 nanoparticles and the development of interfaces between the organic and inorganic phase. In this regard, the concept of silane coupling agent was reported by Plueddemann and et al. [64]. Researchers found that organofunctional silanes are silicon chemicals that contain both organic and inorganic reactivity in the same molecule, and which can be used as coupling agents [65,66]. Coupling agents connect resin and fillers, and improve the physical, mechanical and electrical properties of composites. Moreover, they enhance the wetting of inorganic substrates, decrease the viscosity of the resin during mixing, and ensure smoother surfaces of composites [67,68].
The general formulation of the coupling agent molecule is as X–R, where X interacts with the filler and R is compatible with the polymer. Organosilanes are of the form R–Si–(OR’)3, where OR’ can be methoxy, ethoxy, acetoxy, and R can be alkyl, aryl or organofunctional group [56]. According to this structure, the following steps may take place, as shown in Figure 3:
Hydrolyzation of alkoxy groups obtaining silanol, which reacts with the mineral surface;
The condensation reaction between silanol molecules;
Bond formation between TiO2 nanoparticles and the organofunctional group.
The choice of organosilane is established, taking into account the polymers chemical structure to be compatible. For example: for a phenolic and epoxy resin an epoxy silane, or an amino silane is recommended and for an unsaturated polyester resin a methacrylsilane. The reactivity of the thermosetting polymers should be close to that of organosilane. For a thermoplastic matrix, bonding occurs by diffusion of the organosilane network in the interphase region of the composite [66].
There were silane coupling agents used, such as 3-methacryloxypropyl-trimethoxysilane (MPS) [68], 3-aminopropyltriethoxysiane (APTES) [69], γ-glycidoxypropyltrimethoxysilane (GPS) [70], n-propyltriethoxysilane and 3-methacryloxypropyltrimethoxysilane [71], which change the hydrophilic particles into a hydrophobic surface by providing some molecules with certain hydrophobicity.
Some coupling agent recommendations for the surface modification of TiO2 nanoparticles is given in Table 1.
Silane coupling agents are usually employed to realize chemical modification. These can offer hydrolyzable groups bonding with the inorganic particles. After bond formation, the organosilane functional groups of silane coupling agents form a hydrophobic layer on the surface of the inorganic nanoparticles. Different coupling agents have been used to modify the surface of TiO2 and improve the interfacial interactions necessary for the successful incorporation of these hydrophilic nanoparticles into hydrophobic polymer matrices.
The surface modification of TiO2 has been reported using different silane coupling agents, such as 3-aminopropyltriethoxysilane (APTES). The photocatalytic activity of TiO2 has been shown to increase with increasing the concentration of APTES used [87]. For example, Mallakpour and Barati [88] reported the surface modification of TiO2 nanoparticles by the reaction with APTES. The silane coupling agent was adsorbed on the surface of the nanoparticles at its hydrophilic end and interacted with the hydroxyl groups pre-existing on the surface of the nanoparticles. Thus, it was confirmed that the heat stability of the nanocomposite was improved. Shakeri et al. [89] studied the self-cleaning capability of surfaces covered TiO2 nanoparticles, modified by APTES. They concluded that the surface could degrade the dye used as an organic pollutant due to the obtained coating being stable. Klaysri et al. [90] proposed a one-step synthesis method of APTES-functionalized TiO2 surface. They showed that obtained nanomaterials are capable of the photocatalytic decolonization of methylene blue.
Modification of the surface of TiO2 nanoparticles with silane coupling agents was obtained via reflux in an aqueous solution [75,91]. Chen et al. investigated the interactions between 3-aminopropyltrimethoxysilane (APTMS) and phenyltrimethoxysilane with commercially available TiO2 nanoparticles (Degussa P-25) [91]. They obtained results showing that the silane coupling agents used bind covalently on the surface of the TiO2 nanoparticles. In another study, Zhao et al. reported the cross-linking and chemical bonding mechanisms of APTMS and 3-isocyanatopropyltrimethoxysilane on TiO2 nanoparticles [75].
To improve TiO2 nanoparticles dispersion and enhance the interactions between the nanoparticles and polymeric matrix (polyamide/modified–TiO2 nanocomposites), the surface of TiO2 was modified with a 1,3,5-triazine based silane coupling agent [92].
Caris et al. [93] used conventional emulsion polymerization to encapsulate TiO2 in poly(methyl methacrylate) (PMMA). Sidorenko et al. [84] investigated the radical polymerization of styrene and methyl methacrylate (MMA). This reaction was initiated at the surface of TiO2 particles by adsorbed hydroperoxide macroinitiators. Erdem et al. [94] encapsulated the TiO2 nanoparticles by miniemulsion polymerization of styrene, polybutene-succinimide pentamine being used as the stabilizer at the oil/water interface. Rong et al. [95] used the TiO2 nanoparticles modified by 3-(trimethoxysilyl) propylmethacrylate (MPS) to copolymerize styrene with the methacrylate group of MPS, by free-radical polymerization. Yang and Dan [96] used a similar approach by graft polymerized MMA on the modified surface of the TiO2 nanoparticles.
Milanesi et al. used a mixture of isomeric octyltriethoxysilanes (OTES), highlighting the hydrophobic layer structure. They concluded that the cross-linking (via Si–O–Si bonds) and chemical bonding (via Ti–O–Si bonds) of silanes onto TiO2 nanoparticles occurred [97]. Xiang et al. used 3-methacryloxypropyl-trimethoxysilane (MPS) to modify the TiO2 surface to enhance the compatibility of TiO2 nanoparticles in the poly(butyl acrylate) (PBA) matrix. The modified TiO2 presented good compatibility in the PBA matrix [98]. In another study [83], Xiang showed the hydrophobic surface modification of TiO2 to produce acrylonitrile-styrene-acrylate (ASA) terpolymer–TiO2 composites for cool materials. Wang et al. [99] functionalized the commercial TiO2 nanoparticles in an aqueous solution via ultrasonic treatment at room temperature with 3-(trimethoxysilyl)propyl methacrylate.
Godnjavec et al. have coated TiO2 nanoparticles by 3-glycidyloxypropyltrimethoxysilane (GLYMO) as an additive in a clear polyacrylic coating. According to their results, grafting GLYMO on the nanoparticles surface improved the dispersion, transparency, and UV protection of the clear acrylic coating [100].
Yang et al. [101] reported silanization of TiO2 particles through a sol-gel method. Based on their results, vinyl triethoxysilane (VTES) as a surface modifier improved the stability of dispersion and suspension in tetrachloroethylene. Dalod et al. [50] modified TiO2 nanoparticles with amino silane groups using a hydrothermal method and found that the nanoparticles shape and structure depends on the type of silane coupling groups.
Tangchantra et al. [102] investigated the effect of different silane coupling agents on the surface grafting of TiO2 with hexadecyl trimethoxysilane (HTMS), triethoxyvinylsilane (TEVS), and aminopropyl trimethoxysilane (APS). The results showed that silane coupling agents could modify the surface of TiO2 nanoparticles via the hydrolytic condensation of titanium isopropoxide. The TEVS agent improved the dispersibility of TiO2 particles and showed optimum mechanical properties.
The appropriate surface modification on nanoparticles leads to better dispersion and compatibility in the polymer matrix. The formation of chemical and physical interactions with the polymer matrix could guarantee remarkable mechanical properties of polymeric nanocomposites.

3.3. Properties, Commercial Products and Applications

At the nanoscale size, the material properties may dramatically change and differ significantly from their bulk counterparts.
Particular attention has been paid, in recent years, to obtaining TiO2 with photocatalytic properties [103,104,105,106], optical properties [107], with applications related to the degradation of pollutants [108,109,110], and the realization of the photoelectrochemical cells [111]. Also of interest are titanium dioxide films deposited on various substrates to obtain special characteristics, such as surfaces with self-cleaning properties [112,113].
All applications of TiO2 nanoparticles depend on their crystal structure, morphology, specific surface area, particle size, and form. TiO2 has been widely used in the industry for many years for its numerous and diverse applications, as shown in Table 2.
The applications that can be mentioned are sensors, photo-conductors, additives in plastics, catalysts, photo-/electrochromics and photovoltaics applications, dye-sensitized solar cells, sunscreens, paints, antimicrobial applications, water purification by photocatalysis processes, biosensing, and drug delivery [114]. TiO2 nanoparticles incorporated into outdoor building materials, such as paving stones or paints, can reduce volatile organic compounds and nitrogen oxide concentrations.
TiO2 is a material with multifunctional properties that can be incorporated in polymeric matrices as a filler to develop new nanocomposites with enhanced properties [115].

4. Polymeric Nanocomposites with TiO2 Filler

4.1. Preparation Methods

Polymeric matrix nanocomposites can be obtained using injection molding, compression molding, in situ polymerization, sol-gel, melt mixing and sintering.
In situ polymerization involves the dispersion of inorganic nanoparticles in a monomer phase as a first step, followed by bulk phase polymerization. This process is mainly used for thermosetting polymers. As a result, unstable nanocomposites can be transform into a different morphology than expected. The in situ polymerization method is a simple and inexpensive method. The nanocomposites with the polymer matrix, and inorganic filler with good filler distribution in the polymer matrix, can be obtained [116].
Most compression molding techniques require pre-treatment of the nanoparticles with curing, but injection molding is the most widely used process for obtaining nanocomposite materials. Injection molding can be used in a variety of applications, in both commercial and research fields [117]. Sintering, powder compaction and sol-gel are all alternative techniques to produce polymeric composites. However, the operating conditions (temperature, pressure, time, etc.) are far more than those of injection molding [118]. Some reports were found in the literature focusing on obtaining TiO2 nanocomposites with the polymeric matrix, as shown in Table 3.
Studies on polymer–TiO2 nanocomposites prepared by melt mixing have shown a slight improvement or no change in mechanical properties [119,120]. Somani et al. [121] received highly piezoresistive conducting polyaniline/TiO2 composite by in situ deposition technique at a low temperature. Feng et al. [122] synthesized a composite of polyaniline-encapsulating TiO2 nanoparticles by in situ emulsion polymerization. They investigated and explained the interaction between polyaniline and nano-TiO2 particles, and the nature of chain growth according to Fourier transform infrared (FTIR) spectra. Xia and Wang [123] prepared a polyaniline/nanocrystalline TiO2 composite by ultrasonic irradiation. They think that ultrasonic irradiation provides a new way to prepare 0–3-dimensional conducting polymer/nanocrystalline composites.
Titanium dioxide has been used to reinforce polypropylene (PP) via extrusion, followed by injection molding, by Alghamdi [124]. There were presented to the mechanical and structural aspects of PP for different loading of TiO2 filler (up to 30 wt. %). As the TiO2 weight percent increases, the impact strength decreases. This behaviour is expected because the PP is incompatible with TiO2. The PP phase is non-polar, hydrophobic and has low surface energy, while TiO2 represent the polar phase, hydrophilic and high surface energy for TiO2. The highest resilience value was recorded for the sample with 20% TiO2 (37.09 ± 5.3 J/m).
Mourad et al. [117] studied HDPE nanocomposites with 5% TiO2, obtained by injection molding under the following different processing parameters: temperature, pressure, injection velocity, and injection time. The results showed the influence of processing parameters on the mechanical and thermal properties of HDPE–TiO2 nanocomposites. Mechanical testing revealed that the tensile strength varied from 22.5 to 26.3 MPa, while the Young modulus increased by 8.6% as the molding temperature increased.
Vladuta et al. [125] investigated the effect of the TiO2 nanoparticles on the PET–rubber interface in nanocomposites obtained from waste by compression molding. The modifications in surface energy, morphology and crystalline structure were discussed for samples kept under visible light and UV radiation. TiO2 develops new physical interactions in the composite, but induces, even in visible light, oxidation processes. The results indicated that the optimum concentration for TiO2 to the composites, for obtaining better interface properties, is 0.25 wt. %.
Regardless of the method of obtaining nanocomposites with the polymeric matrix, it is found that the nature of filler has a significant influence on mechanical properties.

4.2. Mechanical Properties

Titanium dioxide is used as a filler in many polymeric matrices because of the improved physical and mechanical properties it yields. Many studies showed improvements in the mechanical strength and modulus of TiO2-filled polymeric nanocomposites compared to the pristine-base matrix. The mechanical properties of the TiO2 nanocomposites depend significantly on their internal structure. The poor compatibility of hydrophilic TiO2 nanoparticles with a hydrophobic polymer matrix may lead to particle aggregates and/or agglomerates. The aggregates create defect sites in the nanocomposites, and the improvement in mechanical properties is not observed. More uniform dispersion of nanoparticles is recommended, using one-dimensional nanoparticles, i.e., nanorods, nanotubes or nanoribbons, particles with a high aspect ratio [46]. Several factors that may influence the mechanical properties of composites with a polymer matrix and inorganic fillers are presented in Figure 4.

4.2.1. The Nature of the Filler

TiO2 fillers affect the basic mechanical properties of the polymer. The effect of TiO2 fillers on composites properties depends on the particle size and shape, concentration and the interaction with the matrix, as shown in Table 3. For example, to increase the modulus and hardness of polymers, micrometre-sized inorganic particles are frequently applied. However, a reduction in the material ductility may take place. By diminishing the particle size or by enhancing the particle volume fraction, the strength can be improved. Still, in some cases, the fracture toughness and modulus remain relatively independent of the particle size. The properties of TiO2 that make it a good filler for composite materials are good dispersibility in the polymer system and good heat stability. Titanium dioxide has a relatively high elastic modulus, which can be frequently combined into various polymers to obtain the composites mechanical gain.
Mikešová and et al. [126] studied the effects of nanoparticles and the properties of the nanocomposites of polypropylene and filler TiO2. They used isotactic polypropylene (PP) as a matrix, and as fillers they used TiO2 in the following different shapes: a commercial titanium dioxide micropowder (mTiO2; a mixture of anatase and rutile), a commercial titanium dioxide nanopowder (nTiO2; anatase modification), and titanate nanotubes (TiNT). More series of samples were obtained with PP unmodified and with PP modified by electron beam irradiation (PP*), resulting in PP*/TiX composites (i.e., PP*/mTiO2, PP*/nTiO2, and PP*/TiNT). These were prepared by melt mixing of PP* with 5 wt. % of TiX. The stiffness and microhardness properties of PP*/TiX systems are improved in the order PP*/mTiO2, PP*/nTiO2, PP*/TiNT, due to the specific surface of the TiX particles.
Nano-sized TiO2 was further studied in starch/(poly[vinyl alcohol]) blends by Sreekumar et al. [127]. The nano-sized TiO2 could provide the composite with superior mechanical properties because of good interfacial adhesion between the polymer matrix and filler.
Bora et al. [128] studied the effect of TiO2 particle concentrations (up to 25 wt. %) on the properties of polyphenylenesulphide (PPS)–TiO2 composites. The increase in TiO2 particle concentrations in the PPS matrix improves the stiffness of the composite. High values of flexural and residual flexural strength were obtained at 10 wt. % TiO2 particle concentrations. Saluja et al. [129] obtained polyester composites filled with TiO2 concentrations up to 25 wt. %. This study shows that the addition of TiO2 particles improves the effective thermal conductivity of polyester–TiO2 composites, the glass transition temperature (Tg), and the reduction in the coefficient of thermal expansion (CTE).
The mechanical properties of nanocomposites depend significantly on their internal structure. In the nanocomposite, TiO2 nanoparticles can appear as agglomerations due to their low compatibility with the hydrophobic polymer matrix.
In this case, the large surface area of the nanowires decreases rapidly, the aggregates create defect sites in the nanocomposites, and no improvement in the mechanical properties is observed. A more uniform dispersion of nanoparticles, using one-dimensional (1D) nanoparticles, i.e., nanorods, nanotubes, or nanoribbons, would improve these properties. Compared with the isometric nanoparticles, a large surface-to-volume ratio of the 1D nanoparticle generally improves the nanocomposites properties. Contrary to the anatase polymer nanocomposites, only a few papers concerning polymers filled with titanate nanotubes have been found in the literature [130,131,132].
The majority of nanoparticle fillers added in the polymer matrix improve mechanical properties such as flexibility, ductility, hardness, and strength and stiffness, even in small amounts.

4.2.2. The Nature of the Polymer Matrix

Polymer–TiO2 nanocomposites have been successfully synthesized in different polymer matrices such as the following thermoplastic polymers: polyacrylate, poly (methyl methacrylate), polyimide, polystyrene, and polyolefines; the following thermosetting polymers: polycarbonate, polyamide 6, epoxy, unsaturated polyester; and silicone elastomer [77,132].
Saritha et al. [133] studied the incorporation of TiO2 in rubber composites. The tensile strength, modulus, and tear strength increased with increasing TiO2 loading. More recently, processing techniques were developed to allow the size of TiO2 to decrease to the nanoscale. Manap et al. [134] demonstrated that TiO2 and multi-walled carbon nanotubes (MWCNT) as filler reinforcements could address the agglomeration issue, by exhibiting even distribution of particles in the TPU matrix. The combination of MWCNT and TiO2 in the TPU matrix enhanced the mechanical and thermal properties significantly, this being a good heat insulator.
In the function of the matrix nature, the percentage by weight of the inorganic filler introduced can remain very low (on the order of 0.5% to 5%) due to the incredibly high surface area-to-volume ratio of the particles. This area can generate a new material behavior, which is widely determined by interfacial interactions, offering unique properties and an entirely new class of materials. Several important types of research in this regard are presented in Table 3.
Table 3. Types of nanocomposites with polymeric matrix and TiO2 filler.
Table 3. Types of nanocomposites with polymeric matrix and TiO2 filler.
CompositesMaterials Methods Results-Mechanical PropertiesRef.
Thermoplastic matrix
Polystyrene (PS)–TiO2 Matrix: polystyrene.
Filler: TiO2 (0.19 µm)
Coupling agent: 3-amino ethoxy silane (0.1, 0.5 and 1 wt. %.).
Obtaining: mixing of matrix with TiO2-coupling agent
Characterization: mechanical tests, SEM analysis.
Values of Young’s modulus, tensile strength, elongation at break, flexural strength increase with linearly filler concentration followed by a decrease beyond 15 wt. %. [45]
Polyphenylene sulfide (PPS)–TiO2Matrix: polyphenylenesulphide (PPS)
Filler: TiO2 (200 nm; 0, 5, 10, 15, 20, and 25 wt. %)
Obtaining: injection molding.
Characterization: solid particle erosion test, three-point bending test, thermal analyzing methods.
The flexural modulus of composites increased with the increase in TiO2 concentration up to 10 wt. %, and then it decreases. TiO2 filler caused to reduce the erosion resistance of the PPS composites.[128]
Polypropylene (PP)–TiO2Matrix: PP pellets;
Filler: TiO2 (0, 10, 20 and 30 wt. %)
Obtaining: injection molding
Characterization: mechanical properties: tensile stress, impact tests; TGA
The highest resilience value recorded for the sample with 20% TiO2 (37.09 ± 5.3 J/m). Tensile stress shows a decrease and the E modulus increase as the weight percent of TiO2 increases. [124]
Polypropylene (PP)–TiO2Matrix: polypropylene (PP)
Filler: TiO2 micropowder; TiO2 nanopowder titanate nanotubes (TiNT)
Obtaining: melt mixing; samples types PP*/TiX (PP*/mTiO2, PP*/nTiO2, PP*/TiNT) and samples with PP unmodified.
Characterization: SEM analysis, TEM analysis, mechanical properties
The stiffness and microhardness of the PP–TiNT nanocomposites increase by 27% and, respectively, 33%. In the PP–nTiO2 nanocomposites, the increase in these mechanical characteristics is lower.[126]
Polypropylene (PP)–TiO2Matrix: PP homopolymer
Filler: TiO2 (0–3 wt. %)
Obtaining: melt compounding;
Characterization: mechanical properties, thermogravimetric analysis, DSC, SEM analysis
The addition of TiO2 nanoparticles increases the mechanical properties of PP fibres. Tenacity is increased by 72.69% for the PP–TiO2 (3 wt. %) nanoparticle. Elongation at break of the PP fibres with TiO2 (1.5 wt. %) indicated an increase of 15.79%.[135]
Polypropylene (PP)-rice husk–TiO2Matrix: polypropylene (PP)
Filler: rice husk and TiO2
Obtaining: injection molding
Characterization: mechanical properties, SEM, TGA
Incorporating inorganic filler TiO2 into PP/RH significantly enhanced the green hybrid PP/RH/TiO2 composites mechanical properties and thermal stabilities. The maximum values of tensile strength and Young modulus were 41.2 MPa for PP/RH (10wt. %)/TiO2 (3wt. %), respectively, for PP/RH (40wt. %)/TiO2 (3wt. %)[136]
polyurethane (TPU)–TiO2 Matrix: polyurethane (TPU) matrix with multi-walled carbon nanotube (MWCNT);
Filler: TiO2 (particle diameter—0.19 μm).
Obtaining: injection molding.
Characterization: mechanical properties: tensile test, DMA, TGA,
The composites have good mechanical properties: tensile stress was 4.46 MPa, elongation at the break—49%, and Young’s Modulus— 9.17 MPa. [134]
thermoplastic polyurethane (TPU)–TiO2Matrix: thermoplastic polyurethane
Filler: TiO2 nano-particles
Coupling agent: aminopropyl trimethoxy silane (APS)
Obtaining: mixing of matrix with filler;
Characterization: elemental analysis, FTIR spectroscopy, TGA, mechanical properties.
For composite with TiO2 (3 wt.%), tensile strength and Young’s modulus were increased by 72% and 48.9, respectively. Higher values were obtained when modified TiO2 was used, at low percentages (1 wt.%). [77]
polybutylene succinate (PBS)–TiO2 Matrix: polybutylene succinate (PBS);
Filler: TiO2 (20 nm; 0, 0.5, 1, 2, 5, and 10 wt. %)
Obtaining: vane extruder.
Characterization: SEM, TEM, XRD, DSC, TGA, DMA; mechanical test, UV transmittance.
TiO2 has little effect on the impact strength of the composite material. The flexural modulus of composites improved by 36.3% with TiO2 (10 wt. %) addition. The tensile modulus of PBS–TiO2 (10 wt. %) was higher by 15.5% than that of pristine PBS.[137]
polyetheretherketone (PEEK)–TiO2 Matrix: PEEK.
Filler: TiO2 powder (1, 3, 5 wt.%)
Obtaining: mixing and extrusion forming;
Characterization: density and Melt Flow Index (MFI) measurement, DSC, UV thermal, mechanical test
E modulus increase with TiO2 content. The PEEK-1% TiO2 sample has a tensile strength higher than that of pristine PEEK. TiO2 (5% vol.) particles act effectively as UV blocker retarding the photo-degradation of PEEK.[138]
poly(ethylene terephthalate) (PET)–TiO2
Poly(lactic acid) (PLA)–TiO2
Matrix: poly(ethylene terephthalate) (PET) and poly(lactic acid) (PLA);
Filler: TiO2 (20 nm);
Obtaining: extrusion forming;
Characterization: analysis—DSC, XRD, SEM, DMTA, UV–Visible test, mechanical test.
The mechanical properties of PET–TiO2 and PLA–TiO2 composites have maximum values at a loading level of 3% TiO2.[139]
poly(L-lactide-co-ε-caprolactone) (PLCL)–TiO2 nanocomposites Matrix: PLCL;
Filler: TiO2 (20 nm)
Coupling agent: silane coupling agent NH2(CH2)3Si(OC2H5)3
Obtaining: solution casting method.
Characterization: analysis—FTIR, DSC, TEM, tensile test, shape memory;
For composite with TiO2 (5%) the ultimate tensile strength and the elongation at break increase to 35.4 MPa and 444.6%, which are 113% and 11% higher than that of pure PLCL.[140]
Poly(L-Lactide) (PLLA)–TiO2Matrix: poly(L-Lactide) (PLLA)
Filler: TiO2 (<25 nm particle size) and Halloysite nanoclay (HNT) (Al2Si2O5(OH)4.2H2O);
Obtaining: compression molding.
Characterization: mechanical test
Young modulus had a significant increase (p ≤ 0.05) with the addition of TiO2 up to 2.5 g TiO2/100 g PLLA. Regarding the tensile strength, better results were also achieved when adding 2.5 g TiO2/100g PLLA.[141]
Poly(lactic acid) (PLA)–TiO2Matrix: PLA (4032D, 1.2–1.6% D-isomer lactide)
Filler: TiO2 (20 nm);
Obtaining: injection molding;
Characterization: SEM, TEM, dynamic rheological measurements, DSC, TGA, tensile testing, UV transmittance
Samples show a higher elongation at break, except for 15 wt. % TiO2. Elongations of nanocomposites with 1–2% TiO2 are about 19.1% and 24% higher than the pristine PLA.[142]
Poly(lactic acid) (PLA)–TiO2Matrix: poly(lactic acid) (PLA)
Filler: TiO2 (1, 3, 5, 10 wt.%)
Coupling agent: c-methacryloxy
propyltrimethoxy-silane)
Obtaining: in situ polymerization, Characterization: DSC, TGA, XDR, SEM, thermal and mechanical propertiesThe tensile strength, elongation at break, and Young’s modulus of PLA–TiO2 (3 wt.%) composites are improved to a certain degree compared with those of pristine PLA.[143]
Thermosetting matrix
epoxy–TiO2 nanocompositesMatrix: mixture (resin + hardener);
Filler: TiO2 (0.5, 1, 2, 3, 4, 5, 8 and 10% vol.);
Obtaining: mixing of resin + hardener and filler;
Characterization: tensile test, dynamic mechanical analysis;
The incorporation of TiO2 nanoparticles into the epoxy resin improved flexural stiffness, flexural strength, and fracture toughness of the polymer. [144]
epoxy–TiO2 nanocompositesMatrix: epoxy resin
Filler: TiO2 (5–40 nm, 0.5–2 wt.%);
Obtaining: mixing of matrix with filler;
Characterization: thermal properties,
mechanical properties, morphology, viscoelastic properties.
TiO2 composites with dimensions between 5–10 nm showed better properties than those with larger dimensions (20–50 nm).[145]
epoxy–TiO2 nanocompositesMatrix: mixture (resin+hardener);
Filler: TiO2 (1, 3, 5, 10 wt.%)
Coupling agent: methyl isobutyl-ketone; dodecylbenzene-sulfonic acid
Obtaining: mixing of matrix, filler and coupling agent;
Characterization: FTIR, SEM, XRD, TGA, mechanical tests
The mechanical properties of materials are found to improve with TiO2, but degrade if the nano-TiO2 exceeds 3%.[146]
epoxy–TiO2 nanocompositesMatrix: epoxy resin (DER 331TM)
Filler: TiO2 (220 nm, 50 nm and 17 nm crystal diameter);
Coupling agent: isophorone diamine (IPDA) + salicylic acid.
Obtaining: mixing of matrix, filler and coupling agent;
Characterization: mechanical test, XPS, SEM
The highest tensile stress values were found at 3 wt. % TiO2 (17 nm and 50 nm) and 5 wt. % TiO2 (220 nm). The maximum flexural properties were found at a lower TiO2 fraction of 1 wt.% only.[147]
epoxy–TiO2 micro and nanocompositesMatrix: epoxy resin:curing agent = 2:1 (wt. %)
Filler: TiO2 (0.2 μm; 1, 5, 10, 15 wt. %); TiO2 (21 nm; 0.5, 1, 3 wt. %).
Obtaining: mixing with an electrical stirrer,
Characterization: tensile test, tensile creep-recovery test, tensile stress relaxation tests, SEM.
TiO2 nanocomposites have better strength properties than TiO2 microcomposites due to the size of the particle.[148]
vinyl ester resins–TiO2 nanocompositesMatrix: vinyl ester:styrene monomers (55:45 wt. %)
Filler: TiO2 (21 nm; 50 m2/g; 1, 2.5, and 5 wt. %).
Coupling agent: polymeric coupling BYK-C 8000
Obtaining: shear mixing and ultrasonication;
Characterization: tensile test, flexural test, impact test, SEM
For nanocomposite with 0–2.5 wt. % TiO2, the tensile strength exhibits increasing tendency, while loading more than 2.5 wt. % leads to its decline.[149]
epoxy resin–polyurethane (EP-PU)–TiO2Matrix: EP-PU epoxy resin;
Filler: TiO2 (0.42 g/cm3; 25 nm)
Coupling agent: isopropyl tri(dioctylpyrophosphate) titanate (TCA201)
Obtaining: mixing EP–PU and TCA201–TiO2
Characterization: FT-IR spectroscopy, SEM analysis, TGA analysis, mechanical properties, dielectric constant
The shear strength reached the maximum value (27.14 MPa) for EP–PU/TiO2 (3 wt. %) and its thermal decomposition temperature increase by 17.48 º C more than that of EP–PU matrix. The dielectric constant and dielectric loss showed 4.27 and 0.02, respectively.[85]
Elastomeric matrix
TiO2–natural rubber compositesMatrix: natural rubber (NR)
Filler: TiO2 (KEMOX RC 800 PG) and the surface-modified nanosilica
Obtaining: hydraulic press under a pressure
Characterization: stress relaxation measurements, SEM, AFM, effect of strain level, effect of ageing
The rate of stress relaxation was higher for silica-filled NR than TiO2-filled NR. This is due to the high degree of agglomeration in silica compared to TiO2. The relaxation rate increased with increasing TiO2 loading.[150]
TiO2–natural rubber compositeMatrix: natural rubber stabilised with ammonia;
Filler: TiO2 dispersion (2, 4 and 6 pphr)
Obtaining: TiO2 dispersions added in matrix;
Characterisation: tesnsile test
The results showed improvement in both elongations at break and tensile strength data at low filler concentration (2 phr).[151]
TiO2–natural rubber compositesMatrix: natural rubber latex centrifuged with ammonia;
Filler: TiO2 (3 mm;.13 g/mL); TiO2 (15–40 nm; 4.26 g/mL);
aditives: zinc oxide, stearic acid, N-cyclohexyl-benzothiazyl-sulphenamide, N2′-propyl-N-phenylenediamine, and S
Obtaining: TiO2 dispersion was immersed in natural rubber latex.
Characterization: tensile test, SEM, TEM, XRD
The tensile strength of nano-sized TiO2-filled natural rubber composites (23.04 MPa) is superior to micro-sized TiO2-filled natural rubber composites (19.62 MPa) (for 6 phr of micro- and nano-s)[152]
TiO2–natural rubber compositesMatrix: natural rubber;
Filler: TiO2-15, 25, 45, 85 wt. %
aditives: stearic acid, sulfur powder and zinc oxide;
Obtaining: compression molding;
Characterization: mechanical properties; dynamic mechanical properties; thermal stability
TiO2 as filler allows obtaining materials with improved mechanical properties and thermal stability compared to the pristine natural rubber vulcanizates. [153]
TiO2–chlorobutyl rubber compositesMatrix: chlorobutyl rubber (CBK 150) with 1.2% Cl;
Filler: TiO2 (10–30 phr.)
Additives: stearic acid, zinc oxide, sulfur, and zinc
Obtaining: mixing in a two-roll mill
Characterization: mechanical properties, morphology (SEM, AFM), thermophysical measurements, diffusion experiments
The tensile strength of the composites increases by 250% when the filler loading goes to 40 phr (tensile modulus the same).[133]
Acrylonitrile–Butadiene–Styrene–TiO2 nanocompositesMatrix: acrylonitrile butadiene styrene (ABS)
Fillers: TiO2 (25–50 nm; 0.5, 2.5, 5 and 10 wt. %) and ATO (size < 50 nm)
Obtaining: mechanical homogeniser.
Characterization: SEM, AFM and Raman analysis, thermal properties, tensile test, flexural tests, micro-hardness tests.
The tensile strength of ABS/TiO2 and ABS/ATO nanocomposites increased by 7% at the 2.5 wt. % TiO2 filler, respectively, by 9.2% at 0.50 wt. % ATO filler. The modulus of elasticity increases up to 5 TiO2 wt. % and then decreases.[31]
When designing new polymer–TiO2 nanoparticle composites, the following aspects should be considered:
Nature of filler and polymer matrix;
Amount of filler;
The distribution of filler, this should not form agglomerates in the samples;
Concentration of coupling agent for modifying of filler surface;
The method of obtaining, which is an essential factor.
The impact resistance of polymer matrices with TiO2 filled is of particular interest to researchers, as long as it represents the weak point of most composite materials. Hardening of thermoplastics by modification with elastomers could be a new way to solve this problem. It is recommended to study new cheaper and more efficient polymer matrices to produce composites with predetermined properties. In this case, we recommend using of polymeric waste as a matrix for obtaining nanocomposites with TiO2 filler.
Polymer nanocomposites give a new way to overcome the limitations of pure polymers or their traditional composites. Nowadays, polymer nanocomposites with TiO2 filled represent an area of interest for many researchers. This article contains information on the nature of the polymer matrix (thermoplastic, thermosetting, elastomeric) and the type of TiO2 filler, processing methods, possible surface modifications of the filler and how they influence the mechanical properties of nanocomposites, thus completing the areas of knowledge for many researchers.

4.3. Advantages, Limits and Applications

Polymeric materials can be used as matrices in TiO2 nanocomposites due to their good thermal stability, environmental resistance (durability), and electrical, chemical and mechanical properties. However, it is well known that some polymers (e.g., epoxy resin, polyamides) are highly brittle. This disadvantage limits the application of these polymers in products that require high impact and fracture strength. TiO2 filler added in the polymer matrix improves the mechanical performance of the polymeric nanocomposites over conventional polymer composites, as shown in Table 4. Finally, typical existing and potential applications are shown in Figure 5.

5. Conclusions

An essential characteristic of polymers is modifying their inherent physical properties by adding fillers, while retaining their characteristic processing ease. By adding inorganic fillers into the polymers matrix, composite materials become stronger, stiffer, electronically conductive, magnetically permeable, flame retardant, more challenging, and more wear-resistant.
After reviewing part of the existing literature on polymeric composites with TiO2 fillers, it is found that the interfacial connection between the filler and polymer matrix is an important element for determining the mechanical properties of the composite.
The addition of TiO2 nanoparticles into the polymeric matrix demonstrates their ability to significantly improve important mechanical properties (tensile modulus, tensile strength, toughness and fracture toughness, fracture energies, flexural modulus, flexural strength, elongation at break, fatigue crack propagation resistance, abrasion, pull-off strength, and fracture surface properties), even at low filler contents.
From the literature, one can conclude that the mechanical properties of the composites with the polymer matrix depend on the particle size, and particle–matrix interface adhesion and loading (type, quantity, filler distribution and orientation, and void content). Along with those properties, the interfacial bonds and the interphase load mechanisms also play an essential role.
Studies performed on polymeric matrix nanocomposites filled with TiO2 nanoparticles were performed to verify the influence of several variables (shape, size, % loading, surface change, etc.) and also to propose various areas of applicability of these nanocomposites.

Author Contributions

Conceptualization, C.C., L.A. and A.E.; methodology, C.C.; validation, C.C., L.A. and A.E.; formal analysis, C.C.; investigation, L.A.; resources, A.E.; data curation, C.C., L.A. and A.E.; writing—original draft preparation, C.C.; writing—review and editing, C.C., A.E., L.A.; visualization, A.E.; supervision, L.A.; project administration, C.C., L.A.; funding acquisition, C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Romanian National Authority for Scientific Research and Innovation, CCCDI-UEFISCDI, project number 169/2020 ERANET-M.-3D-Photocat, within PNCDI III.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mittal, V. Optimization of Polymer Nanocomposite Properties; Wiley: Weinheim, Germany, 2010. [Google Scholar] [CrossRef]
  2. Schadler, L.S.; Brinson, L.C.; Sawyer, W.G. Polymer nanocomposites: A small part of the story. JOM 2007, 59, 53–60. [Google Scholar] [CrossRef]
  3. Young, R.; Kinloch, I.A.; Gong, L.; Novoselov, K. The mechanics of graphene nanocomposites: A review. Compos. Sci. Technol. 2012, 72, 1459–1476. [Google Scholar] [CrossRef]
  4. Oliviera, M.; Machado, A. Preparation of Polymer-Based Nanocomposites by Different Routes. In Nanocomposites: Synthesis, Characterization and Application; Wang, X., Ed.; NOVA Publishers: Hauppauge, NY, USA, 2013. [Google Scholar]
  5. Kumar, R. Polymer-Matrix Composites Types, Applications & Performance; Nova Science Publishers Inc.: Hauppauge, NY, USA, 2014. [Google Scholar]
  6. Seferis, J. Role of the Polymeric Matrix in the Processing and Structural Properties of Composite Materials; Springer: New York, NY, USA, 2013. [Google Scholar]
  7. Venditti, I.; D’Amato, R.; Russo, M.V.; Falconieri, M. Synthesis of conjugated polymeric nanobeads for photonic bandgap materials. Sens. Actuators B Chem. 2007, 126, 35–40. [Google Scholar] [CrossRef]
  8. Soares, I.L.; Chimanowsky, J.P.; Luetkmeyer, L.; Da Silva, E.O.; Souza, D.D.H.S.; Tavares, M.I.B. Evaluation of the Influence of Modified TiO2 Particles on Polypropylene Composites. J. Nanosci. Nanotechnol. 2015, 15, 5723–5732. [Google Scholar] [CrossRef] [PubMed]
  9. Monteiro, M.; Neto, R.C.; Santos, I.C.S.; Da Silva, E.O.; Tavares, M.I.B. Inorganic-organic hybrids based on poly (ε-Caprolactone) and silica oxide and characterization by relaxometry applying low-field NMR. Mater. Res. 2012, 15, 825–832. [Google Scholar] [CrossRef] [Green Version]
  10. Rubab, Z.; Afzal, A.; Siddiqi, H.M.; Saeed, S. Preparation, Characterization, and Enhanced Thermal and Mechanical Properties of Epoxy-Titania Composites. Sci. World J. 2014, 2014, 515739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Kierys, A.; Zaleski, R.; Buda, W.; Pikus, S.; Dziadosz, M.; Goworek, J. Nanostructured polymer–titanium composites and titanium oxide through polymer swelling in titania precursor. Colloid Polym. Sci. 2012, 291, 1463–1470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Suzuki, N.; Kiba, S.; Kamachi, Y.; Miyamoto, N.; Yamauchi, Y. Unusual reinforcement of silicone rubber compounds containing mesoporous silica particles as inorganic fillers. Phys. Chem. Chem. Phys. 2012, 14, 3400–3407. [Google Scholar] [CrossRef]
  13. Da Silva, P.S.R.C.; Tavares, M.I.B. Solvent Effect on the Morphology of Lamellar Nanocomposites Based on HIPS. Mater. Res. 2015, 18, 191–195. [Google Scholar] [CrossRef] [Green Version]
  14. Cunha, A.D.P.C.B.; Tavares, M.I.B.; Silva, E.O.; Zaioncz, S. The Effect of Montmorillonite Clay on the Crystallinity of Poly(vinyl alcohol) Nanocomposites Obtained by Solution Intercalation and In Situ Polymerization. J. Nanosci. Nanotechnol. 2015, 15, 2814–2820. [Google Scholar] [CrossRef]
  15. Motaung, T.; Luyt, A.; Saladino, M.; Caponetti, E. Study of morphology, mechanical properties, and thermal degradation of polycarbonate-titania nanocomposites as function of titania crystalline phase and content. Polym. Compos. 2013, 34, 164–172. [Google Scholar] [CrossRef] [Green Version]
  16. Liaw, W.-C.; Cheng, Y.-L.; Liao, Y.-S.; Chen, C.-S.; Lai, S.-M. Complementary functionality of SiO2 and TiO2 in polyimide/silica–titania ternary hybrid nanocomposites. Polym. J. 2011, 43, 249–257. [Google Scholar] [CrossRef]
  17. Shishkovsky, I.V.; Scherbakov, V.I. Additive manufacturing of polymer composites with nano-titania inclusions. Laser Phys. Lett. 2021, 18, 066001. [Google Scholar] [CrossRef]
  18. Sanes, J.; Sánchez, C.; Pamies, R.; Avilés, M.-D.; Bermúdez, M.-D. Extrusion of Polymer Nanocomposites with Graphene and Graphene Derivative Nanofillers: An Overview of Recent Developments. Materials 2020, 13, 549. [Google Scholar] [CrossRef] [Green Version]
  19. Müller, C.M.; Laurindo, J.B.; Yamashita, F. Composites of thermoplastic starch and nanoclays produced by extrusion and thermopressing. Carbohydr. Polym. 2012, 89, 504–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Kickelbick, G. Hybrid Materials: Synthesis, Characterization, and Applications; John Wiley & Sons: Hoboken, NJ, USA, 2007. [Google Scholar]
  21. Almeida, A.D.S.; Tavares, M.I.B.; Da Silva, E.O.; Neto, R.P.C.; Moreira, L.A. Development of hybrid nanocomposites based on PLLA and low-field NMR characterization. Polym. Test. 2012, 31, 267–275. [Google Scholar] [CrossRef]
  22. Cheraghian, G. Effect of nano titanium dioxide on heavy oil recovery during polymer flooding. Pet. Sci. Technol. 2016, 34, 633–641. [Google Scholar] [CrossRef]
  23. Vidakis, N.; Petousis, M.; Maniadi, A.; Koudoumas, E.; Liebscher, M.; Tzounis, L. Mechanical Properties of 3D-Printed Acrylonitrile–Butadiene–Styrene TiO2 and ATO Nanocomposites. Polymers 2020, 12, 1589. [Google Scholar] [CrossRef] [PubMed]
  24. Vidakis, N.; Maniadi, A.; Petousis, M.; Vamvakaki, M.; Kenanakis, G.; Koudoumas, E. Mechanical and Electrical Properties Investigation of 3D-Printed Acrylonitrile–Butadiene–Styrene Graphene and Carbon Nanocomposites. J. Mater. Eng. Perform. 2020, 29, 1909–1918. [Google Scholar] [CrossRef]
  25. Nasu, A.; Otsubo, Y. Rheology and UV-protecting properties of complex suspensions of titanium dioxides and zinc oxides. J. Colloid Interface Sci. 2007, 310, 617–623. [Google Scholar] [CrossRef]
  26. Uddin, M.; Mondal, D.P.; Morris, C.; Lopez, T.; Diebold, U.; Gonzalez, R.D. An in vitro controlled release study of valproic acid encapsulated in a titania ceramic matrix. Appl. Surf. Sci. 2011, 257, 7920–7927. [Google Scholar] [CrossRef]
  27. Fujihara, K.; Kumar, A.; Jose, R.; Ramakrishna, S.; Uchida, S. Spray deposition of electrospun TiO2 nanorods for dye-sensitized solar cell. Nanotechnology 2007, 18, 365709. [Google Scholar] [CrossRef]
  28. Andronic, L.; Enesca, A. Black TiO2 Synthesis by Chemical Reduction Methods for Photocatalysis Applications. Front. Chem. 2020, 8, 5489. [Google Scholar] [CrossRef]
  29. Saccà, A.; Carbone, A.; Gatto, I.; Pedicini, R.; Freni, A.; Patti, A.; Passalacqua, E. Composites Nafion-titania membranes for Polymer Electrolyte Fuel Cell (PEFC) applications at low relative humidity levels: Chemical physical properties and electrochemical performance. Polym. Test. 2016, 56, 10–18. [Google Scholar] [CrossRef]
  30. Fiorati, A.; Bellingeri, A.; Punta, C.; Corsi, I.; Venditti, I. Silver Nanoparticles for Water Pollution Monitoring and Treatments: Ecosafety Challenge and Cellulose-Based Hybrids Solution. Polymers 2020, 12, 1635. [Google Scholar] [CrossRef]
  31. Pantalei, S.; Zampetti, E.; Macagnano, A.; Bearzotti, A.; Venditti, I.; Russo, M. Enhanced Sensory Properties of a Multichannel Quartz Crystal Microbalance Coated with Polymeric Nanobeads. Sensors 2007, 7, 2920–2928. [Google Scholar] [CrossRef] [Green Version]
  32. Fratoddi, I.; Cartoni, A.; Venditti, I.; Catone, D.; O’Keeffe, P.; Paladini, A.; Toschi, F.; Turchini, S.; Sciubba, F.; Testa, G.; et al. Gold nanoparticles functionalized by rhodamine B isothiocyanate: A new tool to control plasmonic effects. J. Colloid Interface Sci. 2018, 513, 10–19. [Google Scholar] [CrossRef]
  33. Sharpe, L.H. The Interfacial Interactions în Polymeric Composites; Akovali, G., Ed.; Kluwer: Dordrecht, The Netherlands, 1993. [Google Scholar]
  34. Jesson, D.; Watts, J.F. The Interface and Interphase in Polymer Matrix Composites: Effect on Mechanical Properties and Methods for Identification. Polym. Rev. 2012, 52, 321–354. [Google Scholar] [CrossRef]
  35. Wang, M.; Wang, Z.; Li, N.; Liao, J.; Zhao, S.; Wang, J.; Wang, S. Relationship between polymer–filler interfaces in separation layers and gas transport properties of mixed matrix composite membranes. J. Membr. Sci. 2015, 495, 252–268. [Google Scholar] [CrossRef]
  36. Drzal, L.T.; Rich, M.J.; Lloyd, P.F. Adhesion of Graphite Fibers to Epoxy Matrices: I. The Role of Fiber Surface Treatment. J. Adhes. 1983, 16, 1–30. [Google Scholar] [CrossRef]
  37. Berlin, A.A.; Volfson, S.t.A.; Enikilopian, N.S.; Negmatov, S.S. Principles of Polymer Composites; Akademic-Verlag: Berlin, Germany, 1985. [Google Scholar]
  38. Cassidy, P.E.; Yager, B.J. Coupling Agents As Adhesion Promoters în Reviews în Polymer Technology; Skeist, I., Ed.; Marcel Dekker Inc.: New York, NY, USA, 1972; Volume 1. [Google Scholar]
  39. Kim, J.-K.; Mai, Y.-W. Engineered Interfaces in Fiber Reinforced Composites; Elsevier BV: Amsterdam, The Netherlands, 1998. [Google Scholar]
  40. Fourche, G. An overview of the basic aspects of polymer adhesion. Part I: Fundamentals. Polym. Eng. Sci. 1995, 35, 957–967. [Google Scholar] [CrossRef]
  41. Morris, H.R.; Turner, J.F.; Munro, B.; Ryntz, R.A.; Treado, P.J. Chemical Imaging of Thermoplastic Olefin (TPO) Surface Architecture. Langmuir 1999, 15, 2961–2972. [Google Scholar] [CrossRef]
  42. Del Rio, F.W.; De Boer, M.; Knapp, J.A.; Reedy, E.D.; Clews, P.J.; Dunn, M. The role of van der Waals forces in adhesion of micromachined surfaces. Nat. Mater. 2005, 4, 629–634. [Google Scholar] [CrossRef]
  43. Lipatov, Y. Polymer Reinforcement; ChemTec Publishing: Scarborough, ON, Canada, 1995. [Google Scholar]
  44. Qin, R.-Y.; Schreiber, H. Adhesion at partially restructured polymer surfaces. Colloids Surf. A Physicochem. Eng. Asp. 1999, 156, 85–93. [Google Scholar] [CrossRef]
  45. Selvin, T.P.; Kuruvilla, J.; Sabu, T. Mechanical properties of titanium dioxide-filled polystyrene microcomposites. Mater. Lett. 2004, 58, 281–289. [Google Scholar] [CrossRef]
  46. De Armitt, C.; Rothon, R. Fillers and surface treatment. Plast. Addit. Compd. 2002, 4, 12–14. [Google Scholar] [CrossRef]
  47. Fronza, B.M.; Lewis, S.; Shah, P.K.; Barros, M.D.; Giannini, M.; Stansbury, J.W. Modification of filler surface treatment of composite resins using alternative silanes and functional nanogels. Dent. Mater. 2019, 35, 928–936. [Google Scholar] [CrossRef] [PubMed]
  48. Mozetič, M. Surface Modification to Improve Properties of Materials. Materials 2019, 12, 441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Chaudhari, S.; Shaikh, T.; Pandey, P. A Review on Polymer TiO2 Nanocomposites. Int. J. Eng. Res. Appl. 2013, 3, 1386–1391. [Google Scholar]
  50. Dalod, A.R.M.; Henriksen, L.; Grande, T.; Einarsrud, M.-A. Functionalized TiO2 nanoparticles by single-step hydrothermal synthesis: The role of the silane coupling agents. Beilstein J. Nanotechnol. 2017, 8, 304–312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Chen, X.; Mao, S.S. Synthesis of Titanium Dioxide (TiO2) Nanomaterials. J. Nanosci. Nanotechnol. 2006, 6, 906–925. [Google Scholar] [CrossRef]
  52. Tao, P.; Li, Y.; Rungta, A.; Viswanath, A.; Gao, J.; Benicewicz, B.; Siegel, R.W.; Schadler, L.S. TiO2 nanocomposites with high refractive index and transparency. J. Mater. Chem. 2011, 21, 18623–18629. [Google Scholar] [CrossRef]
  53. Díez-Pascual, A.M.; Díez-Vicente, A.L. Nano-TiO2 Reinforced PEEK/PEI Blends as Biomaterials for Load-Bearing Implant Applications. ACS Appl. Mater. Interfaces 2015, 7, 5561–5573. [Google Scholar] [CrossRef] [PubMed]
  54. Byranvand, M.M.; Kharat, A.N.; Fatholahi, L.; Beiranvand, Z.M. A review on synthesis of nano-TiO2 via different methods. JNS 2013, 3, 1–9. [Google Scholar] [CrossRef]
  55. Wang, Y.; He, Y.; Lai, Q.; Fan, M. Review of the progress in preparing nano TiO2: An important environmental engineering material. J. Environ. Sci. 2014, 26, 2139–2177. [Google Scholar] [CrossRef] [PubMed]
  56. Hanemann, T.; Szabó, D.V. Polymer-Nanoparticle Composites: From Synthesis to Modern Applications. Materials 2010, 3, 3468–3517. [Google Scholar] [CrossRef]
  57. Laine, R.M.; Choi, J.; Lee, I. Organic–Inorganic Nanocomposites with Completely Defined Interfacial Interactions. Adv. Mater. 2001, 13, 800–803. [Google Scholar] [CrossRef]
  58. Li, H.; Zhang, Z.; Ma, X.; Hu, M.; Wang, X.; Fan, P. Synthesis and characterization of epoxy resin modified with nano-SiO2 and γ-glycidoxypropyltrimethoxy silane. Surf. Coat. Technol. 2007, 201, 5269–5272. [Google Scholar] [CrossRef]
  59. Xu, X.; Li, B.; Lu, H.; Zhang, Z.; Wang, H. The interface structure of nano-SiO2/PA66 composites and its influence on material’s mechanical and thermal properties. Appl. Surf. Sci. 2007, 254, 1456–1462. [Google Scholar] [CrossRef]
  60. Li, X.; Cao, Z.; Zhang, Z.; Dang, H. Surface-modification in situ of nano-SiO2 and its structure and tribological properties. Appl. Surf. Sci. 2006, 252, 7856–7861. [Google Scholar] [CrossRef]
  61. Iijima, M.; Sato, N.; Lenggoro, I.W.; Kamiya, H. Surface modification of BaTiO3 particles by silane coupling agents in different solvents and their effect on dielectric properties of BaTiO3/epoxy composites. Colloids Surf. A Physicochem. Eng. Asp. 2009, 352, 88–93. [Google Scholar] [CrossRef]
  62. Balas, F.; Kokubo, T.; Kawashita, M.; Nakamura, T. Surface modification of organic polymers with bioactive titanium oxide without the aid of a silane-coupling agent. J. Mater. Sci. Mater. Electron. 2007, 18, 1167–1174. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Chen, H.; Wen, Y.; Yuan, Y.; Wu, W.; Liu, C. Tunable wettability of monodisperse core-shell nano-SiO2 modified with poly(methylhydrosiloxane) and allyl-poly(ethylene glycol). Colloids Surf. A Physicochem. Eng. Asp. 2014, 441, 16–24. [Google Scholar] [CrossRef]
  64. Plueddemann, E.P. Silane Coupling Agents, 2nd ed.; Plenum Press: New York, NY, USA, 1991. [Google Scholar]
  65. Yosomiya, R.; Morimoto, K.; Nakajima, A.; Ikada, Y.; Suzuki, T.; Dharan, C.K.H. Adhesion and Bonding in Composites. J. Eng. Ind. 1991, 113, 117. [Google Scholar] [CrossRef]
  66. Konakanchi, A.; Alla, R.M.; Guduri, V. Silane Coupling Agents—Benevolent Binders in Composites. Trends Biomater. Artif. Organs 2017, 31, 108–113. [Google Scholar]
  67. Shokoohi, S.; Arefazar, A.; Khosrokhavar, R. Silane Coupling Agents in Polymer-based Reinforced Composites: A Review. J. Reinf. Plast. Compos. 2008, 27, 473–485. [Google Scholar] [CrossRef]
  68. Zhang, Y.; Fang, F.; Wang, C.; Wang, L.; Wang, X.; Chu, X.; Li, J.; Fang, X.; Wei, Z.; Wang, X. Hydrophobic modification of ZnO nanostructures surface using silane coupling agent. Polym. Compos. 2014, 35, 1204–1211. [Google Scholar] [CrossRef]
  69. Mallakpour, S.; Madani, M. The effect of the coupling agents KH550 and KH570 on the nanostructure and interfacial interaction of zinc oxide/chiral poly(amide–imide) nanocomposites containing l-leucine amino acid moieties. J. Mater. Sci. 2014, 49, 5112–5118. [Google Scholar] [CrossRef]
  70. Liu, Y.-L.; Su, Y.-H.; Lai, J.-Y. In situ crosslinking of chitosan and formation of chitosan–silica hybrid membranes with using γ-glycidoxypropyltrimethoxysilane as a crosslinking agent. Polymer 2004, 45, 6831–6837. [Google Scholar] [CrossRef]
  71. Lu, Y.; Zhou, S.; Wu, L. De-Agglomeration and Dispersion Behavior of TiO2 Nanoparticles in Organic Media Using 3-Methacryloxypropyltrimethoxysilane as a Surface Modifier. J. Dispers. Sci. Technol. 2012, 33, 497–505. [Google Scholar] [CrossRef]
  72. Dang, Z.-M.; Xia, Y.-J.; Zha, J.-W.; Yuan, J.; Bai, J. Preparation and dielectric properties of surface modified TiO2 /silicone rubber nanocomposites. Mater. Lett. 2011, 65, 3430–3432. [Google Scholar] [CrossRef]
  73. Xu, Q.F.; Liu, Y.; Lin, F.-J.; Mondal, B.; Lyons, A.M. Superhydrophobic TiO2 –Polymer Nanocomposite Surface with UV-Induced Reversible Wettability and Self-Cleaning Properties. ACS Appl. Mater. Interfaces 2013, 5, 8915–8924. [Google Scholar] [CrossRef] [PubMed]
  74. Rahim-Abadi, M.M.; Mahdavian, A.R.; Gharieh, A.; Salehi-Mobarakeh, H. Chemical modification of TiO2 nanoparticles as an effective way for encapsulation in polyacrylic shell via emulsion polymerization. Prog. Org. Coat. 2015, 88, 310–315. [Google Scholar] [CrossRef]
  75. Toh-Ae, P.; Junhasavasdikul, B.; Lopattananon, N.; Sahakaro, K. Surface Modification of TiO2 Nanoparticles by Grafting with Silane Coupling Agent. Adv. Mater. Res. 2013, 844, 276–279. [Google Scholar] [CrossRef]
  76. Ambrósio, J.D.; Balarim, C.V.M.; De Carvalho, G.B. Preparation, characterization, and mechanical/tribological properties of polyamide 11/Titanium dioxide nanocomposites. Polym. Compos. 2014, 37, 1415–1424. [Google Scholar] [CrossRef]
  77. Sabzi, M.; Mirabedini, S.M.; Zohuriaan-Mehr, M.J.; Atai, M. Surface modification of TiO2 nano-particles with silane coupling agent and investigation of its effect on the properties of polyurethane composite coating. Prog. Org. Coat. 2009, 65, 222–228. [Google Scholar] [CrossRef]
  78. Rusu, G.; Rusu, E. Nylon 6/TiO2Composites by in situ Anionic Ring-Opening Polymerization of ϵ-Caprolactam: Synthesis, Characterization, and Properties. Int. J. Polym. Anal. Charact. 2011, 16, 561–583. [Google Scholar] [CrossRef]
  79. Zapata, P.A.; Palza, H.; Delgado, K.; Rabagliati, F.M. Novel antimicrobial polyethylene composites prepared by metallocenic in situ polymerization with TiO2-based nanoparticles. J. Polym. Sci. Part A Polym. Chem. 2012, 50, 4055–4062. [Google Scholar] [CrossRef]
  80. Nguyen, V.G.; Thai, H.; Mai, D.H.; Tran, H.T.; Tran, D.L.; Vu, M.T. Effect of titanium dioxide on the properties of polyethylene/ TiO2 nanocomposites. Compos. Part B Eng. 2013, 45, 1192–1198. [Google Scholar] [CrossRef]
  81. Dzžunuzović, E.; Marinović-Cincović, M.; Vuković, J.; Jeremić, K.; Nedeljkovic, J. Thermal properties of PMMA/ TiO2 nanocomposites prepared byin-situbulk polymerization. Polym. Compos. 2009, 30, 737–742. [Google Scholar] [CrossRef]
  82. Yuvaraj, H.; Kim, W.S.; Kim, J.T.; Kang, I.P.; Gal, Y.-S.; Kim, S.W.; Lim, K.T. Synthesis of Poly(methyl methacrylate) Encapsulated TiO2 Nanocomposite Particles in Supercritical CO2. Mol. Cryst. Liq. Cryst. 2009, 514, 355–365. [Google Scholar] [CrossRef]
  83. Qi, Y.; Xiang, B.; Tan, W.; Zhang, J. Hydrophobic surface modification of TiO2 nanoparticles for production of acrylonitrile-styrene-acrylate terpolymer/ TiO2 composited cool materials. Appl. Surf. Sci. 2017, 419, 213–223. [Google Scholar] [CrossRef]
  84. Weng, C.-C.; Wei, K.-H. Selective Distribution of Surface-Modified TiO2 Nanoparticles in Polystyrene-b-poly (Methyl Methacrylate) Diblock Copolymer. Chem. Mater. 2003, 15, 2936–2941. [Google Scholar] [CrossRef]
  85. Yufei, C.; Zhichao, L.; Junyan, T.; Qingyu, Z.; Yang, H. Characteristics and Properties of TiO2 /EP-PU Composite. J. Nanomater. 2015, 2015, 1–7. [Google Scholar] [CrossRef] [Green Version]
  86. Zhao, J.; Milanova, M.; Warmoeskerken, M.M.; Dutschk, V. Surface modification of TiO2 nanoparticles with silane coupling agents. Colloids Surf. A Physicochem. Eng. Asp. 2012, 413, 273–279. [Google Scholar] [CrossRef]
  87. Wanag, A.; Sienkiewicz, A.; Rokicka-Konieczna, P.; Kusiak-Nejman, E.; Morawski, A.W. Influence of modification of titanium dioxide by silane coupling agents on the photocatalytic activity and stability. J. Environ. Chem. Eng. 2020, 8, 103917. [Google Scholar] [CrossRef]
  88. Mallakpour, S.; Barati, A. Efficient preparation of hybrid nanocomposite coatings based on poly(vinyl alcohol) and silane coupling agent modified TiO2 nanoparticles. Prog. Org. Coat. 2011, 71, 391–398. [Google Scholar] [CrossRef]
  89. Shakeri, A.; Yip, D.; Badv, M.; Imani, S.M.; Sanjari, M.; Didar, T.F. Self-Cleaning Ceramic Tiles Produced via Stable Coating of TiO2 Nanoparticles. Materials 2018, 11, 1003. [Google Scholar] [CrossRef] [Green Version]
  90. Klaysri, R.; Tubchareon, T.; Praserthdam, P. One-step synthesis of amine-functionalized TiO2 surface for photocatalytic decolorization under visible light irradiation. J. Ind. Eng. Chem. 2017, 45, 229–236. [Google Scholar] [CrossRef]
  91. Chen, Q.; Yakovlev, N.L. Adsorption and interaction of organosilanes on TiO2 nanoparticles. Appl. Surf. Sci. 2010, 257, 1395–1400. [Google Scholar] [CrossRef]
  92. Dinari, M.; Haghighi, A. Surface modification of TiO2 nanoparticle by three dimensional silane coupling agent and preparation of polyamide/modified-TiO2 nanocomposites for removal of Cr (VI) from aqueous solutions. Prog. Org. Coat. 2017, 110, 24–34. [Google Scholar] [CrossRef]
  93. Caris, C.; Kuijpers, R.; Van Herk, A.M.; German, A.L. Kinetics of (CO)polymerizations at the surface of inorganic submicron particles in emulsion-like systems. Makromol. Chem. Macromol. Symp. 1990, 35-36, 535–548. [Google Scholar] [CrossRef] [Green Version]
  94. Erdem, B.; Sudol, E.D.; Dimonie, V.L.; El-Aasser, M.S. Encapsulation of inorganic particles via miniemulsion polymerization. II. Preparation and characterization of styrene miniemulsion droplets containing TiO2 particles. J. Polym. Sci. Part A Polym. Chem. 2000, 38, 4431–4440. [Google Scholar] [CrossRef]
  95. Rong, M.Z.; Zhang, M.Q.; Wang, H.B.; Zeng, H.M. Surface modification of magnetic metal nanoparticles through irradiation graft polymerization. Appl. Surf. Sci. 2002, 200, 76–93. [Google Scholar] [CrossRef]
  96. Yang, M.; Dan, Y. Preparation and characterization of poly(methyl methacrylate)/titanium oxide composite particles. Colloid Polym. Sci. 2005, 284, 243–250. [Google Scholar] [CrossRef]
  97. Milanesi, F.; Cappelletti, G.; Annunziata, R.; Bianchi, C.L.; Meroni, D.; Ardizzone, S. Siloxane− TiO2 Hybrid Nanocomposites. The Structure of the Hydrophobic Layer. J. Phys. Chem. C 2010, 114, 8287–8293. [Google Scholar] [CrossRef]
  98. Xiang, B.; Jiang, G.; Zhang, J. Surface modification of TiO2 nanoparticles with silane coupling agent for nanocomposite with poly(butyl acrylate). Plast. Rubber Compos. 2015, 44, 148–154. [Google Scholar] [CrossRef]
  99. Wang, C.; Mao, H.; Wang, C.; Fu, S. Dispersibility and Hydrophobicity Analysis of Titanium Dioxide Nanoparticles Grafted with Silane Coupling Agent. Ind. Eng. Chem. Res. 2011, 50, 11930–11934. [Google Scholar] [CrossRef]
  100. Godnjavec, J.; Znoj, B.; Vince, J.; Steinbucher, M.; Žnidaršič, A.; Venturini, P. Stabilization of rutile TiO2 nanoparticles with Glymo in polyacrylic clear coating. Mater. Tehnol. 2012, 46, 19–24. [Google Scholar]
  101. Yang, C.; Yang, C. Preparation of TiO2 particles and surface silanization modification for electronic ink. J. Mater. Sci. Mater. Electron. 2014, 25, 3285–3289. [Google Scholar] [CrossRef]
  102. Tangchantra, N.; Kruenate, J.; Aumnate, C.; Sooksomsong, T. The Effect of Surface Modification of TiO2 on Mechanical Properties of Polyethylene Composite Film. Adv. Mater. Res. 2010, 93-94, 300–303. [Google Scholar] [CrossRef]
  103. Fujishima, A.; Rao, T.N.; Tryk, D.A. Titanium dioxide photocatalysis. J. Photochem. Photobiol. C Photochem. Rev. 2000, 1, 1–21. [Google Scholar] [CrossRef]
  104. Herrmann, J.-M. Heterogeneous photocatalysis: Fundamentals and applications to the removal of various types of aqueous pollutants. Catal. Today 1999, 53, 115–129. [Google Scholar] [CrossRef]
  105. Nada, A.; Barakat, M.; Hamed, H.; Mohamed, N.; Veziroglu, T. Studies on the photocatalytic hydrogen production using suspended modified TiO2 photocatalysts. Int. J. Hydrogen Energy 2005, 30, 687–691. [Google Scholar] [CrossRef]
  106. Andronic, L.; Enesca, A.; Cazan, C.; Visa, M. TiO2–active carbon composites for wastewater photocatalysis. J. Sol.-Gel. Sci. Technol. 2014, 71, 396–405. [Google Scholar] [CrossRef]
  107. Sharma, S.K.; Vishwas, M.; Rao, K.N.; Mohan, S.; Reddy, D.S.; Gowda, K. Structural and optical investigations of TiO2 films deposited on transparent substrates by sol–gel technique. J. Alloys Compd. 2009, 471, 244–247. [Google Scholar] [CrossRef]
  108. Bayarri, B.; Giménez, J.; Curcó, D.; Esplugas, S. Photocatalytic degradation of 2,4-dichlorophenol by TiO2 /UV: Kinetics, actinometries and models. Catal. Today 2005, 101, 227–236. [Google Scholar] [CrossRef]
  109. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C Photochem. Rev. 2008, 9, 1–12. [Google Scholar] [CrossRef]
  110. Shimizu, N.; Ninomiya, K.; Ogino, C.; Rahman, M.M. Potential uses of titanium dioxide in conjunction with ultrasound for improved disinfection. Biochem. Eng. J. 2010, 48, 416–423. [Google Scholar] [CrossRef]
  111. Tuan, N.M.; Nha, N.T.; Tuyen, N.H. Low-temperature synthesis of nano-TiO2 anatase on nafion membrane for using on DMFC. J. Phys. Conf. Ser. 2009, 187, 012040. [Google Scholar] [CrossRef] [Green Version]
  112. Armstrong, A.R.; Armstrong, G.; Canales-Vazquez, J.; García, R.; Bruce, P.G. Lithium-Ion Intercalation into TiO2-B Nanowires. Adv. Mater. 2005, 17, 862–865. [Google Scholar] [CrossRef]
  113. Richards, B.S.; Cotter, J.E.; Honsberg, C.B. Enhancing the surface passivation of TiO2 coated silicon wafers. Appl. Phys. Lett. 2002, 80, 1123–1125. [Google Scholar] [CrossRef]
  114. Pinto, D.; Bernardo, L.; Amaro, A.; Lopes, S. Mechanical properties of epoxy nanocomposites using titanium dioxide as reinforcement—A review. Constr. Build. Mater. 2015, 95, 506–524. [Google Scholar] [CrossRef]
  115. Sahu, M.; Satapathy, A. Thermal Characteristics of Polypropylene Composites Filled with TiO2. In Proceedings of the ASEAI 2014, Bangkok, Thailand, 11–13 April 2013; Volume 49, pp. 1747–1752. [Google Scholar]
  116. Peng, X.; Ding, E.; Xue, F. In situ synthesis of TiO2/polyethylene terephthalate hybrid nanocomposites at low temperature. Appl. Surf. Sci. 2012, 258, 6564–6570. [Google Scholar] [CrossRef]
  117. Mourad, A.-H.I.; Mozumder, M.S.; Mairpady, A.; Pervez, H.; Kannuri, U.M. On the Injection Molding Processing Parameters of HDPE-TiO2 Nanocomposites. Materials 2017, 10, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zhil’Tsova, T.; Oliveira, M.; Ferreira, J. Relative influence of injection molding processing conditions on HDPE acetabular cups dimensional stability. J. Mater. Process. Technol. 2009, 209, 3894–3904. [Google Scholar] [CrossRef]
  119. Serrano, C.; Cerrada, M.; Fernandez-Garcia, M.; Ressia, J.; Valles, E.M. Rheological and structural details of biocidal iPP-TiO2 nanocomposites. Eur. Polym. J. 2012, 48, 586–596. [Google Scholar] [CrossRef] [Green Version]
  120. Umek, P.; Huskić, M.; Škapin, A.S.; Florjančič, U.; Zupančič, B.; Emri, I.; Arčon, D. Structural and mechanical properties of polystyrene nanocomposites with 1D titanate nanostructures prepared by an extrusion process. Polym. Compos. 2008, 30, 1318–1325. [Google Scholar] [CrossRef]
  121. Somani, P.R.; Marimuthu, R.; Mulik, U.; Sainkar, S.; Amalnerkar, D. High piezoresistivity and its origin in conducting polyaniline/TiO2 composites. Synth. Met. 1999, 106, 45–52. [Google Scholar] [CrossRef]
  122. Feng, W.; Sun, E.; Fujii, A.; Wu, H.; Niihara, K.; Yoshino, K. Synthesis and Characterization of Photoconducting Polyaniline-TiO2 Nanocomposite. Bull. Chem. Soc. Jpn. 2000, 73, 2627–2633. [Google Scholar] [CrossRef]
  123. Xia, H.; Wang, Q. Ultrasonic Irradiation: A Novel Approach To Prepare Conductive Polyaniline/Nanocrystalline Titanium Oxide Composites. Chem. Mater. 2002, 14, 2158–2165. [Google Scholar] [CrossRef]
  124. Alghamdi, M.N. Titanium Dioxide Reinforced Polypropylene Composites: Preparation and Characterization PP-TiO2 Composites. Int. J. Eng. Res. Technol. 2016, 5, 633–637. [Google Scholar]
  125. Vladuta, C.; Andronic, L.; Duta, A. Effect of TiO2, nanoparticles on the interface in the PET-rubber composites. J. Nanosci. Nanotechnol. 2010, 10, 2518–2526. [Google Scholar] [CrossRef] [PubMed]
  126. Mikešová, J.; Šlouf, M.; Gohs, U.; Popelková, D.; Vacková, T.; Vu, N.H.; Kratochvíl, J.; Zhigunov, A. Nanocomposites of polypropylene/titanate nanotubes: Morphology, nucleation effects of nanoparticles and properties. Polym. Bull. 2014, 71, 795–818. [Google Scholar] [CrossRef]
  127. Sreekumar, P.; Al-Harthi, M.A.; De, S. Reinforcement of starch/polyvinyl alcohol blend using nano-titanium dioxide. J. Compos. Mater. 2012, 46, 3181–3187. [Google Scholar] [CrossRef]
  128. Bora, M.Ö.; Çoban, O.; Avcu, E.; Fidan, S.; Sınmazçelik, T. The effect of TIO2 filler content on the mechanical, thermal, and tribological properties of TiO2 /PPS composites. Polym. Compos. 2013, 34, 1591–1599. [Google Scholar] [CrossRef]
  129. Saluja, P.S.; Tiwari, J.K.; Gupta, G. Preparation and Thermal Behaviour of Polyester Composite Filled with TiO2. Int. Res. J. Eng. Technol. 2017, 4, 3135–3141. [Google Scholar]
  130. Byrne, M.T.; McCarthy, J.E.; Bent, M.; Blake, R.; Gun’Ko, Y.K.; Horvath, E.; Konya, Z.; Kukovecz, A.; Kiricsi, I.; Coleman, J.N. Chemical functionalisation of titania nanotubes and their utilisation for the fabrication of reinforced polystyrene composites. J. Mater. Chem. 2007, 17, 2351–2358. [Google Scholar] [CrossRef]
  131. Dong, Y.; Gui, Z.; Hu, Y.; Wu, Y.; Jiang, S. The influence of titanate nanotube on the improved thermal properties and the smoke suppression in poly(methyl methacrylate). J. Hazard. Mater. 2012, 209-210, 34–39. [Google Scholar] [CrossRef] [PubMed]
  132. Alexandru, M. On the morphology and potential application of polydimethylsiloxane-silica-titania composites. Express Polym. Lett. 2011, 5, 188–196. [Google Scholar] [CrossRef]
  133. Saritha, A.; Joseph, K.; Boudenne, A.; Thomas, S. Mechanical, thermophysical, and diffusion properties of TiO2-filled chlorobutyl rubber composites. Polym. Compos. 2011, 32, 1681–1687. [Google Scholar] [CrossRef]
  134. Manap, A.; Mahalingam, S.; Vaithylingam, R.; Abdullah, H. Mechanical, thermal and morphological properties of thermoplastic polyurethane composite reinforced by multi-walled carbon nanotube and titanium dioxide hybrid fillers. Polym. Bull. 2020, 1–18. [Google Scholar] [CrossRef]
  135. Esthappan, S.K.; Kuttappan, S.K.; Joseph, R. Thermal and mechanical properties of polypropylene/titanium dioxide nanocomposite fibers. Mater. Des. 2012, 37, 537–542. [Google Scholar] [CrossRef]
  136. Awang, M.; Mohd, W.R.W.; Sarifuddin, N. Study the effects of an addition of titanium dioxide (TiO2) on the mechanical and thermal properties of polypropylene-rice husk green composites. Mater. Res. Express 2019, 6, 075311. [Google Scholar] [CrossRef]
  137. Huang, J.; Lu, X.; Zhang, N.; Yang, L.; Yan, M.; Liu, H.; Zhang, G.; Qu, J. Study on the properties of nano-TiO2 /polybutylene succinate composites prepared by vane extruder. Polym. Compos. 2013, 35, 53–59. [Google Scholar] [CrossRef]
  138. Bragaglia, M.; Cherubini, V.; Nanni, F. PEEK-TiO2 composites with enhanced UV resistance. Compos. Sci. Technol. 2020, 199, 108365. [Google Scholar] [CrossRef]
  139. Farhoodi, M.; Dadashi, S.; Mousavi, S.M.A.; Sotudeh-Gharebagh, R.; Emam-Djomeh, Z.; Oromiehie, A.; Hemmati, F. Influence of TiO2 Nanoparticle Filler on the Properties of PET and PLA Nanocomposites. Polym. Korea 2012, 36, 745–755. [Google Scholar] [CrossRef] [Green Version]
  140. Lu, X.-L.; Lü, X.-Q.; Wang, J.-Y.; Sun, Z.-J.; Tong, Y.-X. Preparation and shape memory properties of TiO2/PLCL biodegradable polymer nanocomposites. Trans. Nonferrous Met. Soc. China 2013, 23, 120–127. [Google Scholar] [CrossRef]
  141. Alberton, J.; Martelli, S.M.; Fakhouri, F.M.; Soldi, V. Mechanical and moisture barrier properties of titanium dioxide nanoparticles and halloysite nanotubes reinforced polylactic acid (PLA). IOP Conf. Ser. Mater. Sci. Eng. 2014, 64, 012010. [Google Scholar] [CrossRef] [Green Version]
  142. Zhang, H.; Huang, J.; Yang, L.; Chen, R.; Zou, W.; Lin, X.; Qu, J. Preparation, characterization and properties of PLA/TiO2 nanocomposites based on a novel vane extruder. RSC Adv. 2015, 5, 4639–4647. [Google Scholar] [CrossRef]
  143. Zhuang, W.; Liu, J.; Zhang, J.H.; Hu, B.X.; Shen, J. Preparation, characterization, and properties of TiO2/PLA nanocomposites by in situ polymerization. Polym. Compos. 2009, 30, 1074–1080. [Google Scholar] [CrossRef]
  144. Wetzel, B.; Rosso, P.; Haupert, F.; Friedrich, K. Epoxy nanocomposites—Fracture and toughening mechanisms. Eng. Fract. Mech. 2006, 73, 2375–2398. [Google Scholar] [CrossRef]
  145. Chatterjee, A.; Islam, M.S. Fabrication and characterization of TiO2–epoxy nanocomposite. Mater. Sci. Eng. A 2008, 487, 574–585. [Google Scholar] [CrossRef]
  146. Huang, K.S.; Nien, Y.H.; Chen, J.S.; Shieh, T.R.; Chen, J.W. Synthesis and properties of epoxy/TiO2 composite materials. Polym. Compos. 2006, 27, 195–200. [Google Scholar] [CrossRef]
  147. Al-Turaif, H.A. Effect of nano TiO2 particle size on mechanical properties of cured epoxy resin. Prog. Org. Coat. 2010, 69, 241–246. [Google Scholar] [CrossRef]
  148. Papanicolaou, G.; Kontaxis, L.; Manara, A. Viscoelastic behaviour and modelling of nano and micro TiO2 powder-epoxy resin composites. Ciênc. Tecnol. Mater. 2016, 28, 138–146. [Google Scholar] [CrossRef]
  149. Salehian, H.; Jahromi, S.A.J. Effect of titanium dioxide nanoparticles on mechanical properties of vinyl ester-based nanocomposites. J. Compos. Mater. 2015, 49, 2365–2373. [Google Scholar] [CrossRef]
  150. Meera, A.P.; Said, S.; Grohens, Y.; Luyt, A.S.; Thomas, S. Tensile Stress Relaxation Studies of TiO2 and Nanosilica Filled Natural Rubber Composites. Ind. Eng. Chem. Res. 2009, 48, 3410–3416. [Google Scholar] [CrossRef]
  151. Ochigbo, S.S.; Luyt, A.S. Mechanical and Morphological Properties of Films Based on Ultrasound Treated Titanium Dioxide Dispersion/Natural Rubber Latex. Int. J. Compos. Mater. 2011, 1, 7–13. [Google Scholar] [CrossRef] [Green Version]
  152. Hayeemasae, N.; Rathnayake, W.; Ismail, H. Nano-sized TiO2-reinforced natural rubber composites prepared by latex compounding method. J. Vinyl Addit. Technol. 2017, 23, 200–209. [Google Scholar] [CrossRef]
  153. Datta, J.; Kosiorek, P.; Włoch, M. Effect of high loading of titanium dioxide particles on the morphology, mechanical and thermo-mechanical properties of the natural rubber-based composites. Iran. Polym. J. 2016, 25, 1021–1035. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Representation of the interphase between matrix and fillers.
Figure 1. Representation of the interphase between matrix and fillers.
Polymers 13 02017 g001
Figure 2. The formation of the interface between matrix and filler.
Figure 2. The formation of the interface between matrix and filler.
Polymers 13 02017 g002
Figure 3. The interaction between the coupling agent molecule and the filler.
Figure 3. The interaction between the coupling agent molecule and the filler.
Polymers 13 02017 g003
Figure 4. The factors that influence the mechanical properties of composites materials.
Figure 4. The factors that influence the mechanical properties of composites materials.
Polymers 13 02017 g004
Figure 5. Applications of the TiO2 nanocomposites.
Figure 5. Applications of the TiO2 nanocomposites.
Polymers 13 02017 g005
Table 1. Surface modification of TiO2 nanoparticles.
Table 1. Surface modification of TiO2 nanoparticles.
Modification Agent of TiO2 SurfaceChemical Structure Polymer–TiO2 NanocompositeRef
3-(trimethoxysilyl)propyl methacrylate, KH–570 Polymers 13 02017 i001silicone rubber–TiO2 nanocomposite[72]
fluoro silaneH3Si–FHDPE–TiO2 nanocomposite[73]
glycidyl methacrylate Polymers 13 02017 i002methyl methacrylate–butyl acrylate/dimethylaminoethyl methacrylate–butyl acrylate–acrylic acid–TiO2 nanoparticles [74]
bis-(3-triethoxysilylpropyl) tetrasulfide (TESPT) Polymers 13 02017 i003rubber–TiO2 nanocomposite[75]
3-amino propyl trimethoxy silane Polymers 13 02017 i004PA11–TiO2 nanocomposite;PU-TiO2 composites;[76,77]
3-amino propyl triethoxy silane Polymers 13 02017 i005nylon 6/TiO2 composites;PS–TiO2 microcompositespolyurethane–TiO2 composites;polyamide–TiO2 nanocomposites[45,77,78]
hexadecyl trimethoxy silane Polymers 13 02017 i006PE–TiO2 nanocomposite[79]
vinyl trimethoxy silane (VTMS) Polymers 13 02017 i007LDPE–TiO2 nanocomposite[80]
6-palmitate ascorbic acid Polymers 13 02017 i008PMMA–TiO2 nanocomposite[81]
3-methacryloxy propyl trimethoxy silane Polymers 13 02017 i009PMMA–TiO2 nanocomposite;acrylonitrile–styrene-acrylate terpolymer–TiO2 composite;PS-b-PMMA–TiO2 nanocomposite[82,83]
cetyl trimethylammonium chloride (TMAC) amphiphilics Polymers 13 02017 i010PS-b-PMMA–TiO2 nanocomposite[84]
isopropyl tri(dioctylpyrophosphate) titanate (TCA201) Polymers 13 02017 i011EP-PU/TiO2 composite[85]
3-isocyanato propyl trimethoxy silane Polymers 13 02017 i012polymer–TiO2 [86]
Table 2. Some properties of TiO2 and applications.
Table 2. Some properties of TiO2 and applications.
ApplicationProperties
PhotocatalysisParticularly in anatase from under ultraviolet light
Self-cleaning and anti-fogging glassSpiked with nitrogen ions or droplet with metal oxides under UV–visible light
Hydrolysis catalystSuper hydrophilicity, deodorizing, sterilizing, anti-fouling; chemical resistance
Dye-sensitized solar cellsStrong oxidative potential for develop OH radicals
Pigments, opacifiers, cosmetic, UV absorberBrightness, high reflective index, high reflective optical, perfect white, opacity, nontoxic to human life
Table 4. Advantages of polymer nanocomposites over conventional polymer composites.
Table 4. Advantages of polymer nanocomposites over conventional polymer composites.
Polymer NanocompositesConventional Polymer Composites
Fillers separation are in nm, and properties will be affected by size effects of nanofiller;
Small amounts of TiO2 filler are enough (less than 10%) to achieve desired properties;
Properties are obtained without sacrificing the inherent properties of the polymer or adding excessive weight;
Improvements in properties even in low amount is due to nanosized of TiO2 filler and interphase region;
Using nanosized particles can reduce the likelihood of finding defects, such as grain boundaries, voids, dislocations, and imperfections.
Fillers are separated in μm, and there is not that much of size effect;
High concentrations should be needed as compared to nanofiller case;
Fillers can unfavorably impact other benefits of polymers, such as appearance, ductility and toughness;
There is not that much improvement in properties even for a large amount of fillers;
It is difficult even observed in conventional polymer composites.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cazan, C.; Enesca, A.; Andronic, L. Synergic Effect of TiO2 Filler on the Mechanical Properties of Polymer Nanocomposites. Polymers 2021, 13, 2017. https://doi.org/10.3390/polym13122017

AMA Style

Cazan C, Enesca A, Andronic L. Synergic Effect of TiO2 Filler on the Mechanical Properties of Polymer Nanocomposites. Polymers. 2021; 13(12):2017. https://doi.org/10.3390/polym13122017

Chicago/Turabian Style

Cazan, Cristina, Alexandru Enesca, and Luminita Andronic. 2021. "Synergic Effect of TiO2 Filler on the Mechanical Properties of Polymer Nanocomposites" Polymers 13, no. 12: 2017. https://doi.org/10.3390/polym13122017

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