Next Article in Journal
Effect of Sulfonation Degree and PVDF Content on the Structure and Transport Properties of SPEEK/PVDF Blend Membranes
Previous Article in Journal
Highly Efficient Polydopamine-coated Poly(methyl methacrylate) Nanofiber Supported Platinum–Nickel Bimetallic Catalyst for Formaldehyde Oxidation at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Effect of Nanofillers on the Functional Properties of Biopolymer-Based Films: A Review

1
Institute of Chemistry, University of Agriculture in Cracow, Balicka Street 122, PL-30-149 Kraków, Poland
2
Department of Animal Products Processing, University of Agriculture, Balicka Street 122, PL-30-149 Kraków, Poland
3
Department of Chemistry and Biochemistry, Faculty of AgriSciences, Mendel University in Brno, Zemedelska 1, CZ-613 00 Brno, Czech Republic
4
Central European Institute of Technology, Brno University of Technology, Purkynova 123, CZ-612 00 Brno, Czech Republic
*
Author to whom correspondence should be addressed.
Polymers 2019, 11(4), 675; https://doi.org/10.3390/polym11040675
Submission received: 21 March 2019 / Revised: 9 April 2019 / Accepted: 9 April 2019 / Published: 12 April 2019

Abstract

:
Waste from non-degradable plastics is becoming an increasingly serious problem. Therefore, more and more research focuses on the development of materials with biodegradable properties. Bio-polymers are excellent raw materials for the production of such materials. Bio-based biopolymer films reinforced with nanostructures have become an interesting area of research. Nanocomposite films are a group of materials that mainly consist of bio-based natural (e.g., chitosan, starch) and synthetic (e.g., poly(lactic acid)) polymers and nanofillers (clay, organic, inorganic, or carbon nanostructures), with different properties. The interaction between environmentally friendly biopolymers and nanofillers leads to the improved functionality of nanocomposite materials. Depending on the properties of nanofillers, new or improved properties of nanocomposites can be obtained such as: barrier properties, improved mechanical strength, antimicrobial, and antioxidant properties or thermal stability. This review compiles information about biopolymers used as the matrix for the films with nanofillers as the active agents. Particular emphasis has been placed on the influence of nanofillers on functional properties of biopolymer films and their possible use within the food industry and food packaging systems. The possible applications of those nanocomposite films within other industries (medicine, drug and chemical industry, tissue engineering) is also briefly summarized.

1. Introduction

Nanocomposite films are a new generation of packaging materials with combination of bio-based polymer and fillers that have at least one nanometer scale dimension. In nanocomposite films, the biopolymer acts as a matrix, while the nanofillers are dispersed therein to improve the functional properties. Nanocomposites have a set of improved properties, such as mechanical, antimicrobial, or physical properties. These properties do not occur naturally in the biopolymers themselves, therefore they are gained due to addition of nanocomposites [1]. The components of biopolymer films are characterized by high availability and good biodegradability. However, one-component films have relatively poor mechanical properties and high water vapor permeability, which may cause them not to be of sufficient quality to be used as packaging materials. This can be reduced by mixing two biopolymers with one another [2]. Currently, a novel method to improve the properties of biopolymer films is the use of nanofillers, which can fulfill not only the reinforcing function but also could act as an active ingredient. In recent years, the concept of active agents for biopolymer films has received much more attention. Such active ingredients in biopolymer films can extend the shelf-life of food products, through exhibiting antimicrobial and/or antioxidant activities [3,4]. The preservative effect of biopolymer films with nanofillers have been reported for the whole spectrum of food products including vegetables and fruits, mushrooms, dairy products, meat and meat products, and fish and other seafood [5,6,7,8,9,10,11,12,13,14,15]. The development of nanotechnology has led to the design of nanocomposite film materials in which nanofillers play an active role. The aim of this review is to summarize recent advances and achievements regarding the addition of nanofillers—such as clay, metals, metal oxides, polymer nanoparticles, nanocellulose, etc.—into biopolymer films with a focus on their functional properties and a possible use in food packaging systems. Additionally, this review shows recent applications of biopolymer films with nanofillers on wound dressing, drug, and enzyme delivery systems and tissue engineering.

2. Types of Biopolymers and Nanofillers

New trends and perspectives in nanotechnology, have facilitated the path to use nanofillers as active agents in the packaging industry. This solution applies to the production of biodegradable packaging material, which is based on biopolymers and nanofillers. This type of nanocomposite material is not capable of replacing synthetic packaging materials, because it has many disadvantages, i.e., weak mechanical properties, too high hydrophilicity, and susceptibility to decomposition. However, they are interesting alternatives to plastic materials and can be used for example in areas where plastic recovery is not economically feasible [16].
Currently, there is a growing interest in the combination of biopolymer and nanofiller-based materials. Nanofillers can have different shapes and sizes, but their individual particles size is, by definition of the nanomaterials, below 100 nm [1]. The size of nanofillers is very beneficial for nanocomposite materials because they are based on a large surface area which leads to a large interphase or boundary area between the biopolymer matrix and nanofiller. Due to such interaction, the biopolymer matrix is modified, which could contribute to the improvement of mechanical, thermal, and barrier properties of bionanocomposite materials [17].

2.1. Types of Biopolymers Matrix

Bio-based polymers, which are considered the matrix base during preparation of nanocomposite films, have many advantages. They are renewable, biodegradable, multi-functional, and biocompatible [18].
Bio-based polymers can be classified into two major categories:
  • Natural bio-based polymers, including:
    -
    polymers extracted from agricultural resources:
    polysaccharides:
    • neutral: e.g., cellulose, hemicellulose, starch
    • cationic: e.g., chitin, chitosan
    • anionic: alginic acid, hyaluronic acid
    • of bacterial origin: e.g., pullulan, carrageenan;
    proteins: e.g., gelatin and whey protein;
  • Other bio-based polymers: e.g., lipid, lignin, natural rubber, urushiol, DNA, etc. polymers produced directly from microorganisms bacterial cellulose; polyhydroxyalkanoates, poly-ɛ-caprolactones.
  • Synthetic bio-based polymers, including:
    natural-based or bio-based synthetic polymers, the monomers of which are derived from renewable resources (e.g., poly(lactic acid)-PLA);
    partially bio-based polymers such as polyethylene (PE), poly(ethylene terephtalate) (PET) and polyamide (PA), etc. [19,20,21,22].
Among biopolymers, starch and chitosan are the most often used polysaccharides in the production of biopolymer films. Chitosan is a highly interesting film forming base due to its biological (antimicrobial and antioxidant activity) and physical (thermal or mechanical) properties [23]. The use of starch as film forming coatings eliminates the problem of environmental pollution because it is an economical, non-toxic, and environmentally friendly biopolymer. However, starch’s poor mechanical properties, moisture sensitivity, and weak barrier properties restricts its use as a commercial packaging materials [24]. Pullulan is a neutral polysaccharide obtained from the fermentation medium of the fungus-like yeast Aureobasidium pullulans [25]. It has superb optical properties, and applications of this type of non-toxic, biodegradable polymer for use in novel optical materials may be encouraging [26].
Proteins such as gelatin are also used in the preparation of films. However, proteins have poor mechanical properties and very high sensitivity to moisture which are problems in the commercialization of this type of film [27]. Beeswax is a good barrier against moisture migration, because it is hydrophobic and has a firmly packed crystalline structure, which is why it is often added to a polysaccharide and protein matrix to reduce the parameters of water vapor permeability of biopolymer films [28,29].
Poly(lactic acid) or polylactide is derived from renewable biomass products and wastes such as corn starch. This polymer is compostable, non-toxic, biocompatible, thermoplastic, and has desirable mechanical properties [30]. However, PLA has some limitations for packaging applications such as weak water vapor permeability, low thermal stability, and high rigidity [31]. Polyvinyl alcohol (PVA) is a non-toxic polymer, which has excellent chemical resistance and physical properties [32]. The major drawback of this synthetic polymer is its low mechanical strength [33]. Polyhydroxyalkanoates (PHAs) are a family of biopolyesters produced by a wide variety of bacteria, which are often used as a film component [34]. Due to its biodegradability and compatibility with various polymers and nanofillers, poly(ɛ-caprolactone) (PCL) has recently gained interest [35]. However, due to the hydrophobicity and crystallinity of this polymer, PCL undergoes very slow biodegradation as a result microorganisms in the environment [36].
As noted, there are many types of biopolymer films with each having their own advantages and disadvantages (Table 1). Some drawbacks of the polysaccharide or protein films can be overcome. Blending the different biopolymers together is one of the promising methods of improving the functional properties of biopolymer-based films [37,38]. The second is to use cross-linking of biopolymer films with various nanofillers.

2.2. Types of Nanofillers

There are four types of nanofillers: clays, organic, inorganic, and carbon nanostructure. The organic nanofillers include natural biopolymers (e.g. chitosan, cellulose), whereas inorganic agents are either a metal (e.g. silver) or metal oxide (e.g. ZnO and TiO2) [4,67]. Carbon nanostructures can be classified into fullerenes, graphene, carbon nanotubes, and nanofibers [68]. A list of examples of properties of various nanofillers is shown in Table 2.
The hybrid materials, which are materials made from hybrid of organic and inorganic materials, attracts more and more attention in various fields of research. In comparison to conventional materials, organic–inorganic hybrid materials have features derived from the inorganic part (rigidity, dimensional stability and thermal stability) as well as the properties of organic materials (tenacity and workability) [94].

2.2.1. Clay and Organic Nanofillers

Natural clays are inexpensive materials and exist as aluminum silicate, which consists of fine-grained minerals. Currently, nanoclays are the most commonly used nanoparticles, and they exist in the form of sheets/platelets, which have at least one dimension in the nanoscale range. The most often used nanoclay is montmorillonite (MMT), which consists of two tetrahedral silica sheets connected to an edge-divided, eight-sided aluminum oxide sheet [16]. Another type of nanoclay is halloysite (Hal), which is a natural aluminosilicate clay mineral with hollow, cylindrical-shaped nanotubes [95]. The strong mechanical properties of Hal have been integrated with the antimicrobial and photocatalytic properties of metal oxides and metal nanoparticles by grafting them on Hal [95,96,97].
Natural biopolymer nanofibrils, which are composed of various biopolymer molecules—i.e., cellulose, collagen, and chitin—are often used due to their biocompatibility, biodegradability, durability, availability, and unique mechanical properties [98]. Nanocellulose, which is extracted from cellulose, is characterized by the reactive surface of hydroxyl groups and thus can be adapted to different surface properties [99,100]. There are three types of cellulose used on the nanoscale as an additive to biopolymer films: nanocrystalline cellulose, nanofibrillated cellulose, and bacterial nanocellulose [100]. These types of nanocellulose differ in their morphology, degree of crystallinity, particle size, and some other properties. These differences result from their sources and various extraction methods used [100]. Nanocrystalline cellulose with its high crystallinity and short-rod shape is also known as cellulose nanocrystals or cellulose nanowhiskers and is usually extracted from cellulose fibrils by acid hydrolysis [101]. The source of cellulose nanocrystals may also affect the functionality of the nanofiller [102].
Nanofibrillated cellulose is extracted from cellulose fibers using mechanical methods. This long, flexible, and tangled nanocellulose is also known as cellulose microfibril, microfibrillated cellulose, cellulose nanofiber, cellulose nanofibril, and nanofibrillar cellulose [100].
Bacterial nanocellulose (BNC), which is mainly extracted from cultures of the Gram-negative bacteria, Gluconacetobacter xylinus, has a higher molecular weight and crystallinity than cellulose from plant sources [103]. The different types of nanocellulose have been incorporated to many biopolymer films [104,105,106,107].
Chitosan and chitin nanoparticles obtained respectively from chitosan or chitin have gained attention as nanofillers, due to their attractive surface area, biocompatibility, non-toxicity and film forming ability [108,109,110,111,112]. An important limitation in the use of chitosan nanoparticles is their poor stability. It can be improved by controlling the conditions e.g., by changing the structure with chemical agents. Poor solubility of chitosan nanoparticles is the next limitation, which is a major problem in the encapsulation of hydrophobic drugs [113].

2.2.2. Inorganic Nanofillers

Due to their better mechanical, thermal, physical, biological, and chemical properties than bulk materials, nanoparticles (NPs) are an interesting solution as functional agents for biopolymers films. Copper nanoparticles (CuNPs) are known to possess antimicrobial activity on wide spectrum of microbes. Released copper ions have a high redox potential and the ability to destroy and cause apoptosis of microbial cell components [114]. Owing to their excellent physicochemical and biological properties, silver nanoparticles (AgNPs) have gained increasingly more attention for use in wound healing [115], food packaging [116], and in medical applications [117]. Selenium nanoparticles (SeNPs) show significantly reduced toxicity compared to selenium. In addition, SeNPs can be used as a platform for transporting different drugs to target destinations [118].
Metal oxide nanoparticles such as zinc oxide (ZnO NPs), titanium oxide (TiO2 NPs), silica (SiO2 NPs), aluminum oxide (Al2O3 NPs), cerium oxide (CeO2 NPs), iron oxide (Fe2O3 NPs), and copper oxide (CuO NPs) have been added to biopolymer films as active nanofillers. TiO2 NPs and ZnO NPs have photocatalytic antibacterial properties, caused by reactive oxygen species (ROS) formation after exposure to UV-light [119]. Additionally, ZnO NPs can release zinc ions that damage bacterial cells, showing that antimicrobial properties do not have to be dependent on UV exposure [120]. The low hydrophilicity of TiO2 prevents the penetration of moisture into the biopolymer matrix. Moreover, the barrier effect of TiO2 leads to reduction in water vapor permeability [121]. CuO NPs have antimicrobial, antibacterial, and antioxidant activity and exhibit a UV-blocking effect [122,123]. Tin oxide (SnO2) nanoparticles have electrical, thermal, mechanical and gas barrier properties [124,125]. The addition of non-toxic and neutral Al2O3 NPs to the biopolymer matrix provides a wide range of promising properties due to its small particles, significant surface area, and good activity [126,127].

2.2.3. Carbon Nanofillers

Graphene-based materials have gained attention due to their properties which include excellent mechanical properties and substantial electron mobility. Graphene, in comparison to other carbon-based nanomaterials (for example carbon nanotubes), is characterized by a larger surface area that could facilitate interactions with the polymer matrix [128,129]. Graphene oxide (GO) is the most promising nanofiller, among the graphene-family nanomaterials, because it has a lower tendency to agglomerate than pristine graphene [130]. The biggest limitations, for the use of GO, are its intrinsic zero band-gap energy and low solubility in organic and aqueous solvents [131]. The chemical functionalization of graphene oxide leads to its modification and thus increases the possibilities of its potential application. Reduced GO (RGO) has very good mechanical, physical, electrical, and thermal properties, which makes it attractive nanofiller in biopolymer films [132].

2.2.4. Other Nanofillers

Semiconductor quantum dots (QDs) are attracting increasing attention due to their optical properties and high photostability [133]. Due to their strong tendency of oxidation and agglomeration, it is desirable to improve the compatibility and stability of this type of nanofillers in biopolymers [134]. Graphene quantum dots (GQDs) incorporated in biopolymer films have advantages such as availability, biodegradability, low price, and low production cost, making the nanocomposite films a preferred candidate for use in optoelectronics applications [135].

3. Effects of Nanofillers on the Functional Properties of Biopolymer-Based Films

The graphical illustration of nanofillers use within the biopolymer matrix along with their functional properties is presented in Figure 1.

3.1. Effects of Nanofillers on the Physical Properties of Polymer-Based Films

Water resistance is an important functional parameter of biopolymer films. In many studies, water resistance of the films is evaluated by their water solubility, swelling degree, water content, and water vapor permeability [87,136].
The addition of nanofillers decreases the solubility of biopolymer films due to the ratio of dimensions and crystalline areas of the fillers [87,137]. Noshirvani and co-workers (2018) confirmed that the lower solubility of the nanocomposite films is caused by the hydrogen bond between the hydroxyl groups of starch, PVA, and cellulose nanocrystals. The three-dimensional network generated leads to the strengthening of the network and reduces its solubility [138]. The same trend was observed in kefiran-caboxymethyl cellulose films with CuO NPs [139] and in kefiran films with Al2O3 NPs [127]. Moreover, the type of nanoparticles also plays an important role in the solubility of the film. A decrease in the solubility of the film can be attributed to the very low solubility of the nanoparticles compared to the polymer chains, which leads to a reduction in the hydrophilicity of the biopolymer matrix [121]. The interactions between halloysite nanoclay and soluble soybean polysaccharide caused a decrease in water solubility due to reduction in the availability of hydroxyl groups to interact with water [140].
The empty spaces in the structural network of nanocomposites can be occupied by water molecules. This phenomenon is considered the water content of the film [87,141]. The addition of nanofillers into the films can cause a decrease in water content. This behavior could be related to the interaction between nanofillers and functional groups of the biopolymer chain, which can lead to a reduction in the available spaces in the biopolymer matrix [87]. Noshirvani and co-workers (2018) developed starch-PVA films with cellulose nanocrystals (CNC). They noticed that the addition of the nanofiller caused a decrease in the moisture content of the film. Such behavior was demonstrated by the crystal structure of the CNC, which causes lower water uptake than the polymer matrix and the formation of strong hydrogen interactions between the CNC and the starch-PVA matrix [138]. The addition of nano-SiO2 into whey protein isolate–pullulan matrix decreased moisture content of nanocomposite films. The strong interactions between components caused the diffusion of water molecules to nanocomposites [142].
An important property of biopolymer films, and in particular polysaccharide films, is the swelling degree, which determines the amount of absorbed water. As the value of the swelling degree increases, the tolerance of film to water increases [143]. In general, the addition of nanofillers improves the swelling degree and increases the water resistance of the films. This is attributed to the strong hydrogen interactions between the nanofiller/nanofiller and the nanofillers/biopolymer matrix [143]. However, the reverse trend can also be noticed. The addition of chitin nanofiber (CHNF) caused an incremental swelling degree of chitosan films, which was due to increased amount of OH- groups from CHNF and resulted in increased water absorption [144]. The same phenomenon was observed in the chitosan/starch films with halloysite nanotubes, which was attributed to the increased porosity and hydrophilicity of the nanocomposite films [145].
The water vapor permeability (WVP) is very important for fresh food products and for products where dehydration and absorption of moisture should be avoided. The values of WVP of packaging systems should be at the lowest possible level [28,87,146]. The water vapor permeability in biopolymer films is influenced by the chemical nature of macromolecules, porosity and crystallinity, degree of cross-linking, relative humidity, and addition of a plasticizer [147,148]. The addition of nanofillers to biopolymer films has an impact on the water vapor barrier properties. The low WVP of films is an important feature in the packaging of food products as they can reduce the moisture transfer between the inner and outer packaging environment [149]. The presence of impermeable nanoparticles in the biopolymer matrix prevents the mobility of the biopolymer chain, and consequently can lead to a reduction in the water vapor permeability of the nanocomposite films [28]. In addition, the hydrophobic character of nanoparticles (e.g., TiO2) affects the reduction of the WVP of the film, due to the low aspect ratio of this nanoparticle and the irregular strengthening of the chains [87]. Shankar et al. concluded that the presence of halloysite nanotubes and AgNPs creates an increased tortuous path for the passage of water vapor molecules through the alginate matrix. On the other hand, exceeding the critical concentration of nanofillers increases the water vapor permeability of nanocomposite films [150]. Similar patterns of changes in WVP properties have been observed in many studies [87,95,143,151,152,153].
The type of nanofiller used also has an effect on the WVP of films. Zahedi and co-authors (2018) established that sodium MMT nanoclay is more effective in the reduction of the WVP of carboxylmethyl cellulose films than ZnONPs. These differences may result from various structural and arrangement features. ZnONPs have a hexagonal close packaged structure, while MMT is a layered silicate structure [154]. Nanoclays impede the diffusion of the water vapor due to impermeable layers of this type of nanofiller [155,156]. However, the addition of another type of nanoclay, laponite—synthetic hectorite-like clay—to kafirin film, does not significantly affect the WVP of the tested films. The hydrophilic clay, due to the presence of Si-OH groups, can affect the hydrophilicity of the surface [157]. The same trend was observed in collagen films with laponite [158].
In related literature, the opposite effect on the water barrier properties of biopolymer films is also reported. The addition of MMT/alkylammonium (hexa- and tetra-decyltrimethylammonium) and MMT/chitosan to cellulose acetate coatings increased the WVP value of the film [159]. Also the capping agent used in the preparation of the nanoparticles is critical for the WVP of the film. CMC was used as a capping agent in preparation of ZnO NPs. The nanoparticles caused an increase in the WVP of the gelatin films, which is attributed to the hydrophilic nature of CMC used as capping agent [160]. Silva and co-workers (2019) prepared new nanocomposites based on laponite and cellulose nanofibers. They concluded that the increase in WVP is related to the hydrophilic nature of laponite [161].
Oxygen permeability (OP) is one of the crucial parameters of films used in food packaging. Low OP values are preferred, because oxygen can cause deterioration in the quality of packed food products [162]. The addition of nanofillers into biopolymer films may cause a decrease in the OP values. Wu et al. (2019) stated that the AgNPs addition into nanocellulose films with grape seed extracts caused the improvement in barrier properties of the tested films. AgNPs filled the interspaces of matrix and hindered the transfer of O2 molecules through the film [163]. The presence of nanocellulose also causes the reduction of OP of the film, which can be attributed to the formation of a dense network structure of the film matrix [152]. A similar phenomenon was observed with the addition of halloysite nanoclay into potato starch films. The reduction in the permeability values may result from the fact that gases have an extended diffusion path [164].

3.2. Effects of Nanofillers on the Mechanical Properties of Polymer-Based Films

Two main parameters which are often used to determine the mechanical properties of biopolymer films are tensile strength (TS, MPa), which is used to measure strength, and the percentage of elongation at break (EAB, %) which is used to determine the elasticity of the film. These parameters must meet certain standards to maintain integrity during packaging [149]. The addition of nanoparticles to biopolymer films significantly affects their mechanical properties. This may be due to the fact that nanoparticles have a very large specific surface area that can affect interfacial strength and degree of dispersion. Even distribution of nanoparticles within the biopolymer matrix results in a specific transfer of stress through the shear mechanism from the biopolymer matrix to the nanoparticles and could result in effective load transfer and increased strength of the film [31,165].
The phenomenon of strengthening the tensile strength may also result from the interaction between the nanofiller and the biopolymer matrix, where hydrogen and covalent bonds between nanoparticles and, for example, hydroxyl groups of biopolymer are formed, which leads to the strengthening of molecular forces between nanoparticles and the biopolymer [143,166].
Nanofillers can fill the free space between biopolymer chains, which increases the intermolecular attraction force, making the biopolymer matrix very dense and less permeable [143]. Improvement in TS of polymer films by addition of nanofillers has been observed in films from konjac glucomannan [167], fenugreek seed gum [168], alginate [95], chitosan [169], agar [150], starch [121,143], soy protein [152], whey protein [87], and CMC [170].
Reinforcement of biopolymer films with different nanofillers have been studied by others. In the presence of nanoclay in a gelatin matrix, the values of TS of gelatin films improved [171]. The same trend was observed in starch-polyvinyl alcohol films with cellulose nanocrystals (CNC) [138]. The authors observed improvement of mechanical properties of these films due to the stiffness of the CNC and the formation of hydrogen bonds between the CNC and the polymer matrix. Popescu and co-workers found that incorporating rigid molecules of CNC into starch-PVA films induces orientation of the polymeric molecules and favors the formation of H-bonds between the polymeric chains [172]. Oun and Rhim (2016) isolated cellulose nanocrystals from rice, wheat, and barley straws. They concluded that properties of carboxymethyl cellulose (CMC) films varied depending on the type of cellulose nanocrystals. The addition of rice and barley straw into CMC film improved tensile strength parameters. However, the best results of WVP were achieved in CMC films with wheat straw [102]. Also, the combination of different nanofillers affects the mechanical properties of films. The binary films (PLA-AgNPs and PLA-CNC) showed lower values of mechanical properties than ternary films (PLA/AgNPs/CNC). The addition of AgNPs and CNC into PLA films improves their tensile strength and modulus values, which is due to synergistic effect between AgNPs and CNC [107].
The changes in elongation at break (EAB) exhibit the opposite trend than TS, when the nanofillers are added into the film polymer matrix. Since the addition of nanoparticles causes a decrease in the EAB value. Therefore, the addition of nanoparticles can increase the strength of the film but do not affect its flexibility [166]. A similar trend was observed in soy protein films with nanocellulose. The reinforcement may affect the limited movement of the polymer chain by reducing the free volume in the polymer matrix [152].
The critical concentration of nanofillers, above which the TS values of films decrease, should also be mentioned. Too high concentration of NPs causes the biopolymer matrix to be unable to evenly distribute the NPs, resulting in the reduction of strength [173]. The same problem can be observed with the addition of other types of nanofillers such as nanocrystals, which in too high concentration tend to grow together into larger crystals [105] or nanoclays which due to their poor dispersion and high surface energy can cause a decrease in the TS [168]. The decrease in TS at high nanofiller concentrations may also be related to crosslinking decrease in blend film and the damage to the polymer network layers due to the presence of large NP agglomerates [87].
When nanofillers are added to the films, they can cause lower resistance of a polymer matrix and its fracture. This can lead to the formation of micropores in the matrix, change of the crack growth of the path and transformation of the matrix shape. The formation of small clusters of nanoparticles increases the number of cracks and contributes to creating holes with a role in the energy loss and flexibility mechanism. The reduction of the polymer–nanoparticle interface surface may be due to the increase in the size of the nanoparticles. Thus, concentration at higher levels of nanoparticles leads to an optimal level of plasticity and reduced elasticity [87].
The increase in the TS value and the decrease in the EAB value may also be related to the moisture content of the tested biopolymer films. Water can act as a plasticizer in films and a decrease in plasticizer content reduces the elasticity of the film, which may lead to an increase in TS and Young modulus (YM) and a reduction in EAB [171,174].
Vicentini and co-workers (2018) studied the effect of modified carbon nanostructures (p-methoxyphenyl functionalized, carbon nanohorns, and reduced graphene oxide) in poly(L-lactic acid) composites. They concluded that the mechanical properties depended on the type of nanofiller used, and these differences can be attributed to the different size and shape of nanofillers [175].
The presence of graphene oxide in the cellulose carbamate matrix caused improvement in tensile strength of nanocomposite films. The roughness of the larger surface of graphene oxide sheets ensures better adhesion with cellulose carbamate, which could lead to stronger mechanical blocking [129]. A similar enhancement in TS was reported in chitosan/graphene oxide nanosheet films [176]. Son et al. (2015) compared mechanical properties of sodium carboxymethyl cellulose/reduced graphene oxide and sodium carboxymethyl cellulose/graphene oxide films. They concluded that reduced graphene oxide could interact more effectively with sodium carboxymethyl cellulose than graphene oxide because reduced graphene oxide reacts with sodium carboxymethyl cellulose through strong van der Waals interactions and hydrogen bonds due to the presence of oxygen-containing functional groups on the edge of the reduced graphene oxide sheet [177].
The type of biopolymer also affects the mechanical properties (modulus, strength, and toughness) of nanocomposite films. Achaby et al. (2018) studied the effect cellulose nanocrystals addition on the mechanical properties of alginate, chitosan, and κ-carrageenan films. The increase in mechanical strength of the alginate films with the addition of CNC was caused by strong hydrogen and ionic interactions between the free hydroxyl groups of the CNC and the hydroxyl and carboxyl groups of the alginate. The same trend was found in κ-carrageenan films with CNC, which the authors explained as a homogeneous CNC dispersion in the carrageenan matrix due to hydrogen bonds, which creates interfacial interaction between the CNC and the κ-carrageenan [178]. Moreover, Kassab and co-workers (2019) concluded that the increase in the mechanical properties of carrageenan films is related to the specific properties of CNC which was extracted from sugarcane bagasse. The authors explained that the morphology (needle shape), surface functionality, large aspect ratio, and high degree of crystallinity of CNC are responsible for improving the mechanical properties of the κ-carrageenan film [179].
Different conclusions were observed in chitosan nanocomposite films. The addition of up to 5 wt % of CNC caused an increase of mechanical properties parameters, which can be attributed to the interaction between the anionic sulphate groups of the CNC and the cationic amide groups of the chitosan. In addition, the nano-reinforcing effect of CNC was achieved through effective stress transfer of the CNC-chitosan interface. The addition of CNC at the highest level (8 wt %) caused a decrease in mechanical strength, which was related to the agglomeration phenomenon of CNC in the chitosan matrix. The addition of CNC also affects the elongation at break. The reduction of the EAB of each type of film is due to the rigid behavior of the CNC and its interactions with biopolymers that lead to a reduction in macromolecular biopolymer chains [178]. Kanmani and Rhim obtained three types of nanocomposite films based on agar, carrageenan, and CMC with ZnONPs. The addition of ZnONPs to the biopolymer films resulted in a decrease in TS and YM. In contrast, ZnONPs addition significantly increased the EAB of biopolymer films. The presence of ZnONPs did not interrupt the movement of polymer chains [180].

3.3. Effects of Nanofillers on the Antimicrobial Activity of Polymer-Based Films

The antimicrobial packaging system is intended to ensure the safety and prolong storage stability of food by preventing the development of microorganisms. The packaging materials can acquire antibacterial activity through:
(1)
the addition of antibacterial substances to the biopolymer matrix;
(2)
use of inherently antimicrobial polymer (for example polymer resins or chitosan);
(3)
irradiation of a biopolymer matrix which results in the formation of reactive oxygen species (ROS) [181,182].
The antibacterial agents can be divided into two types: bactericidal, which kill bacteria; and bacteriostatic, which cause the inhibition of bacteria growth. The incorporation of nanoparticles into biopolymer films increased their antimicrobial activity. The mechanism of antimicrobial activity of nanoparticles depends on the type of nanoparticles as well as on the species and the natural properties of the bacteria. The bacteria’s sensitivity to NPs is influenced by the bacterial cell wall structure and their growth rate. Fast-growing bacteria are more sensitive to NPs than slow-growing bacteria [87,183]. Moreover, the antibacterial activity of nanoparticles is greater in Gram-negative bacteria than in Gram-positive bacteria, which is due to the bacterial cell structure. Gram-positive bacteria have a thick three-dimensional layer of peptidoglycan (~20–80 nm) that creates a specific barrier to nanoparticles and prevents them from entering inside the cell. A thin layer of peptidoglycan (~7–8 nm) in Gram-positive bacteria is an ineffective barrier against released nanoparticles ions [184]. There are several hypotheses of the mechanism of antibacterial action of nanoparticles. The first of these concerns the interaction of negatively charged nanoparticles with biologically active molecules that are found on the surface of cells, which in turn can lead to the leakage of cellular components [85]. The second type of mechanism concerns the interactions of nanoparticles with bacterial DNA, resulting in cell death. The potential mechanisms of antimicrobial activities of nanofillers are illustrated by the example of nanoparticles shown in Figure 2.
In addition, each type of nanoparticles could work differently. In the presence of ZnONPs, H2O2 is generated [185], whereas sulfur ions, which are generated from SNPs, cause the formation of toxic H2S, which interacts with the -SH group and generates stress ROS [85,186]. The type of nanoparticles also affects their bactericidal activity. Small-sized nanoparticles have a stronger bactericidal effect [187], while a positive surface charge of metallic nanoparticles results in easier interaction with a negative charge of bacterial surfaces [188].
In related literature, many antibacterial aspects of nanostructures have been reported. The influence of various types of nanoclays (Cloisite Na+, Cloisite 20A and Cloisite 30B) on the antibacterial properties of whey protein isolate (WPI) films was investigated. Only the presence of Cloisite 30B showed a bacteriostatic effect on L. monocytogenes, while none of the different types of nanoclays had any antibacterial effect against E. coli [189]. The antibacterial effect of Cloisite 30B can be attributed to the presence of a quaternary ammonium group in the silicate layer, which destroys bacterial cell membranes and causes cell lysis. The same behavior was observed in PLA films with Cloisite 30B [190]. Although hydrophilic NaMMT or organically modified montmorillonite (OrgMMT) do not show any particular antibacterial action, it has been shown that their addition to the film enhances the antibacterial effect of chitosan-PVA films [191].
Chitin nanofibrils (CNF) could be characterized as bacteriostatic rather than as a bactericidal agent. The addition of CNF into carrageenan matrix improved antimicrobial activity of nanocomposite films. The proposed mechanism of CNF activity consists of blocking access to nutrients and oxygen, which causes the bacteria to flocculate [192]. Salari and coworkers (2018) developed active chitosan films with bacterial cellulose nanocrystals (BCNC) and silver nanoparticles AgNPs. The antibacterial and antifungal properties of this type of film were related to the presence of AgNPs, while BCNC and chitosan had no antibacterial or antifungal activity [184]. The mechanism of antibacterial behavior of ZnO nanorods could have a physical basis. Nanorods could act like needles that easily pass through the bacterial cell wall [185]. The same results were reported in sago starch films with ZnO nanorods [193].
The high antimicrobial activity of GO is due to the physical destruction of cell membranes due to exceptionally sharp edges of GO and chemical oxidation through generation of ROS. In comparison to other carbon nanomaterials, graphene oxide has low cost of production and causes mild cytotoxicity to mammalian cells in low doses [194,195].
Increased attention is being paid to hybrid materials that give new properties to nanocomposite materials. Hybridization of metal nanoparticles and polymer can improve the properties of individual components. The addition of the hybrid of ZnO NPs-chitosan to starch films resulted in an increase in antibacterial activity against E. coli and S. aureus, compared to starch films alone [196]. Moreover, when a hybrid of PVA/GO/starch AgNPs was used in PVA matrix, an increase in the antibacterial properties have been observed, which might be due to the synergistic effects of GO and AgNPs in PVA matrix [197]. The addition of GO and CNC stimulates the antibacterial activity of PLA film, resulting in greater activity than single nanofillers [198].
Table 3 presents some recent studies on the impact of nanofillers on functional properties of nanocomposite films.

4. Functional Application of Biopolymer-Based Films with Nanofillers

Although many countries have made legal regulations regarding the contact of nanomaterials with food or with human skin, those regulations vary from country to country. Moreover, the definition of ’nanomaterial’ is not even unified between countries. The detailed review of the legal status regulations for nanomaterials in EU has been recently performed by Rauscher et al. [253]. Nanomaterials that are allowed for use in contact with humans are required to have a detailed risk assessment [254]. This chapter is devoted to nanocomposite materials and their use in various areas of life.

4.1. Food Preservation Application

Packaging materials made from natural biopolymers are characterized by poor mechanical properties and high WVP parameters. The utilization of nanotechnology in this field may help to improve these parameters as well as give them completely new active properties. The main task of the nanocomposite packaging is to increase the shelf-life of food during storage and distribution [67].
Chitosan–gelatin films with the addition of AgNPs have been used to protect red grapes. Preliminary studies confirmed the active nature of such coatings, extending the storage time of grapes. The addition of 0.05% AgNPs in the chitosan–gelatin films resulted in the extension of the red grapes shelf-life with no signs of molds for 14 days, whereas the concentration of 0.1% AgNPs prolonged the shelf-life to 18 days [200]. In another case, AgNPs were incorporated into chitosan films and prolonged the storage time of litchi from 4 to 7 days [5]. The incorporation of MMT into starch–cashew tree gum increased protection against moisture loss of cashew nut kernels. However, the MMT additive did not achieve the critical value of peroxides for acceptance (10 or 20–30 mEq O2 kg−1) [6].
Soft white cheese was packed into chitosan-PVA films with TiO2 NPs and stored at 7 °C for 30 days. After 15 days of storage, the number of coliforms significantly decreased in the samples coated with films with 2%, 4%, and 8% TiO2 (by 1.47, 1.47, 1.30 log cfu/g cheese, respectively, compared to the control (1.90 log cfu/g). The results showed that bionanocomposite films with strong antimicrobial activity against gram positive (S. aureus), gram negative (P. aeruginosa, E. coli) bacteria and fungi (C. albicans) could be used as an environmentally friendly material for food packaging. Additionally, the lack of migration of TiO2 from the film to cheese was confirmed, which proves the high safety of this type of material [7]. Nanostructured chitosan–manolaurin films caused a reduction of L. monocytogenes population (by 2.3–2.4 log) on ultrafiltered cheese after 14 days [8].
The addition of ZnO nanoparticles into mahua oil-based polyurethane/chitosan film significantly influenced antibacterial properties when these films were used for packing carrot slices. After 9 days of storage, carrot slices packed in mahua oil-based polyurethane/chitosan+ ZnO NPs films had a lower degree of bacteria growth (by approx. 0.3–0.6 log cfu/g) than pieces of carrot slices without film and with PE film [213]. Antimicrobial activity against E. coli and L. monocytogenes was recorded for cooked minced fish paste packaged in PLA films with ZnO NPs and after 10 days of storage, the number of colonies of both bacteria was reduced to zero. The antimicrobial activity might have resulted from direct contact of microorganisms with ZnO NPs or Zn2+ ions emitted from the film [31]. The addition of ZnO NPs into chitosan/CMC also positive influenced packaged Egyptian white soft cheese (rheological properties, color measurements, moisture, pH, and titratable acidity) during storage [211]. The starch–halloysite–nisin nanocomposite films effectively protected Minas Frescal cheese against post-process contamination with L. monocytogenes [236]. Echeverria and co-workers (2018) investigated the effect of MMT and clove essential oil in starch films on storage of muscle fillets of bluefin tuna (Thunnus thynnus). They noticed that the presence of clay prolonged the antimicrobial and antioxidant effects of clove essential oil, while no migration of clay’s metal (Si and Al) into the muscle of the fish were observed [9]. Mathew and co-workers (2019) developed PVA/MMT K10 clay/AgNPs films and used them to improve the shelf-life of chicken sausages. The results showed that this type of nanocomposite packaging system is suitable for sausages, inhibiting the growth of total aerobic bacteria [10].
Starch/cellulose nanocrystal/grape pomade extract ‘Viognier’, nanocomposite films exhibited the highest antibacterial effect against L. monocytogenes inoculated on the ready-to-eat chicken meats during a 10-day storage period at 4 °C [11]. Strong antimicrobial properties were exhibited by chitosan–starch–cellulose nanofibrils films during storage of fresh beef sirloin [12]. The antimicrobial properties and high biocompatibility of chitosan/GO/TiO2 NPs nanocomposite films were observed when used in packaging strawberries and mangos. The results indicated that nanocomposite wrap effectively delays loss of weight in fruits, with weight loss below 5% after 7 days of storage, and prevents bacterial contamination in food products [13].
The influence of nanofillers on the properties of LDPE synthetic film was also studied. The use of CuNPs as an active additive to LDPE films positively influenced the shelf-life of Peda (Indian sweet dairy product) [151]. Luo et al. (2015) incorporated nano-SiO2 into LDPE films and used this type of packaging for storing chilled white shrimps. LDPE+nano-SiO2 are an active packaging material with antimicrobial and enzyme-inhibiting properties, in which shrimps remain fresh for 8 days [14]. Donglu and co-authors obtained a nanocomposite packaging material from LDPE and SiO2/TiO2 /Ag NPs, which regulated CO2 and O2 levels as well as inhibited the growth of microorganisms and scavenge ethylene during storage of mushrooms (Flammulina velutipes) [15].
Sodium CMC nanocomposite films with photoluminescent ZnS NPs exhibited a blue emission centered at 445 nm under UV light excitation, which can be easily incorporated into paper sheets in order to prepare protective paper, which can be used as a food packaging [255]. Rice straw nanofibrillated cellulose (NFC)/CHNPs nanocomposite films were used to coat baggasse paper. The coating of paper sheets confirmed improvement in paper sheet properties, improving the tensile strength, grease-proof properties, antibacterial activity, and reduced porosity [110]. El-Wakil and co-workers used gluten/CNC/TiO2 NPs nanocomposite films to coat paper sheets. The paper sheets coated with the aforementioned nanocomposite showed antimicrobial activity against S. cervisiae, E. coli, and S. aureus [256]. Nevertheless, further research is necessary to confirm the use of paper sheets coated with nanocomposite as an active packaging system, because there is still a lack of relevant in vivo studies.
In addition, the effectiveness of halloysite nanotubes as nanocontainers for active substances - sodium sorbate (a commonly used food preservative) - embedded in the zein matrix was tested. Preliminary studies have shown that it is possible to obtain an active packaging that has special packaging systems for the controlled release of active substances [257]. The presence of nanoclay in carnauba wax coatings enhanced the sensory acceptability and nutritional quality and effectively prevented the weight loss of ‘Valencia’ oranges (Citrus sinensis L. Osbeck) during storage at 4 °C for 8 weeks [258].
Currently, a large limitation in the commercial use of nanocomposite materials is the level of transition of nanomaterials from packaging to the food matrix. Although a number of studies have been carried out on the effects of using nanocomposites on prolonged storage of various types of food, consumers and legislation bodies expect confirmation of the safety of using nanocomposite packaging. One of the main concerns related to the use of nanocomposites is the emergence of new, diverse strains of bacteria and allergens, as well as the increased degree of adsorption of nanomaterials within the environment [67]. There are studies in which the migration of nanoparticles from food packaging has been studied [259,260]. Biodegradable starch/clay nanocomposite films could be used as food packaging owing to their low overall migration limit [260]. In addition, contact tests using food simulants are carried out to check the safety of the obtained nanocomposite materials. For example, the chitosan films with SeNPs met the required regulations and European directives on food packaging (EN1186-1, 2002), according to Commission Regulation No. 10/2011 [261].

4.2. Wound Dressing

The presence of ROS in wound often hampers wound healing. Cerium oxide nanoparticles (CeO2 NPs) with strong antioxidant properties may contribute to active ROS purification. One of the methods for the production of wound dressings is the electrospinning technique, which consists in obtaining fibers ranging from tens of nanometers to micrometers [262]. Naseri-Nosor and co-workers (2017) developed CeO2 NPs with poly (ɛ-caprolactone)/gelatin films as potential wound dressing material. After 2 weeks, wounds treated with PCL/gelatin+ 1.5 % CeO2 exhibited complete wound closure compared to control film without CeO2 [263].
To prevent damage caused by the removal of wound dressings, plasma treated with electrospun polycaprolactone was coated with gelatin–silver nanoparticle membranes using the multi-immersion technique. Compared to monolayer coating, this type of multilayer coating (six-time coated membrane) accelerated the healing of wounds by 3 days. Moreover, the membrane did not adhere to the wound during peeling off, preventing damage to the newly formed tissue [264]. The antibacterial chitosan/poly(vinyl pyrrolidone)(PVP)/nanocellulose films exhibited good compatible property and low cytotoxicity with blood. Moreover, this type of nanocomposite films could be used as wound dressing material due to their swelling, thermal, and mechanical properties [71]. The reinforcement of nanostarch in chitosan/PVP/1%-Stearic acid films caused the enhancement in the healing effect of albino rats [71].
Antimicrobial agent, chlorhexidine (CLX), was intercalated between the layers of MMT in the chitosan films. The chitosan films with MMT containing 5% CLX were non cytotoxic, whereas in the case of 1% CLX films only film with neat CLX resulted in cytotoxicity. The localized and prolonged release of CLX on the chitosan-MMT matrix could be a good solution in the area of wound dressings [265]. Cacciotti and coworkers (2018) designed poly(lactic) acid fibrous membranes with addition of H2S slow releasing donors extracted from garlic. The obtained antibacterial fibrous scaffolds/patches not only have the ability to stimulate cMSC (cardiac mesenchymal stem cells) proliferation, but also reduce oxidative damage [266].
Currently, nanotechnology is continuing to gain interest in the field of skin regeneration. On a laboratory scale, nanofillers placed in a polymer matrix accelerate wound healing, due to their antibacterial activity and improved angiogenesis. There are several important issues regarding the use of nanofillers in the wound healing process. Above all is the issue of dressing stability, so that nanofillers do not form a covalent bond with the polymer chain. In addition, there is still little data which would study the long-term safety of the use of nanofillers. Therefore, the mechanism of nanofillers should be thoroughly studied in order to be able to manage the cell behaviors. Understanding the mechanism of nanofillers will create a broader perspective on the safety of their use, which is why the long-term control over the performance of nanofillers is required [267].

4.3. Drug and Enzyme Delivery System

A drug delivery system is based on the delivery of a pharmaceutical drug to a specific site of a disease with minimized toxicological risk. Nanotechnology supports this by producing nano-scale materials, which can be successfully used as carriers [268]. A gelatin film with magnetic nanoparticles can be used as a control system for drug release, where the use of a magnetic field is important. Research has confirmed that without a magnetic field, this type of film acts as a blockage of drug release, while under the influence of a magnetic field, it opens the diffusion pathways [207]. Javanbakht and Namazi (2018) obtained flexible nanocomposite CMC hydrogel films with graphene quantum dots and anticancer drug, doxorubicin, release property. This type of nanocomposite caused pH-sensitivity and consecutively prolonged the release of doxorubicin and showed non-significant toxicity against blood cancer cells (K562) [269].
Nanocomposite films not only can be carriers of drugs, but also for example, can transport proteolytic enzymes during the wine production process. Bromelain produced from pineapple stem is a proteolytic enzyme, which reduces the haze potential of white wines. Chitosan films with different nanoclays (MMT, sepiolite, and bentonite) were produced as supports for the covalent immobilization of proteolytic enzymes. The addition of these clays affected the amount of immobilized protein, which was higher for all the nanocomposite films compared to the clay-free sample (with the exception of modified bentonite). On the other hand, the addition of nanoclays negatively affected the catalytic properties of the immobilized protease [270].

4.4. Tissue Engineering Application

Tissue engineering relies on the regeneration of damaged tissues or organ reconstruction. One of the methods used in tissue engineering is the use of materials that act as implantable scaffolds. Nanotechnology proposes to create nanocomposites that offer different structures and properties [271].
Hassan and co-workers (2018) obtained nanocomposite scaffolds based on chitosan and CMC AgNPs decorated on carboxylated CNWs. Those scaffolds exhibited both mechanical properties, due to carboxylated CNWs addition, and antimicrobial properties due to AgNPs. This type of nanocomposite exhibited sufficient protein adsorption and mineralization capacity which can overcome bone related infections like osteomyelitis [272]. Selenium nanoparticles (SeNPs), with their excellent antioxidant activity were used as nanofiller in chitosan films to produce electrical conductivity in cardiac patches. Due to its electrical and mechanical properties, this innovative material can be used in cardiac tissue engineering [273]. Nanocomposite films, obtained from PVA and zirconium phosphate doped with Ca, Mg, and Ti nanoparticles were used for scaffold-guided tissue engineering application. It was observed that PVA films with Ca- doped ZrP nanoparticles had increased biological activity, whereas PVA films with Ti-doped ZrP nanoparticles had relatively higher mechanical properties [274]. Hydroxyapatite (HA) is the main component of inorganic bones and is a very good material for the development of internal scaffolds for bone repair [275]. HA nanoparticles promote adhesion, proliferation, and osteogenic differentiation of osteoblast-like cells therefore they were used in preparation of chitosan-HA nanocomposite films for bone tissue engineering application [276]. Arumugam and co-workers (2019) developed novel, porous, mechanically stable, hydrophilic nanocomposite polyvinylidene fluoride/poly(methyl methacrylate)/hydroxyapatite/TiO2 film scaffolds. The in vitro study confirmed its excellent osteocompatibility, so this type of nanocomposite material could be used as potential material for bone repair applications [277]. The poly(L-lactic acid)/Ca-deficient-hydroxyapatite mats obtained by the electrospinning method can be used as a potential scaffold for bone marrow mesenchymal stem cell culture [278]. D’Angelo and coworkers (2012) investigated potential responses of multipotent (human-bone-marrow-derived mesenchymal stem cells) and pluripotent stem cells (murine-induced pluripotent stem cells and murine embryonic stem cells) to poly(L-lactic acid)/Ca-deficient-hydroxyapatite mats. They observed that the osteogenic differentiation effect of mats was not dependent on the type of stem cells [279].
In another work, HA was added to GO-chitosan films by using the layer-by-layer technique along with the biomimetic mineralization procedure. The results indicate the potential use of aspirin-loaded GO-chitosan-HA films in the field of bone tissue engineering [280].

4.5. Other Applications

Nanocomposite materials are also used in other industries. The addition of ZnO nanoparticles into chitosan/PVA films indicate their efficient removal of AB 1 azo dye (dye component) from aqueous solution [214]. The MWCNT was functionalized by valine and starch to improve the compatibility of MWCNT with chitosan/PVA matrix. The results indicate that this nanocomposite film is a very interesting adsorbent for the removal of Cd(II) from aqueous solutions [239].
Due to their inherent optical properties, the presence of carbon quantum dots in CMC films effectively caused ultraviolet light to convert into blue light. Transparent sunlight conversion CMC film with carbon quantum dots could be used in agriculture planting [281]. Mitta et al. (2018) obtained DNA thin films incorporated with AuNPs to demonstrate efficient UV photodetectors. Due to the presence of Au NPs, photodetectors were characterized by stability and durability [282].

5. The Future of Nanomaterials

According to the regulations of the European Commission, nanomaterials are covered by a regulatory framework. The presence of nanomaterials, including in food and cosmetics, must be indicated on the label in the list of ingredients. In addition, consumers must have access to online resources and databases on nanomaterials [253]. Due to the promising properties of nanomaterials, increasingly more new materials are appearing and the assessment of their risks requires an individual approach to each nanomaterial. There are major concerns that the high surface-to-volume ratio of nanomaterials may result in their having higher reactivity and potential toxicity. Due to the concerns about the safety of using nanomaterials, further research is necessary to give an unambiguous answer as to whether and which nanomaterials can be a viable alternative to traditional materials for application in many different areas [226,253].

Author Contributions

All the authors contributed equally.

Acknowledgments

The financial support from the projects CEITEC 2020 (LQ 1601) and EFRR project (no. CZ.02.1.01/0.0/0.0/16_025/0007314) is highly acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

NPsnanoparticles
CMCcarboxymethyl cellulose
PVApoly(vinyl) alcohol
PLApoly(lactic acid)
TStensile strength
EABelongation at break
YMYoung’s modulus
EMelastic modulus
WVPwater vapor permeability
OPoxygen permeability
WSwater solubility
SRswelling ratio
MCmoisture content
WCAwater contact angle

References

  1. George, J.; Ishida, H. A review on the very high nanofiller-content nanocomposites: Their preparation methods and properties with high aspect ratio fillers. Prog. Polym. Sci. 2018, 86, 1–39. [Google Scholar] [CrossRef]
  2. El Achaby, M.; El Miri, N.; Aboulkas, A.; Zahouily, M.; Bilal, E.; Barakat, A.; Solhy, A. Processing and properties of eco-friendly bio-nanocomposite films filled with cellulose nanocrystals from sugarcane bagasse. Int. J. Biol. Macromol. 2017, 96, 340–352. [Google Scholar] [CrossRef]
  3. Cacciotti, I.; Mori, S.; Cherubini, V.; Nanni, F. Eco-sustainable systems based on poly(lactic acid), diatomite and coffee grounds extract for food packaging. Int. J. Biol. Macromol. 2018, 112, 567–575. [Google Scholar] [CrossRef] [PubMed]
  4. Sothornvit, R. Nanostructured materials for food packaging systems: New functional properties. Curr. Opin. Food Sci. 2019. [Google Scholar] [CrossRef]
  5. Wu, Z.; Huang, X.; Li, Y.-C.; Xiao, H.; Wang, X. Novel chitosan films with laponite immobilized Ag nanoparticles for active food packaging. Carbohydr. Polym. 2018, 199, 210–218. [Google Scholar] [CrossRef]
  6. Pinto, A.M.B.; Santos, T.M.; Caceres, C.A.; Lima, J.R.; Ito, E.N.; Azeredo, H.M.C. Starch-cashew tree gum nanocomposite films and their application for coating cashew nuts. LWT Food Sci. Technol. 2015, 62, 549–554. [Google Scholar] [CrossRef]
  7. Youssef, A.M.; El-Sayed, S.M.; Salama, H.H.; El-Sayed, H.S.; Dufresne, A. Evaluation of bionanocomposites as packaging material on properties of soft white cheese during storage period. Carbohydr. Polym. 2015, 132, 274–285. [Google Scholar] [CrossRef]
  8. Lotfi, M.; Tajik, H.; Moradi, M.; Forough, M.; Divsalar, E.; Kuswandi, B. Nanostructured chitosan/ monolaurin film: Preparation, characterization and antimicrobial activity against Listeria monocytogenes on ultrafiltered white cheese. LWT 2018, 92, 576–583. [Google Scholar] [CrossRef]
  9. Echeverría, I.; López-Caballero, M.E.; Gómez-Guillén, M.C.; Mauri, A.N.; Montero, M.P. Active nanocomposite films based on soy proteins-montmorillonite- clove essential oil for the preservation of refrigerated bluefin tuna (Thunnus thynnus) fillets. Int. J. Food Microbiol. 2018, 266, 142–149. [Google Scholar] [CrossRef] [PubMed]
  10. Mathew, S.; Snigdha, S.; Mathew, J.; Radhakrishnan, E.K. Biodegradable and active nanocomposite pouches reinforced with silver nanoparticles for improved packaging of chicken sausages. Food Packag. Shelf Life 2019, 19, 155–166. [Google Scholar] [CrossRef]
  11. Xu, Y.; Rehmani, N.; Alsubaie, L.; Kim, C.; Sismour, E.; Scales, A. Tapioca starch active nanocomposite films and their antimicrobial effectiveness on ready-to-eat chicken meat. Food Packag. Shelf Life 2018, 16, 86–91. [Google Scholar] [CrossRef]
  12. Yu, Z.; Alsammarraie, F.K.; Nayigiziki, F.X.; Wang, W.; Vardhanabhuti, B.; Mustapha, A.; Lin, M. Effect and mechanism of cellulose nanofibrils on the active functions of biopolymer-based nanocomposite films. Food Res. Int. 2017, 99, 166–172. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, W.; Xie, W.; Huang, X.; Chen, X.; Huang, N.; Wang, X.; Liu, J. The graphene oxide and chitosan biopolymer loads TiO2 for antibacterial and preservative research. Food Chem. 2017, 221, 267–277. [Google Scholar] [CrossRef] [PubMed]
  14. Luo, Z.; Qin, Y.; Ye, Q. Effect of nano-TiO2-LDPE packaging on microbiological and physicochemical quality of Pacific white shrimp during chilled storage. Int. J. Food Sci. Technol. 2015, 50, 1567–1573. [Google Scholar] [CrossRef]
  15. Fang, D.; Yang, W.; Kimatu, B.M.; Mariga, A.M.; Zhao, L.; An, X.; Hu, Q. Effect of nanocomposite-based packaging on storage stability of mushrooms (Flammulina velutipes). Innov. Food Sci. Emerg. Technol. 2016, 33, 489–497. [Google Scholar] [CrossRef]
  16. Majeed, K.; Jawaid, M.; Hassan, A.; Abu Bakar, A.; Abdul Khalil, H.P.S.; Salema, A.A.; Inuwa, I. Potential materials for food packaging from nanoclay/natural fibres filled hybrid composites. Mater. Des. 2013, 46, 391–410. [Google Scholar] [CrossRef]
  17. Othman, S.H. Bio-nanocomposite Materials for Food Packaging Applications: Types of Biopolymer and Nano-sized Filler. Agric. Agric. Sci. Procedia 2014, 2, 296–303. [Google Scholar] [CrossRef]
  18. Xiong, R.; Grant, A.M.; Ma, R.; Zhang, S.; Tsukruk, V.V. Naturally-derived biopolymer nanocomposites: Interfacial design, properties and emerging applications. Mater. Sci. Eng. R Rep. 2018, 125, 1–41. [Google Scholar] [CrossRef]
  19. Garavand, F.; Rouhi, M.; Razavi, S.H.; Cacciotti, I.; Mohammadi, R. Improving the integrity of natural biopolymer films used in food packaging by crosslinking approach: A review. Int. J. Biol. Macromol. 2017, 104, 687–707. [Google Scholar] [CrossRef]
  20. Sharma, B.; Malik, P.; Jain, P. Biopolymer reinforced nanocomposites: A comprehensive review. Mater. Today Commun. 2018, 16, 353–363. [Google Scholar] [CrossRef]
  21. Rouf, T.B.; Kokini, J.L. Biodegradable biopolymer–graphene nanocomposites. J. Mater. Sci. 2016, 51, 9915–9945. [Google Scholar] [CrossRef]
  22. Zare, Y.; Shabani, I. Polymer/metal nanocomposites for biomedical applications. Mater. Sci. Eng. C 2016, 60, 195–203. [Google Scholar] [CrossRef] [PubMed]
  23. Mujtaba, M.; Morsi, R.E.; Kerch, G.; Elsabee, M.Z.; Kaya, M.; Labidi, J.; Khawar, K.M. Current advancements in chitosan-based film production for food technology: A review. Int. J. Biol. Macromol. 2019, 121, 889–904. [Google Scholar] [CrossRef]
  24. Ogunsona, E.; Ojogbo, E.; Mekonnen, T. Advanced material applications of starch and its derivatives. Eur. Polym. J. 2018, 108, 570–581. [Google Scholar] [CrossRef]
  25. Singh, R.S.; Kaur, N.; Rana, V.; Kennedy, J.F. Pullulan: A novel molecule for biomedical applications. Carbohydr. Polym. 2017, 171, 102–121. [Google Scholar] [CrossRef]
  26. Horinaka, J.-I.; Hashimoto, Y.; Takigawa, T. Optical and mechanical properties of pullulan films studied by uniaxial stretching. Int. J. Biol. Macromol. 2018, 118, 584–587. [Google Scholar] [CrossRef] [PubMed]
  27. Hassan, B.; Chatha, S.A.S.; Hussain, A.I.; Zia, K.M.; Akhtar, N. Recent advances on polysaccharides, lipids and protein based edible films and coatings: A review. Int. J. Biol. Macromol. 2018, 109, 1095–1107. [Google Scholar] [CrossRef] [PubMed]
  28. Bahrami, A.; Rezaei Mokarram, R.; Sowti Khiabani, M.; Ghanbarzadeh, B.; Salehi, R. Physico-mechanical and antimicrobial properties of tragacanth/hydroxypropyl methylcellulose/beeswax edible films reinforced with silver nanoparticles. Int. J. Biol. Macromol. 2018. [Google Scholar] [CrossRef] [PubMed]
  29. Velickova, E.; Winkelhausen, E.; Kuzmanova, S.; Alves, V.D.; Moldão-Martins, M. Impact of chitosan-beeswax edible coatings on the quality of fresh strawberries (Fragaria ananassa cv Camarosa) under commercial storage conditions. LWT Food Sci. Technol. 2013, 52, 80–92. [Google Scholar] [CrossRef]
  30. Nofar, M.; Sacligil, D.; Carreau, P.J.; Kamal, M.R.; Heuzey, M.-C. Poly (lactic acid) blends: Processing, properties and applications. Int. J. Biol. Macromol. 2019, 125, 307–360. [Google Scholar] [CrossRef]
  31. Shankar, S.; Wang, L.-F.; Rhim, J.-W. Incorporation of zinc oxide nanoparticles improved the mechanical, water vapor barrier, UV-light barrier, and antibacterial properties of PLA-based nanocomposite films. Mater. Sci. Eng. C 2018, 93, 289–298. [Google Scholar] [CrossRef] [PubMed]
  32. Thong, C.C.; Teo, D.C.L.; Ng, C.K. Application of polyvinyl alcohol (PVA) in cement-based composite materials: A review of its engineering properties and microstructure behavior. Constr. Build. Mater. 2016, 107, 172–180. [Google Scholar] [CrossRef]
  33. Hoque, M.A.; Ahmed, M.R.; Rahman, G.T.; Rahman, M.T.; Islam, M.A.; Khan, M.A.; Hossain, M.K. Fabrication and comparative study of magnetic Fe and α-Fe2O3 nanoparticles dispersed hybrid polymer (PVA + Chitosan) novel nanocomposite film. Results Phys. 2018, 10, 434–443. [Google Scholar] [CrossRef]
  34. Martínez-Sanz, M.; Villano, M.; Oliveira, C.; Albuquerque, M.G.E.; Majone, M.; Reis, M.; Lopez-Rubio, A.; Lagaron, J.M. Characterization of polyhydroxyalkanoates synthesized from microbial mixed cultures and of their nanobiocomposites with bacterial cellulose nanowhiskers. New Biotechnol. 2014, 31, 364–376. [Google Scholar] [CrossRef]
  35. Bianco, A.; Di Federico, E.; Cacciotti, I. Electrospun poly(ε-caprolactone)-based composites using synthesized β-tricalcium phosphate. Polym. Adv. Technol. 2011, 22, 1832–1841. [Google Scholar] [CrossRef]
  36. Liu, M.; Zhang, T.; Long, L.; Zhang, R.; Ding, S. Efficient enzymatic degradation of poly (ɛ-caprolactone) by an engineered bifunctional lipase-cutinase. Polym. Degrad. Stab. 2019, 160, 120–125. [Google Scholar] [CrossRef]
  37. Bianco, A.; Calderone, M.; Cacciotti, I. Electrospun PHBV/PEO co-solution blends: Microstructure, thermal and mechanical properties. Mater. Sci. Eng. C 2013, 33, 1067–1077. [Google Scholar] [CrossRef]
  38. Cacciotti, I.; Calderone, M.; Bianco, A. Tailoring the properties of electrospun PHBV mats: Co-solution blending and selective removal of PEO. Eur. Polym. J. 2013, 49, 3210–3222. [Google Scholar] [CrossRef]
  39. Wang, S.; Lu, A.; Zhang, L. Recent advances in regenerated cellulose materials. Prog. Polym. Sci. 2016, 53, 169–206. [Google Scholar] [CrossRef]
  40. Wang, Y.; Liu, L.; Chen, P.; Zhang, L.; Lu, A. Cationic hydrophobicity promotes dissolution of cellulose in aqueous basic solution by freezing-thawing. Phys. Chem. Chem. Phys. PCCP 2018, 20, 14223–14233. [Google Scholar] [CrossRef]
  41. Wu, Y.; Li, Q.; Zhang, X.; Li, Y.; Li, B.; Liu, S. Cellulose-based peptidopolysaccharides as cationic antimicrobial package films. Int. J. Biol. Macromol. 2019, 128, 673–680. [Google Scholar] [CrossRef]
  42. Dutta, P.K.; Tripathi, S.; Mehrotra, G.K.; Dutta, J. Perspectives for chitosan based antimicrobial films in food applications. Food Chem. 2009, 114, 1173–1182. [Google Scholar] [CrossRef]
  43. Devlieghere, F.; Vermeulen, A.; Debevere, J. Chitosan: Antimicrobial activity, interactions with food components and applicability as a coating on fruit and vegetables. Food Microbiol. 2004, 21, 703–714. [Google Scholar] [CrossRef]
  44. Rambabu, K.; Bharath, G.; Banat, F.; Show, P.L.; Cocoletzi, H.H. Mango leaf extract incorporated chitosan antioxidant film for active food packaging. Int. J. Biol. Macromol. 2019, 126, 1234–1243. [Google Scholar] [CrossRef]
  45. Jiang, Z.; Neetoo, H.; Chen, H. Efficacy of freezing, frozen storage and edible antimicrobial coatings used in combination for control of Listeria monocytogenes on roasted turkey stored at chiller temperatures. Food Microbiol. 2011, 28, 1394–1401. [Google Scholar] [CrossRef]
  46. Thakur, R.; Pristijono, P.; Scarlett, C.J.; Bowyer, M.; Singh, S.P.; Vuong, Q.V. Starch-based films: Major factors affecting their properties. Int. J. Biol. Macromol. 2019. [Google Scholar] [CrossRef] [PubMed]
  47. Silva, A.P.M.; Oliveira, A.V.; Pontes, S.M.A.; Pereira, A.L.S.; Souza Filho, M.d.s.M.; Rosa, M.F.; Azeredo, H.M.C. Mango kernel starch films as affected by starch nanocrystals and cellulose nanocrystals. Carbohydr. Polym. 2019, 211, 209–216. [Google Scholar] [CrossRef] [PubMed]
  48. Younis, H.G.R.; Zhao, G. Physicochemical properties of the edible films from the blends of high methoxyl apple pectin and chitosan. Int. J. Biol. Macromol. 2019, 131, 1057–1066. [Google Scholar] [CrossRef] [PubMed]
  49. Spatafora Salazar, A.S.; Sáenz Cavazos, P.A.; Mújica Paz, H.; Valdez Fragoso, A. External factors and nanoparticles effect on water vapor permeability of pectin-based films. J. Food Eng. 2019, 245, 73–79. [Google Scholar] [CrossRef]
  50. Niu, B.; Shao, P.; Chen, H.; Sun, P. Structural and physiochemical characterization of novel hydrophobic packaging films based on pullulan derivatives for fruits preservation. Carbohydr. Polym. 2019, 208, 276–284. [Google Scholar] [CrossRef]
  51. Chu, Y.; Xu, T.; Gao, C.; Liu, X.; Zhang, N.; Feng, X.; Liu, X.; Shen, X.; Tang, X. Evaluations of physicochemical and biological properties of pullulan-based films incorporated with cinnamon essential oil and Tween 80. Int. J. Biol. Macromol. 2019, 122, 388–394. [Google Scholar] [CrossRef]
  52. Kim, J.-Y.; Choi, Y.-G.; Byul Kim, S.R.; Lim, S.-T. Humidity stability of tapioca starch–pullulan composite films. Food Hydrocoll. 2014, 41, 140–145. [Google Scholar] [CrossRef]
  53. Salama, H.E.; Abdel Aziz, M.S.; Sabaa, M.W. Novel biodegradable and antibacterial edible films based on alginate and chitosan biguanidine hydrochloride. Int. J. Biol. Macromol. 2018, 116, 443–450. [Google Scholar] [CrossRef] [PubMed]
  54. Li, K.; Zhu, J.; Guan, G.; Wu, H. Preparation of chitosan-sodium alginate films through layer-by-layer assembly and ferulic acid crosslinking: Film properties, characterization, and formation mechanism. Int. J. Biol. Macromol. 2019, 122, 485–492. [Google Scholar] [CrossRef] [PubMed]
  55. Molinaro, S.; Cruz-Romero, M.; Sensidoni, A.; Morris, M.; Lagazio, C.; Kerry, J.P. Combination of high-pressure treatment, mild heating and holding time effects as a means of improving the barrier properties of gelatin-based packaging films using response surface modeling. Innov. Food Sci. Emerg. Technol. 2015, 30, 15–23. [Google Scholar] [CrossRef]
  56. Hanani, Z.A.N.; Yee, F.C.; Nor-Khaizura, M.A.R. Effect of pomegranate (Punica granatum L.) peel powder on the antioxidant and antimicrobial properties of fish gelatin films as active packaging. Food Hydrocoll. 2019, 89, 253–259. [Google Scholar] [CrossRef]
  57. Chiumarelli, M.; Hubinger, M.D. Stability, solubility, mechanical and barrier properties of cassava starch – Carnauba wax edible coatings to preserve fresh-cut apples. Food Hydrocoll. 2012, 28, 59–67. [Google Scholar] [CrossRef]
  58. Torres, F.G.; Arroyo, J.J.; Troncoso, O.P. Bacterial cellulose nanocomposites: An all-nano type of material. Mater. Sci. Eng. C 2019, 98, 1277–1293. [Google Scholar] [CrossRef]
  59. Silva, R.D.; Sierakowski, M.R.; Bassani, H.P.; Zawadzki, S.F.; Pirich, C.L.; Ono, L.; de Freitas, R.A. Hydrophilicity improvement of mercerized bacterial cellulose films by polyethylene glycol graft. Int. J. Biol. Macromol. 2016, 86, 599–605. [Google Scholar] [CrossRef]
  60. Chang, H.-M.; Prasannan, A.; Tsai, H.-C.; Jhu, J.-J. Ex vivo evaluation of biodegradable poly(ɛ-caprolactone) films in digestive fluids. Appl. Surf. Sci. 2014, 313, 828–833. [Google Scholar] [CrossRef]
  61. Mosnáček, J.; Borská, K.; Danko, M.; Janigová, I. Photochemically promoted degradation of poly(ɛ-caprolactone) film. Mater. Chem. Phys. 2013, 140, 191–199. [Google Scholar] [CrossRef]
  62. Wang, Z.; Yao, Z.; Zhou, J.; He, M.; Jiang, Q.; Li, A.; Li, S.; Liu, M.; Luo, S.; Zhang, D. Improvement of polylactic acid film properties through the addition of cellulose nanocrystals isolated from waste cotton cloth. Int. J. Biol. Macromol. 2019, 129, 878–886. [Google Scholar] [CrossRef] [PubMed]
  63. Murariu, M.; Dubois, P. PLA composites: From production to properties. Adv. Drug Deliv. Rev. 2016, 107, 17–46. [Google Scholar] [CrossRef] [PubMed]
  64. Mhiri, S.; Mignard, N.; Abid, M.; Prochazka, F.; Majeste, J.-C.; Taha, M. Thermally reversible and biodegradable polyglycolic-acid-based networks. Eur. Polym. J. 2017, 88, 292–310. [Google Scholar] [CrossRef]
  65. You, Y.; Min, B.-M.; Lee, S.J.; Lee, T.S.; Park, W.H. In vitro degradation behavior of electrospun polyglycolide, polylactide, and poly(lactide-co-glycolide). J. Appl. Polym. Sci. 2005, 95, 193–200. [Google Scholar] [CrossRef]
  66. Fu, C.; Hu, X.; Yang, Z.; Shen, L.; Zheng, Z. Preparation and properties of waterborne bio-based polyurethane/siloxane cross-linked films by an in situ sol–gel process. Prog. Org. Coat. 2015, 84, 18–27. [Google Scholar] [CrossRef]
  67. Youssef, A.M.; El-Sayed, S.M. Bionanocomposites materials for food packaging applications: Concepts and future outlook. Carbohydr. Polym. 2018, 193, 19–27. [Google Scholar] [CrossRef]
  68. Nasrollahzadeh, M.; Issaabadi, Z.; Sajjadi, M.; Sajadi, S.M.; Atarod, M. Chapter 2—Types of Nanostructures. In Interface Science and Technology; Nasrollahzadeh, M., Sajadi, S.M., Sajjadi, M., Issaabadi, Z., Atarod, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; Volume 28, pp. 29–80. [Google Scholar]
  69. Fathi Achachlouei, B.; Zahedi, Y. Fabrication and characterization of CMC-based nanocomposites reinforced with sodium montmorillonite and TiO2 nanomaterials. Carbohydr. Polym. 2018, 199, 415–425. [Google Scholar] [CrossRef]
  70. Zhu, T.T.; Zhou, C.H.; Kabwe, F.B.; Wu, Q.Q.; Li, C.S.; Zhang, J.R. Exfoliation of montmorillonite and related properties of clay/polymer nanocomposites. Appl. Clay Sci. 2019, 169, 48–66. [Google Scholar] [CrossRef]
  71. Poonguzhali, R.; Basha, S.K.; Kumari, V.S. Synthesis and characterization of chitosan-PVP-nanocellulose composites for in-vitro wound dressing application. Int. J. Biol. Macromol. 2017, 105, 111–120. [Google Scholar] [CrossRef] [PubMed]
  72. Adel, A.; El-Shafei, A.; Ibrahim, A.; Al-Shemy, M. Extraction of oxidized nanocellulose from date palm (Phoenix dactylifera L.) sheath fibers: Influence of CI and CII polymorphs on the properties of chitosan/bionanocomposite films. Ind. Crops Prod. 2018, 124, 155–165. [Google Scholar] [CrossRef]
  73. Curvello, R.; Raghuwanshi, V.S.; Garnier, G. Engineering nanocellulose hydrogels for biomedical applications. Adv. Coll. Interface Sci. 2019, 267, 47–61. [Google Scholar] [CrossRef] [PubMed]
  74. Yu, S.; Xu, X.; Feng, J.; Liu, M.; Hu, K. Chitosan and chitosan coating nanoparticles for the treatment of brain disease. Int. J. Pharm. 2019, 560, 282–293. [Google Scholar] [CrossRef]
  75. Martins, G.O.; Segalla Petrônio, M.; Furuyama Lima, A.M.; Martinez Junior, A.M.; de Oliveira Tiera, V.A.; de Freitas Calmon, M.; Leite Vilamaior, P.S.; Han, S.W.; Tiera, M.J. Amphipathic Chitosans Improve the Physicochemical Properties of siRNA-Chitosan Nanoparticles at Physiological Conditions. Carbohydr. Polym. 2019. [Google Scholar] [CrossRef]
  76. Pan, C.; Qian, J.; Fan, J.; Guo, H.; Gou, L.; Yang, H.; Liang, C. Preparation nanoparticle by ionic cross-linked emulsified chitosan and its antibacterial activity. Colloids Surf. A Physicochem. Eng. Aspects 2019, 568, 362–370. [Google Scholar] [CrossRef]
  77. Divya, M.; Kiran, G.S.; Hassan, S.; Selvin, J. Biogenic synthesis and effect of silver nanoparticles (AgNPs) to combat catheter-related urinary tract infections. Biocatal. Agric. Biotechnol. 2019, 18, 101037. [Google Scholar] [CrossRef]
  78. Roy, S.; Shankar, S.; Rhim, J.-W. Melanin-mediated synthesis of silver nanoparticle and its use for the preparation of carrageenan-based antibacterial films. Food Hydrocoll. 2019, 88, 237–246. [Google Scholar] [CrossRef]
  79. Kadam, D.; Momin, B.; Palamthodi, S.; Lele, S.S. Physicochemical and functional properties of chitosan-based nano-composite films incorporated with biogenic silver nanoparticles. Carbohydr. Polym. 2019, 211, 124–132. [Google Scholar] [CrossRef]
  80. Farshchi, E.; Pirsa, S.; Roufegarinejad, L.; Alizadeh, M.; Rezazad, M. Photocatalytic/biodegradable film based on carboxymethyl cellulose, modified by gelatin and TiO2-Ag nanoparticles. Carbohydr. Polym. 2019. [Google Scholar] [CrossRef]
  81. Jamróz, E.; Kopel, P.; Juszczak, L.; Kawecka, A.; Bytesnikova, Z.; Milosavljević, V.; Kucharek, M.; Makarewicz, M.; Adam, V. Development and characterisation of furcellaran-gelatin films containing SeNPs and AgNPs that have antimicrobial activity. Food Hydrocoll. 2018, 83, 9–16. [Google Scholar] [CrossRef]
  82. Kalishwaralal, K.; Jeyabharathi, S.; Sundar, K.; Muthukumaran, A. A novel one-pot green synthesis of selenium nanoparticles and evaluation of its toxicity in zebrafish embryos. Artif. Cells Nanomed. Biotechnol. 2016, 44, 471–477. [Google Scholar] [CrossRef] [PubMed]
  83. Shankar, S.; Teng, X.; Rhim, J.-W. Properties and characterization of agar/CuNP bionanocomposite films prepared with different copper salts and reducing agents. Carbohydr. Polym. 2014, 114, 484–492. [Google Scholar] [CrossRef]
  84. Roy, S.; Rhim, J.-W.; Jaiswal, L. Bioactive agar-based functional composite film incorporated with copper sulfide nanoparticles. Food Hydrocoll. 2019, 93, 156–166. [Google Scholar] [CrossRef]
  85. Shankar, S.; Rhim, J.-W. Preparation of sulfur nanoparticle-incorporated antimicrobial chitosan films. Food Hydrocoll. 2018, 82, 116–123. [Google Scholar] [CrossRef]
  86. Natarajan, S.; Lakshmi, D.S.; Thiagarajan, V.; Mrudula, P.; Chandrasekaran, N.; Mukherjee, A. Antifouling and anti-algal effects of chitosan nanocomposite (TiO2/Ag) and pristine (TiO2 and Ag) films on marine microalgae Dunaliella salina. J. Environ. Chem. Eng. 2018, 6, 6870–6880. [Google Scholar] [CrossRef]
  87. Alizadeh-Sani, M.; Khezerlou, A.; Ehsani, A. Fabrication and characterization of the bionanocomposite film based on whey protein biopolymer loaded with TiO2 nanoparticles, cellulose nanofibers and rosemary essential oil. Ind. Crops Prod. 2018, 124, 300–315. [Google Scholar] [CrossRef]
  88. Jbeli, A.; Hamden, Z.; Bouattour, S.; Ferraria, A.M.; Conceição, D.S.; Ferreira, L.F.V.; Chehimi, M.M.; do Rego, A.M.B.; Rei Vilar, M.; Boufi, S. Chitosan-Ag-TiO2 films: An effective photocatalyst under visible light. Carbohydr. Polym. 2018, 199, 31–40. [Google Scholar] [CrossRef]
  89. Noshirvani, N.; Ghanbarzadeh, B.; Mokarram, R.R.; Hashemi, M.; Coma, V. Preparation and characterization of active emulsified films based on chitosan-carboxymethyl cellulose containing zinc oxide nano particles. Int. J. Biol. Macromol. 2017, 99, 530–538. [Google Scholar] [CrossRef] [PubMed]
  90. Abdeen, Z.I.; El Farargy, A.F.; Negm, N.A. Nanocomposite framework of chitosan/polyvinyl alcohol/ZnO: Preparation, characterization, swelling and antimicrobial evaluation. J. Mol. Liquids 2018, 250, 335–343. [Google Scholar] [CrossRef]
  91. Ponnamma, D.; Cabibihan, J.-J.; Rajan, M.; Pethaiah, S.S.; Deshmukh, K.; Gogoi, J.P.; Pasha, S.K.K.; Ahamed, M.B.; Krishnegowda, J.; Chandrashekar, B.N.; et al. Synthesis, optimization and applications of ZnO/polymer nanocomposites. Mater. Sci. Eng. C 2019, 98, 1210–1240. [Google Scholar] [CrossRef] [PubMed]
  92. Tripathi, R.; Narayan, A.; Bramhecha, I.; Sheikh, J. Development of multifunctional linen fabric using chitosan film as a template for immobilization of in-situ generated CeO2 nanoparticles. Int. J. Biol. Macromol. 2019, 121, 1154–1159. [Google Scholar] [CrossRef]
  93. Wang, C.; Murugadoss, V.; Kong, J.; He, Z.; Mai, X.; Shao, Q.; Chen, Y.; Guo, L.; Liu, C.; Angaiah, S.; et al. Overview of carbon nanostructures and nanocomposites for electromagnetic wave shielding. Carbon 2018, 140, 696–733. [Google Scholar] [CrossRef]
  94. Wu, Q.; Sun, Y.; Gao, J.; Dong, S.; Luo, G.; Li, H.; Zhao, L. Applications of hybrid organic–inorganic materials in chiral separation. TrAC Trends Anal. Chem. 2017, 95, 140–148. [Google Scholar] [CrossRef]
  95. Shankar, S.; Kasapis, S.; Rhim, J.-W. Alginate-based nanocomposite films reinforced with halloysite nanotubes functionalized by alkali treatment and zinc oxide nanoparticles. Int. J. Biol. Macromol. 2018, 118, 1824–1832. [Google Scholar] [CrossRef] [PubMed]
  96. Zhang, Y.; Chen, Y.; Zhang, H.; Zhang, B.; Liu, J. Potent antibacterial activity of a novel silver nanoparticle-halloysite nanotube nanocomposite powder. J. Inorg. Biochem. 2013, 118, 59–64. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, Y.; Zhang, Y.; Liu, J.; Zhang, H.; Wang, K. Preparation and antibacterial property of polyethersulfone ultrafiltration hybrid membrane containing halloysite nanotubes loaded with copper ions. Chem. Eng. J. 2012, 210, 298–308. [Google Scholar] [CrossRef]
  98. Ling, S.; Chen, W.; Fan, Y.; Zheng, K.; Jin, K.; Yu, H.; Buehler, M.J.; Kaplan, D.L. Biopolymer nanofibrils: Structure, modeling, preparation, and applications. Prog. Polym. Sci. 2018, 85, 1–56. [Google Scholar] [CrossRef]
  99. Abdul Khalil, H.P.S.; Bhat, A.H.; Ireana Yusra, A.F. Green composites from sustainable cellulose nanofibrils: A review. Carbohydr. Polym. 2012, 87, 963–979. [Google Scholar] [CrossRef]
  100. Phanthong, P.; Reubroycharoen, P.; Hao, X.; Xu, G.; Abudula, A.; Guan, G. Nanocellulose: Extraction and application. Carbon Resour. Convers. 2018, 1, 32–43. [Google Scholar] [CrossRef]
  101. Mishra, S.; Kharkar, P.S.; Pethe, A.M. Biomass and waste materials as potential sources of nanocrystalline cellulose: Comparative review of preparation methods (2016–Till date). Carbohydr. Polym. 2019, 207, 418–427. [Google Scholar] [CrossRef]
  102. Oun, A.A.; Rhim, J.-W. Isolation of cellulose nanocrystals from grain straws and their use for the preparation of carboxymethyl cellulose-based nanocomposite films. Carbohydr. Polym. 2016, 150, 187–200. [Google Scholar] [CrossRef] [PubMed]
  103. Niamsap, T.; Lam, N.T.; Sukyai, P. Production of hydroxyapatite-bacterial nanocellulose scaffold with assist of cellulose nanocrystals. Carbohydr. Polym. 2019, 205, 159–166. [Google Scholar] [CrossRef] [PubMed]
  104. Bao, Y.; Zhang, H.; Luan, Q.; Zheng, M.; Tang, H.; Huang, F. Fabrication of cellulose nanowhiskers reinforced chitosan-xylan nanocomposite films with antibacterial and antioxidant activities. Carbohydr. Polym. 2018, 184, 66–73. [Google Scholar] [CrossRef] [PubMed]
  105. Rouhi, M.; Razavi, S.H.; Mousavi, S.M. Optimization of crosslinked poly(vinyl alcohol) nanocomposite films for mechanical properties. Mater. Sci. Eng. C 2017, 71, 1052–1063. [Google Scholar] [CrossRef] [PubMed]
  106. Marín-Silva, D.A.; Rivero, S.; Pinotti, A. Chitosan-based nanocomposite matrices: Development and characterization. Int. J. Biol. Macromol. 2019, 123, 189–200. [Google Scholar] [CrossRef] [PubMed]
  107. Cacciotti, I.; Fortunati, E.; Puglia, D.; Kenny, J.M.; Nanni, F. Effect of silver nanoparticles and cellulose nanocrystals on electrospun poly(lactic) acid mats: Morphology, thermal properties and mechanical behavior. Carbohydr. Polym. 2014, 103, 22–31. [Google Scholar] [CrossRef]
  108. Olivera, S.; Muralidhara, H.B.; Venkatesh, K.; Guna, V.K.; Gopalakrishna, K.; Kumar, Y. Potential applications of cellulose and chitosan nanoparticles/composites in wastewater treatment: A review. Carbohydr. Polym. 2016, 153, 600–618. [Google Scholar] [CrossRef]
  109. Azarifar, M.; Ghanbarzadeh, B.; Sowti Khiabani, M.; Akhondzadeh Basti, A.; Abdulkhani, A.; Noshirvani, N.; Hosseini, M. The optimization of gelatin-CMC based active films containing chitin nanofiber and Trachyspermum ammi essential oil by response surface methodology. Carbohydr. Polym. 2019, 208, 457–468. [Google Scholar] [CrossRef]
  110. Hassan, E.A.; Hassan, M.L.; Abou-zeid, R.E.; El-Wakil, N.A. Novel nanofibrillated cellulose/chitosan nanoparticles nanocomposites films and their use for paper coating. Ind. Crops Prod. 2016, 93, 219–226. [Google Scholar] [CrossRef]
  111. Ma, Q.; Liang, T.; Cao, L.; Wang, L. Intelligent poly (vinyl alcohol)-chitosan nanoparticles-mulberry extracts films capable of monitoring pH variations. Int. J. Biol. Macromol. 2018, 108, 576–584. [Google Scholar] [CrossRef]
  112. Vahedikia, N.; Garavand, F.; Tajeddin, B.; Cacciotti, I.; Jafari, S.M.; Omidi, T.; Zahedi, Z. Biodegradable zein film composites reinforced with chitosan nanoparticles and cinnamon essential oil: Physical, mechanical, structural and antimicrobial attributes. Colloids Surf. B Biointerfaces 2019, 177, 25–32. [Google Scholar] [CrossRef] [PubMed]
  113. Naskar, S.; Sharma, S.; Kuotsu, K. Chitosan-based nanoparticles: An overview of biomedical applications and its preparation. J. Drug Delivery Sci. Technol. 2019, 49, 66–81. [Google Scholar] [CrossRef]
  114. Sengan, M.; Subramaniyan, S.B.; Arul Prakash, S.; Kamlekar, R.; Veerappan, A. Effective elimination of biofilm formed with waterborne pathogens using copper nanoparticles. Microb. Pathog. 2019, 127, 341–346. [Google Scholar] [CrossRef]
  115. Kumar, S.S.D.; Rajendran, N.K.; Houreld, N.N.; Abrahamse, H. Recent advances on silver nanoparticle and biopolymer-based biomaterials for wound healing applications. Int. J. Biol. Macromol. 2018, 115, 165–175. [Google Scholar] [CrossRef] [PubMed]
  116. Carbone, M.; Donia, D.T.; Sabbatella, G.; Antiochia, R. Silver nanoparticles in polymeric matrices for fresh food packaging. J. King Saud Univ. Sci. 2016, 28, 273–279. [Google Scholar] [CrossRef]
  117. Park, S.-B.; Lih, E.; Park, K.-S.; Joung, Y.K.; Han, D.K. Biopolymer-based functional composites for medical applications. Prog. Polym. Sci. 2017, 68, 77–105. [Google Scholar] [CrossRef]
  118. Khurana, A.; Tekula, S.; Saifi, M.A.; Venkatesh, P.; Godugu, C. Therapeutic applications of selenium nanoparticles. Biomed. Pharmacother. 2019, 111, 802–812. [Google Scholar] [CrossRef]
  119. Garcia, C.V.; Shin, G.H.; Kim, J.T. Metal oxide-based nanocomposites in food packaging: Applications, migration, and regulations. Trends Food Sci. Technol. 2018, 82, 21–31. [Google Scholar] [CrossRef]
  120. Król, A.; Pomastowski, P.; Rafińska, K.; Railean-Plugaru, V.; Buszewski, B. Zinc oxide nanoparticles: Synthesis, antiseptic activity and toxicity mechanism. Adv. Colloid Interface Sci. 2017, 249, 37–52. [Google Scholar] [CrossRef]
  121. Oleyaei, S.A.; Zahedi, Y.; Ghanbarzadeh, B.; Moayedi, A.A. Modification of physicochemical and thermal properties of starch films by incorporation of TiO2 nanoparticles. Int. J. Biol. Macromol. 2016, 89, 256–264. [Google Scholar] [CrossRef]
  122. Deka, B.K.; Kong, K.; Seo, J.; Kim, D.; Park, Y.-B.; Park, H.W. Controlled growth of CuO nanowires on woven carbon fibers and effects on the mechanical properties of woven carbon fiber/polyester composites. Compos. Part A Appl. Sci. Manuf. 2015, 69, 56–63. [Google Scholar] [CrossRef]
  123. Ebnalwaled, A.A.; Ismaiel, A.M. Developing novel UV shielding films based on PVA/Gelatin/0.01CuO nanocomposite: On the properties optimization using γ-irradiation. Measurement 2019, 134, 89–100. [Google Scholar] [CrossRef]
  124. Zhao, Q.; Ma, L.; Zhang, Q.; Wang, C.; Xu, X. SnO2-Based Nanomaterials: Synthesis and Application in Lithium-Ion Batteries and Supercapacitors. J. Nanomater. 2015, 2015, 805147. [Google Scholar] [CrossRef]
  125. Baniasad, A.; Ghorbani, M. Thermal stability enhancement of modified carboxymethyl cellulose films using SnO2 nanoparticles. Int. J. Biol. Macromol. 2016, 86, 901–906. [Google Scholar] [CrossRef]
  126. Bala, T.; Armstrong, G.; Laffir, F.; Thornton, R. Titania–silver and alumina–silver composite nanoparticles: Novel, versatile synthesis, reaction mechanism and potential antimicrobial application. J. Colloid Interface Sci. 2011, 356, 395–403. [Google Scholar] [CrossRef] [PubMed]
  127. Moradi, Z.; Esmaiili, M.; Almasi, H. Development and characterization of kefiran—Al2O3 nanocomposite films: Morphological, physical and mechanical properties. Int. J. Biol. Macromol. 2019, 122, 603–609. [Google Scholar] [CrossRef] [PubMed]
  128. Tong, Y.; Lin, Y.; Wang, S.; Song, M. A study of crystallisation of poly (ethylene oxide) and polypropylene on graphene surface. Polymer 2015, 73, 52–61. [Google Scholar] [CrossRef]
  129. Gan, S.; Zakaria, S.; Syed Jaafar, S.N. Enhanced mechanical properties of hydrothermal carbamated cellulose nanocomposite film reinforced with graphene oxide. Carbohydr. Polym. 2017, 172, 284–293. [Google Scholar] [CrossRef]
  130. Kwak, H.W.; Shin, M.; Yun, H.; Lee, K.H. Preparation of Silk Sericin/Lignin Blend Beads for the Removal of Hexavalent Chromium Ions. Int. J. Mol. Sci. 2016, 17, 1466. [Google Scholar] [CrossRef]
  131. Abdulkhani, A.; Daliri Sousefi, M.; Ashori, A.; Ebrahimi, G. Preparation and characterization of sodium carboxymethyl cellulose/silk fibroin/graphene oxide nanocomposite films. Polym. Test. 2016, 52, 218–224. [Google Scholar] [CrossRef]
  132. Kakaei, K.; Marzang, K. One – Step synthesis of nitrogen doped reduced graphene oxide with NiCo nanoparticles for ethanol oxidation in alkaline media. J. Colloid Interface Sci. 2016, 462, 148–153. [Google Scholar] [CrossRef] [PubMed]
  133. Liang, R.; Yan, D.; Tian, R.; Yu, X.; Shi, W.; Li, C.; Wei, M.; Evans, D.G.; Duan, X. Quantum Dots-Based Flexible Films and Their Application as the Phosphor in White Light-Emitting Diodes. Chem. Mater. 2014, 26, 2595–2600. [Google Scholar] [CrossRef]
  134. Guo, Y.; Ge, X.; Guan, J.; Wu, L.; Zhao, F.; Li, H.; Mu, X.; Jiang, Y.; Chen, A. A novel method for fabricating hybrid bio-based nanocomposites film with stable fluorescence containing CdTe quantum dots and montmorillonite-chitosan nanosheets. Carbohydr. Polym. 2016, 145, 13–19. [Google Scholar] [CrossRef]
  135. Javanbakht, S.; Namazi, H. Solid state photoluminescence thermoplastic starch film containing graphene quantum dots. Carbohydr. Polym. 2017, 176, 220–226. [Google Scholar] [CrossRef] [PubMed]
  136. Hosseini, S.F.; Rezaei, M.; Zandi, M.; Farahmandghavi, F. Development of bioactive fish gelatin/chitosan nanoparticles composite films with antimicrobial properties. Food Chem. 2016, 194, 1266–1274. [Google Scholar] [CrossRef] [PubMed]
  137. Chen, Y.; Liu, C.; Chang, P.R.; Anderson, D.P.; Huneault, M.A. Pea starch-based composite films with pea hull fibers and pea hull fiber-derived nanowhiskers. Polym. Eng. Sci. 2009, 49, 369–378. [Google Scholar] [CrossRef]
  138. Noshirvani, N.; Hong, W.; Ghanbarzadeh, B.; Fasihi, H.; Montazami, R. Study of cellulose nanocrystal doped starch-polyvinyl alcohol bionanocomposite films. Int. J. Biol. Macromol. 2018, 107, 2065–2074. [Google Scholar] [CrossRef] [PubMed]
  139. Hasheminya, S.-M.; Mokarram, R.R.; Ghanbarzadeh, B.; Hamishekar, H.; Kafil, H.S.; Dehghannya, J. Influence of simultaneous application of copper oxide nanoparticles and Satureja Khuzestanica essential oil on properties of kefiran–carboxymethyl cellulose films. Polym. Test. 2019, 73, 377–388. [Google Scholar] [CrossRef]
  140. Alipoormazandarani, N.; Ghazihoseini, S.; Mohammadi Nafchi, A. Preparation and characterization of novel bionanocomposite based on soluble soybean polysaccharide and halloysite nanoclay. Carbohydr. Polym. 2015, 134, 745–751. [Google Scholar] [CrossRef]
  141. Zolfi, M.; Khodaiyan, F.; Mousavi, M.; Hashemi, M. The improvement of characteristics of biodegradable films made from kefiran–whey protein by nanoparticle incorporation. Carbohydr. Polym. 2014, 109, 118–125. [Google Scholar] [CrossRef]
  142. Hassannia-Kolaee, M.; Khodaiyan, F.; Pourahmad, R.; Shahabi-Ghahfarrokhi, I. Development of ecofriendly bionanocomposite: Whey protein isolate/pullulan films with nano-SiO2. Int. J. Biol. Macromol. 2016, 86, 139–144. [Google Scholar] [CrossRef]
  143. Qin, Y.; Zhang, S.; Yu, J.; Yang, J.; Xiong, L.; Sun, Q. Effects of chitin nano-whiskers on the antibacterial and physicochemical properties of maize starch films. Carbohydr. Polym. 2016, 147, 372–378. [Google Scholar] [CrossRef] [PubMed]
  144. Jafari, H.; Pirouzifard, M.; Khaledabad, M.A.; Almasi, H. Effect of chitin nanofiber on the morphological and physical properties of chitosan/silver nanoparticle bionanocomposite films. Int. J. Biol. Macromol. 2016, 92, 461–466. [Google Scholar] [CrossRef] [PubMed]
  145. Devi, N.; Dutta, J. Development and in vitro characterization of chitosan/starch/halloysite nanotubes ternary nanocomposite films. Int. J. Biol. Macromol. 2019, 127, 222–231. [Google Scholar] [CrossRef]
  146. Vejdan, A.; Ojagh, S.M.; Adeli, A.; Abdollahi, M. Effect of TiO2 nanoparticles on the physico-mechanical and ultraviolet light barrier properties of fish gelatin/agar bilayer film. LWT Food Sci. Technol. 2016, 71, 88–95. [Google Scholar] [CrossRef]
  147. Rostamzad, H.; Paighambari, S.Y.; Shabanpour, B.; Ojagh, S.M.; Mousavi, S.M. Improvement of fish protein film with nanoclay and transglutaminase for food packaging. Food Packag. Shelf Life 2016, 7, 1–7. [Google Scholar] [CrossRef]
  148. Davachi, S.M.; Shekarabi, A.S. Preparation and characterization of antibacterial, eco-friendly edible nanocomposite films containing Salvia macrosiphon and nanoclay. Int. J. Biol. Macromol. 2018, 113, 66–72. [Google Scholar] [CrossRef]
  149. Cazón, P.; Velazquez, G.; Ramírez, J.A.; Vázquez, M. Polysaccharide-based films and coatings for food packaging: A review. Food Hydrocoll. 2017, 68, 136–148. [Google Scholar] [CrossRef]
  150. Shankar, S.; Rhim, J.-W. Preparation and characterization of agar/lignin/silver nanoparticles composite films with ultraviolet light barrier and antibacterial properties. Food Hydrocoll. 2017, 71, 76–84. [Google Scholar] [CrossRef]
  151. Lomate, G.B.; Dandi, B.; Mishra, S. Development of antimicrobial LDPE/Cu nanocomposite food packaging film for extended shelf life of peda. Food Packag. Shelf Life 2018, 16, 211–219. [Google Scholar] [CrossRef]
  152. Han, Y.; Yu, M.; Wang, L. Soy protein isolate nanocomposites reinforced with nanocellulose isolated from licorice residue: Water sensitivity and mechanical strength. Ind. Crops Prod. 2018, 117, 252–259. [Google Scholar] [CrossRef]
  153. Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S. Development and characterization of sugar palm nanocrystalline cellulose reinforced sugar palm starch bionanocomposites. Carbohydr. Polym. 2018, 202, 186–202. [Google Scholar] [CrossRef]
  154. Zahedi, Y.; Fathi-Achachlouei, B.; Yousefi, A.R. Physical and mechanical properties of hybrid montmorillonite/zinc oxide reinforced carboxymethyl cellulose nanocomposites. Int. J. Biol. Macromol. 2018, 108, 863–873. [Google Scholar] [CrossRef]
  155. Rhim, J.-W. Effect of clay contents on mechanical and water vapor barrier properties of agar-based nanocomposite films. Carbohydr. Polym. 2011, 86, 691–699. [Google Scholar] [CrossRef]
  156. Klangmuang, P.; Sothornvit, R. Combination of beeswax and nanoclay on barriers, sorption isotherm and mechanical properties of hydroxypropyl methylcellulose-based composite films. LWT Food Sci. Technol. 2016, 65, 222–227. [Google Scholar] [CrossRef]
  157. Olivera, N.; Rouf, T.B.; Bonilla, J.C.; Carriazo, J.G.; Dianda, N.; Kokini, J.L. Effect of LAPONITE® addition on the mechanical, barrier and surface properties of novel biodegradable kafirin nanocomposite films. J. Food Eng. 2019, 245, 24–32. [Google Scholar] [CrossRef]
  158. Valencia, G.A.; Luciano, C.G.; Lourenço, R.V.; Bittante, A.M.Q.B.; do Amaral Sobral, P.J. Morphological and physical properties of nano-biocomposite films based on collagen loaded with laponite®. Food Packag. Shelf Life 2019, 19, 24–30. [Google Scholar] [CrossRef]
  159. Rodríguez, F.J.; Galotto, M.J.; Guarda, A.; Bruna, J.E. Modification of cellulose acetate films using nanofillers based on organoclays. J. Food Eng. 2012, 110, 262–268. [Google Scholar] [CrossRef]
  160. Shankar, S.; Teng, X.; Li, G.; Rhim, J.-W. Preparation, characterization, and antimicrobial activity of gelatin/ZnO nanocomposite films. Food Hydrocoll. 2015, 45, 264–271. [Google Scholar] [CrossRef]
  161. Silva, J.M.; Barud, H.S.; Meneguin, A.B.; Constantino, V.R.L.; Ribeiro, S.J.L. Inorganic-organic bio-nanocomposite films based on Laponite and Cellulose Nanofibers (CNF). Appl. Clay Sci. 2019, 168, 428–435. [Google Scholar] [CrossRef]
  162. Indumathi, M.P.; Saral Sarojini, K.; Rajarajeswari, G.R. Antimicrobial and biodegradable chitosan/cellulose acetate phthalate/ZnO nano composite films with optimal oxygen permeability and hydrophobicity for extending the shelf life of black grape fruits. Int. J. Biol. Macromol. 2019. [Google Scholar] [CrossRef]
  163. Wu, Z.; Deng, W.; Luo, J.; Deng, D. Multifunctional nano-cellulose composite films with grape seed extracts and immobilized silver nanoparticles. Carbohydr. Polym. 2019, 205, 447–455. [Google Scholar] [CrossRef]
  164. Sadegh-Hassani, F.; Mohammadi Nafchi, A. Preparation and characterization of bionanocomposite films based on potato starch/halloysite nanoclay. Int. J. Biol. Macromol. 2014, 67, 458–462. [Google Scholar] [CrossRef] [PubMed]
  165. Franceschini, I.; Selmin, F.; Pagani, S.; Minghetti, P.; Cilurzo, F. Nanofiller for the mechanical reinforcement of maltodextrins orodispersible films. Carbohydr. Polym. 2016, 136, 676–681. [Google Scholar] [CrossRef] [PubMed]
  166. Vaezi, K.; Asadpour, G.; Sharifi, H. Effect of ZnO nanoparticles on the mechanical, barrier and optical properties of thermoplastic cationic starch/montmorillonite biodegradable films. Int. J. Biol. Macromol. 2019, 124, 519–529. [Google Scholar] [CrossRef] [PubMed]
  167. Wu, C.; Li, Y.; Du, Y.; Wang, L.; Tong, C.; Hu, Y.; Pang, J.; Yan, Z. Preparation and characterization of konjac glucomannan-based bionanocomposite film for active food packaging. Food Hydrocoll. 2019, 89, 682–690. [Google Scholar] [CrossRef]
  168. Memiş, S.; Tornuk, F.; Bozkurt, F.; Durak, M.Z. Production and characterization of a new biodegradable fenugreek seed gum based active nanocomposite film reinforced with nanoclays. Int. J. Biol. Macromol. 2017, 103, 669–675. [Google Scholar] [CrossRef] [PubMed]
  169. Nouri, A.; Yaraki, M.T.; Ghorbanpour, M.; Agarwal, S.; Gupta, V.K. Enhanced Antibacterial effect of chitosan film using Montmorillonite/CuO nanocomposite. Int. J. Biol. Macromol. 2018, 109, 1219–1231. [Google Scholar] [CrossRef] [PubMed]
  170. El Miri, N.; Abdelouahdi, K.; Barakat, A.; Zahouily, M.; Fihri, A.; Solhy, A.; El Achaby, M. Bio-nanocomposite films reinforced with cellulose nanocrystals: Rheology of film-forming solutions, transparency, water vapor barrier and tensile properties of films. Carbohydr. Polym. 2015, 129, 156–167. [Google Scholar] [CrossRef]
  171. Farahnaky, A.; Dadfar, S.M.M.; Shahbazi, M. Physical and mechanical properties of gelatin–clay nanocomposite. J. Food Eng. 2014, 122, 78–83. [Google Scholar] [CrossRef]
  172. Popescu, M.-C.; Dogaru, B.-I.; Goanta, M.; Timpu, D. Structural and morphological evaluation of CNC reinforced PVA/Starch biodegradable films. Int. J. Biol. Macromol. 2018, 116, 385–393. [Google Scholar] [CrossRef]
  173. Arfat, Y.A.; Ahmed, J.; Hiremath, N.; Auras, R.; Joseph, A. Thermo-mechanical, rheological, structural and antimicrobial properties of bionanocomposite films based on fish skin gelatin and silver-copper nanoparticles. Food Hydrocoll. 2017, 62, 191–202. [Google Scholar] [CrossRef]
  174. Marvizadeh, M.M.; Oladzadabbasabadi, N.; Mohammadi Nafchi, A.; Jokar, M. Preparation and characterization of bionanocomposite film based on tapioca starch/bovine gelatin/nanorod zinc oxide. Int. J. Biol. Macromol. 2017, 99, 1–7. [Google Scholar] [CrossRef]
  175. Vicentini, N.; Gatti, T.; Salerno, M.; Hernandez Gomez, Y.S.; Bellon, M.; Gallio, S.; Marega, C.; Filippini, F.; Menna, E. Effect of different functionalized carbon nanostructures as fillers on the physical properties of biocompatible poly(l-lactic acid) composites. Mater. Chem. Phys. 2018, 214, 265–276. [Google Scholar] [CrossRef]
  176. Ahmed, J.; Mulla, M.; Arfat, Y.A.; Thai T, L.A. Mechanical, thermal, structural and barrier properties of crab shell chitosan/graphene oxide composite films. Food Hydrocoll. 2017, 71, 141–148. [Google Scholar] [CrossRef]
  177. Son, Y.-R.; Rhee, K.Y.; Park, S.-J. Influence of reduced graphene oxide on mechanical behaviors of sodium carboxymethyl cellulose. Compos. Part B Eng. 2015, 83, 36–42. [Google Scholar] [CrossRef]
  178. El Achaby, M.; Kassab, Z.; Barakat, A.; Aboulkas, A. Alfa fibers as viable sustainable source for cellulose nanocrystals extraction: Application for improving the tensile properties of biopolymer nanocomposite films. Ind. Crops Prod. 2018, 112, 499–510. [Google Scholar] [CrossRef]
  179. Kassab, Z.; Aziz, F.; Hannache, H.; Ben Youcef, H.; El Achaby, M. Improved mechanical properties of k-carrageenan-based nanocomposite films reinforced with cellulose nanocrystals. Int. J. Biol. Macromol. 2019, 123, 1248–1256. [Google Scholar] [CrossRef]
  180. Kanmani, P.; Rhim, J.-W. Properties and characterization of bionanocomposite films prepared with various biopolymers and ZnO nanoparticles. Carbohydr. Polym. 2014, 106, 190–199. [Google Scholar] [CrossRef] [PubMed]
  181. Mlalila, N.; Hilonga, A.; Swai, H.; Devlieghere, F.; Ragaert, P. Antimicrobial packaging based on starch, poly(3-hydroxybutyrate) and poly(lactic-co-glycolide) materials and application challenges. Trends Food Sci. Technol. 2018, 74, 1–11. [Google Scholar] [CrossRef]
  182. Ma, Z.; Garrido-Maestu, A.; Jeong, K.C. Application, mode of action, and in vivo activity of chitosan and its micro- and nanoparticles as antimicrobial agents: A review. Carbohydr. Polym. 2017, 176, 257–265. [Google Scholar] [CrossRef] [PubMed]
  183. Lu, C.; Brauer, M.J.; Botstein, D. Slow growth induces heat-shock resistance in normal and respiratory-deficient yeast. Mol. Biol. Cell 2009, 20, 891–903. [Google Scholar] [CrossRef] [PubMed]
  184. Salari, M.; Sowti Khiabani, M.; Rezaei Mokarram, R.; Ghanbarzadeh, B.; Samadi Kafil, H. Development and evaluation of chitosan based active nanocomposite films containing bacterial cellulose nanocrystals and silver nanoparticles. Food Hydrocoll. 2018, 84, 414–423. [Google Scholar] [CrossRef]
  185. Mohammadi Nafchi, A.; Moradpour, M.; Saeidi, M.; Alias, A.K. Effects of nanorod-rich ZnO on rheological, sorption isotherm, and physicochemical properties of bovine gelatin films. LWT Food Sci. Technol. 2014, 58, 142–149. [Google Scholar] [CrossRef]
  186. Roy Choudhury, S.; Goswami, A. Supramolecular reactive sulphur nanoparticles: A novel and efficient antimicrobial agent. J. Appl. Microbiol. 2013, 114, 1–10. [Google Scholar] [CrossRef] [PubMed]
  187. Besinis, A.; De Peralta, T.; Handy, R.D. The antibacterial effects of silver, titanium dioxide and silica dioxide nanoparticles compared to the dental disinfectant chlorhexidine on Streptococcus mutans using a suite of bioassays. Nanotoxicology 2014, 8, 1–16. [Google Scholar] [CrossRef] [PubMed]
  188. Seil, J.T.; Webster, T.J. Antimicrobial applications of nanotechnology: Methods and literature. Int. J. Nanomed. 2012, 7, 2767–2781. [Google Scholar] [CrossRef]
  189. Sothornvit, R.; Rhim, J.-W.; Hong, S.-I. Effect of nano-clay type on the physical and antimicrobial properties of whey protein isolate/clay composite films. J. Food Eng. 2009, 91, 468–473. [Google Scholar] [CrossRef]
  190. Rhim, J.-W.; Hong, S.-I.; Ha, C.-S. Tensile, water vapor barrier and antimicrobial properties of PLA/nanoclay composite films. LWT Food Sci. Technol. 2009, 42, 612–617. [Google Scholar] [CrossRef]
  191. Giannakas, A.; Vlacha, M.; Salmas, C.; Leontiou, A.; Katapodis, P.; Stamatis, H.; Barkoula, N.-M.; Ladavos, A. Preparation, characterization, mechanical, barrier and antimicrobial properties of chitosan/PVOH/clay nanocomposites. Carbohydr. Polym. 2016, 140, 408–415. [Google Scholar] [CrossRef] [PubMed]
  192. Shankar, S.; Reddy, J.P.; Rhim, J.-W.; Kim, H.-Y. Preparation, characterization, and antimicrobial activity of chitin nanofibrils reinforced carrageenan nanocomposite films. Carbohydr. Polym. 2015, 117, 468–475. [Google Scholar] [CrossRef] [PubMed]
  193. Nafchi, A.M.; Alias, A.K.; Mahmud, S.; Robal, M. Antimicrobial, rheological, and physicochemical properties of sago starch films filled with nanorod-rich zinc oxide. J. Food Eng. 2012, 113, 511–519. [Google Scholar] [CrossRef]
  194. Chen, J.; Wang, X.; Han, H. A new function of graphene oxide emerges: Inactivating phytopathogenic bacterium Xanthomonas oryzae pv. Oryzae. J. Nanoparticle Res. 2013, 15, 1658. [Google Scholar] [CrossRef]
  195. Yousefi, M.; Dadashpour, M.; Hejazi, M.; Hasanzadeh, M.; Behnam, B.; de la Guardia, M.; Shadjou, N.; Mokhtarzadeh, A. Anti-bacterial activity of graphene oxide as a new weapon nanomaterial to combat multidrug-resistance bacteria. Mater. Sci. Eng. C 2017, 74, 568–581. [Google Scholar] [CrossRef] [PubMed]
  196. Hu, X.; Jia, X.; Zhi, C.; Jin, Z.; Miao, M. Improving the properties of starch-based antimicrobial composite films using ZnO-chitosan nanoparticles. Carbohydr. Polym. 2019. [Google Scholar] [CrossRef] [PubMed]
  197. Usman, A.; Hussain, Z.; Riaz, A.; Khan, A.N. Enhanced mechanical, thermal and antimicrobial properties of poly(vinyl alcohol)/graphene oxide/starch/silver nanocomposites films. Carbohydr. Polym. 2016, 153, 592–599. [Google Scholar] [CrossRef]
  198. Pal, N.; Dubey, P.; Gopinath, P.; Pal, K. Combined effect of cellulose nanocrystal and reduced graphene oxide into poly-lactic acid matrix nanocomposite as a scaffold and its anti-bacterial activity. Int. J. Biol. Macromol. 2017, 95, 94–105. [Google Scholar] [CrossRef]
  199. Kanmani, P.; Rhim, J.-W. Physicochemical properties of gelatin/silver nanoparticle antimicrobial composite films. Food Chem. 2014, 148, 162–169. [Google Scholar] [CrossRef]
  200. Kumar, S.; Shukla, A.; Baul, P.P.; Mitra, A.; Halder, D. Biodegradable hybrid nanocomposites of chitosan/gelatin and silver nanoparticles for active food packaging applications. Food Packag. Shelf Life 2018, 16, 178–184. [Google Scholar] [CrossRef]
  201. Shah, A.; Hussain, I.; Murtaza, G. Chemical synthesis and characterization of chitosan/silver nanocomposites films and their potential antibacterial activity. Int. J. Biol. Macromol. 2018, 116, 520–529. [Google Scholar] [CrossRef]
  202. Lin, S.; Chen, L.; Huang, L.; Cao, S.; Luo, X.; Liu, K. Novel antimicrobial chitosan–cellulose composite films bioconjugated with silver nanoparticles. Ind. Crops Prod. 2015, 70, 395–403. [Google Scholar] [CrossRef]
  203. Hajji, S.; Salem, R.B.S.-B.; Hamdi, M.; Jellouli, K.; Ayadi, W.; Nasri, M.; Boufi, S. Nanocomposite films based on chitosan–poly(vinyl alcohol) and silver nanoparticles with high antibacterial and antioxidant activities. Process Saf. Environ. Prot. 2017, 111, 112–121. [Google Scholar] [CrossRef]
  204. Youssef, A.M.; Abdel-Aziz, M.S.; El-Sayed, S.M. Chitosan nanocomposite films based on Ag-NP and Au-NP biosynthesis by Bacillus Subtilis as packaging materials. Int. J. Biol. Macromol. 2014, 69, 185–191. [Google Scholar] [CrossRef]
  205. Dairi, N.; Ferfera-Harrar, H.; Ramos, M.; Garrigós, M.C. Cellulose acetate/AgNPs-organoclay and/or thymol nano-biocomposite films with combined antimicrobial/antioxidant properties for active food packaging use. Int. J. Biol. Macromol. 2019, 121, 508–523. [Google Scholar] [CrossRef] [PubMed]
  206. Sarwar, M.S.; Niazi, M.B.K.; Jahan, Z.; Ahmad, T.; Hussain, A. Preparation and characterization of PVA/nanocellulose/Ag nanocomposite films for antimicrobial food packaging. Carbohydr. Polym. 2018, 184, 453–464. [Google Scholar] [CrossRef] [PubMed]
  207. Marín, T.; Montoya, P.; Arnache, O.; Pinal, R.; Calderón, J. Development of magnetite nanoparticles/gelatin composite films for triggering drug release by an external magnetic field. Mater. Des. 2018, 152, 78–87. [Google Scholar] [CrossRef]
  208. Kloster, G.A.; Muraca, D.; Moscoso Londoño, O.; Knobel, M.; Marcovich, N.E.; Mosiewicki, M.A. Structural analysis of magnetic nanocomposites based on chitosan. Polym. Test. 2018, 72, 202–213. [Google Scholar] [CrossRef]
  209. He, Q.; Zhang, Y.; Cai, X.; Wang, S. Fabrication of gelatin–TiO2 nanocomposite film and its structural, antibacterial and physical properties. Int. J. Biol. Macromol. 2016, 84, 153–160. [Google Scholar] [CrossRef]
  210. Siripatrawan, U.; Kaewklin, P. Fabrication and characterization of chitosan-titanium dioxide nanocomposite film as ethylene scavenging and antimicrobial active food packaging. Food Hydrocoll. 2018, 84, 125–134. [Google Scholar] [CrossRef]
  211. Youssef, A.M.; El-Sayed, S.M.; El-Sayed, H.S.; Salama, H.H.; Dufresne, A. Enhancement of Egyptian soft white cheese shelf life using a novel chitosan/carboxymethyl cellulose/zinc oxide bionanocomposite film. Carbohydr. Polym. 2016, 151, 9–19. [Google Scholar] [CrossRef]
  212. Qiu, B.; Xu, X.-F.; Deng, R.-H.; Xia, G.-Q.; Shang, X.-F.; Zhou, P.-H. Construction of chitosan/ZnO nanocomposite film by in situ precipitation. Int. J. Biol. Macromol. 2019, 122, 82–87. [Google Scholar] [CrossRef] [PubMed]
  213. Saral Sarojini, K.; Indumathi, M.P.; Rajarajeswari, G.R. Mahua oil-based polyurethane/chitosan/nano ZnO composite films for biodegradable food packaging applications. Int. J. Biol. Macromol. 2019, 124, 163–174. [Google Scholar] [CrossRef]
  214. Kumar, S.; Krishnakumar, B.; Sobral, A.J.F.N.; Koh, J. Bio-based (chitosan/PVA/ZnO) nanocomposites film: Thermally stable and photoluminescence material for removal of organic dye. Carbohydr. Polym. 2019, 205, 559–564. [Google Scholar] [CrossRef]
  215. Oun, A.A.; Rhim, J.-W. Carrageenan-based hydrogels and films: Effect of ZnO and CuO nanoparticles on the physical, mechanical, and antimicrobial properties. Food Hydrocoll. 2017, 67, 45–53. [Google Scholar] [CrossRef]
  216. Yadav, M. Study on thermal and mechanical properties of cellulose/iron oxide bionanocomposites film. Compos. Commun. 2018, 10, 1–5. [Google Scholar] [CrossRef]
  217. Swaroop, C.; Shukla, M. Mechanical, Optical and Antibacterial Properties of Polylactic Acid/Polyethylene Glycol Films Reinforced with MgO Nanoparticles. Mater. Today Proc. 2018, 5, 20711–20718. [Google Scholar] [CrossRef]
  218. Jafarzadeh, S.; Ariffin, F.; Mahmud, S.; Alias, A.K.; Hosseini, S.F.; Ahmad, M. Improving the physical and protective functions of semolina films by embedding a blend nanofillers (ZnO-nr and nano-kaolin). Food Packag. Shelf Life 2017, 12, 66–75. [Google Scholar] [CrossRef]
  219. Ejaz, M.; Arfat, Y.A.; Mulla, M.; Ahmed, J. Zinc oxide nanorods/clove essential oil incorporated Type B gelatin composite films and its applicability for shrimp packaging. Food Packag. Shelf Life 2018, 15, 113–121. [Google Scholar] [CrossRef]
  220. Akbariazam, M.; Ahmadi, M.; Javadian, N.; Mohammadi Nafchi, A. Fabrication and characterization of soluble soybean polysaccharide and nanorod-rich ZnO bionanocomposite. Int. J. Biol. Macromol. 2016, 89, 369–375. [Google Scholar] [CrossRef]
  221. Abutalib, M.M. Effect of zinc oxide nanorods on the structural, thermal, dielectric and electrical properties of polyvinyl alcohol/carboxymethyle cellulose composites. Phys. B Condensed Matter 2019, 557, 108–116. [Google Scholar] [CrossRef]
  222. Perumal, A.B.; Sellamuthu, P.S.; Nambiar, R.B.; Sadiku, E.R. Development of polyvinyl alcohol/chitosan bio-nanocomposite films reinforced with cellulose nanocrystals isolated from rice straw. Appl. Surf. Sci. 2018, 449, 591–602. [Google Scholar] [CrossRef]
  223. Mujtaba, M.; Salaberria, A.M.; Andres, M.A.; Kaya, M.; Gunyakti, A.; Labidi, J. Utilization of flax (Linum usitatissimum) cellulose nanocrystals as reinforcing material for chitosan films. Int. J. Biol. Macromol. 2017, 104, 944–952. [Google Scholar] [CrossRef] [PubMed]
  224. Rong, S.Y.; Mubarak, N.M.; Tanjung, F.A. Structure-property relationship of cellulose nanowhiskers reinforced chitosan biocomposite films. J. Environ. Chem. Eng. 2017, 5, 6132–6136. [Google Scholar] [CrossRef]
  225. Tang, Y.; Zhang, X.; Zhao, R.; Guo, D.; Zhang, J. Preparation and properties of chitosan/guar gum/nanocrystalline cellulose nanocomposite films. Carbohydr. Polym. 2018, 197, 128–136. [Google Scholar] [CrossRef] [PubMed]
  226. Oun, A.A.; Rhim, J.-W. Preparation and characterization of sodium carboxymethyl cellulose/cotton linter cellulose nanofibril composite films. Carbohydr. Polym. 2015, 127, 101–109. [Google Scholar] [CrossRef]
  227. El Achaby, M.; El Miri, N.; Hannache, H.; Gmouh, S.; Ben youcef, H.; Aboulkas, A. Production of cellulose nanocrystals from vine shoots and their use for the development of nanocomposite materials. Int. J. Biol. Macromol. 2018, 117, 592–600. [Google Scholar] [CrossRef] [PubMed]
  228. González, A.; Gastelú, G.; Barrera, G.N.; Ribotta, P.D.; Álvarez Igarzabal, C.I. Preparation and characterization of soy protein films reinforced with cellulose nanofibers obtained from soybean by-products. Food Hydrocoll. 2019, 89, 758–764. [Google Scholar] [CrossRef]
  229. Ma, X.; Cheng, Y.; Qin, X.; Guo, T.; Deng, J.; Liu, X. Hydrophilic modification of cellulose nanocrystals improves the physicochemical properties of cassava starch-based nanocomposite films. LWT 2017, 86, 318–326. [Google Scholar] [CrossRef]
  230. Chi, K.; Catchmark, J.M. Improved eco-friendly barrier materials based on crystalline nanocellulose/chitosan/carboxymethyl cellulose polyelectrolyte complexes. Food Hydrocoll. 2018, 80, 195–205. [Google Scholar] [CrossRef]
  231. Koosha, M.; Hamedi, S. Intelligent Chitosan/PVA nanocomposite films containing black carrot anthocyanin and bentonite nanoclays with improved mechanical, thermal and antibacterial properties. Prog. Org. Coat. 2019, 127, 338–347. [Google Scholar] [CrossRef]
  232. Wang, L.-F.; Rhim, J.-W. Functionalization of halloysite nanotubes for the preparation of carboxymethyl cellulose-based nanocomposite films. Appl. Clay Sci. 2017, 150, 138–146. [Google Scholar] [CrossRef]
  233. Biddeci, G.; Cavallaro, G.; Di Blasi, F.; Lazzara, G.; Massaro, M.; Milioto, S.; Parisi, F.; Riela, S.; Spinelli, G. Halloysite nanotubes loaded with peppermint essential oil as filler for functional biopolymer film. Carbohydr. Polym. 2016, 152, 548–557. [Google Scholar] [CrossRef] [PubMed]
  234. Kanmani, P.; Rhim, J.-W. Physical, mechanical and antimicrobial properties of gelatin based active nanocomposite films containing AgNPs and nanoclay. Food Hydrocoll. 2014, 35, 644–652. [Google Scholar] [CrossRef]
  235. Wang, H.; Wang, L.; Ye, S.; Song, X. Construction of Bi2WO6–TiO2/starch nanocomposite films for visible-light catalytic degradation of ethylene. Food Hydrocoll. 2019, 88, 92–100. [Google Scholar] [CrossRef]
  236. Meira, S.M.M.; Zehetmeyer, G.; Scheibel, J.M.; Werner, J.O.; Brandelli, A. Starch-halloysite nanocomposites containing nisin: Characterization and inhibition of Listeria monocytogenes in soft cheese. LWT Food Sci. Technol. 2016, 68, 226–234. [Google Scholar] [CrossRef]
  237. Wilpiszewska, K.; Antosik, A.K.; Spychaj, T. Novel hydrophilic carboxymethyl starch/montmorillonite nanocomposite films. Carbohydr. Polym. 2015, 128, 82–89. [Google Scholar] [CrossRef]
  238. Abdollahi, M.; Alboofetileh, M.; Rezaei, M.; Behrooz, R. Comparing physico-mechanical and thermal properties of alginate nanocomposite films reinforced with organic and/or inorganic nanofillers. Food Hydrocoll. 2013, 32, 416–424. [Google Scholar] [CrossRef]
  239. Mallakpour, S.; Nezamzadeh Ezhieh, A. Preparation and characterization of chitosan-poly(vinyl alcohol) nanocomposite films embedded with functionalized multi-walled carbon nanotube. Carbohydr. Polym. 2017, 166, 377–386. [Google Scholar] [CrossRef]
  240. Bibi, S.; Jamil, A.; Yasin, T.; Rafiq, M.A.; Nawaz, M.; Price, G.J. Ultrasound promoted synthesis and properties of chitosan nanocomposites containing carbon nanotubes and silver nanoparticles. Eur. Polym. J. 2018, 105, 297–303. [Google Scholar] [CrossRef]
  241. Ebrahimzadeh, S.; Ghanbarzadeh, B.; Hamishehkar, H. Physical properties of carboxymethyl cellulose based nano-biocomposites with Graphene nano-platelets. Int. J. Biol. Macromol. 2016, 84, 16–23. [Google Scholar] [CrossRef]
  242. He, Y.; Wang, X.; Wu, D.; Gong, Q.; Qiu, H.; Liu, Y.; Wu, T.; Ma, J.; Gao, J. Biodegradable amylose films reinforced by graphene oxide and polyvinyl alcohol. Mater. Chem. Phys. 2013, 142, 1–11. [Google Scholar] [CrossRef]
  243. Roy, S.; Rhim, J.-W. Preparation of carrageenan-based functional nanocomposite films incorporated with melanin nanoparticles. Colloids Surf. B Biointerfaces 2019, 176, 317–324. [Google Scholar] [CrossRef] [PubMed]
  244. Yun, Y.-H.; Youn, H.-G.; Shin, J.-Y.; Yoon, S.-D. Preparation of functional chitosan-based nanocomposite films containing ZnS nanoparticles. Int. J. Biol. Macromol. 2017, 104, 1150–1157. [Google Scholar] [CrossRef] [PubMed]
  245. Kundu, S.; Das, A.; Basu, A.; Abdullah, M.F.; Mukherjee, A. Guar gum benzoate nanoparticle reinforced gelatin films for enhanced thermal insulation, mechanical and antimicrobial properties. Carbohydr. Polym. 2017, 170, 89–98. [Google Scholar] [CrossRef] [PubMed]
  246. Antoniou, J.; Liu, F.; Majeed, H.; Zhong, F. Characterization of tara gum edible films incorporated with bulk chitosan and chitosan nanoparticles: A comparative study. Food Hydrocoll. 2015, 44, 309–319. [Google Scholar] [CrossRef]
  247. Yang, W.; Owczarek, J.S.; Fortunati, E.; Kozanecki, M.; Mazzaglia, A.; Balestra, G.M.; Kenny, J.M.; Torre, L.; Puglia, D. Antioxidant and antibacterial lignin nanoparticles in polyvinyl alcohol/chitosan films for active packaging. Ind. Crops Prod. 2016, 94, 800–811. [Google Scholar] [CrossRef]
  248. Oun, A.A.; Rhim, J.-W. Effect of oxidized chitin nanocrystals isolated by ammonium persulfate method on the properties of carboxymethyl cellulose-based films. Carbohydr. Polym. 2017, 175, 712–720. [Google Scholar] [CrossRef]
  249. Oun, A.A.; Rhim, J.-W. Preparation of multifunctional chitin nanowhiskers/ZnO-Ag NPs and their effect on the properties of carboxymethyl cellulose-based nanocomposite film. Carbohydr. Polym. 2017, 169, 467–479. [Google Scholar] [CrossRef]
  250. Silva, N.H.C.S.; Vilela, C.; Almeida, A.; Marrucho, I.M.; Freire, C.S.R. Pullulan-based nanocomposite films for functional food packaging: Exploiting lysozyme nanofibers as antibacterial and antioxidant reinforcing additives. Food Hydrocoll. 2018, 77, 921–930. [Google Scholar] [CrossRef]
  251. Banerjee, S.; Bagchi, B.; Bhandary, S.; Kool, A.; Hoque, N.A.; Biswas, P.; Pal, K.; Thakur, P.; Das, K.; Karmakar, P.; et al. Antimicrobial and biocompatible fluorescent hydroxyapatite-chitosan nanocomposite films for biomedical applications. Colloids Surf. B Biointerfaces 2018, 171, 300–307. [Google Scholar] [CrossRef]
  252. Gopi, S.; Amalraj, A.; Jude, S.; Thomas, S.; Guo, Q. Bionanocomposite films based on potato, tapioca starch and chitosan reinforced with cellulose nanofiber isolated from turmeric spent. J. Taiwan Inst. Chem. Eng. 2019. [Google Scholar] [CrossRef]
  253. Rauscher, H.; Rasmussen, K.; Sokull-Klüttgen, B. Regulatory Aspects of Nanomaterials in the EU. Chemie Ingenieur Technik 2017, 89, 224–231. [Google Scholar] [CrossRef]
  254. European Union (Ed.) Regulation (EU) No 528/2012 of the European Parliament and of the Council of 22 May 2012 Concerning the Making Available on the Market and Use of Biocidal Products; Official Journal of European Union: Brussels, Belgium, 2012; p. L167/161. [Google Scholar]
  255. Luna-Martínez, J.F.; Hernández-Uresti, D.B.; Reyes-Melo, M.E.; Guerrero-Salazar, C.A.; González-González, V.A.; Sepúlveda-Guzmán, S. Synthesis and optical characterization of ZnS–sodium carboxymethyl cellulose nanocomposite films. Carbohydr. Polym. 2011, 84, 566–570. [Google Scholar] [CrossRef]
  256. El-Wakil, N.A.; Hassan, E.A.; Abou-Zeid, R.E.; Dufresne, A. Development of wheat gluten/nanocellulose/titanium dioxide nanocomposites for active food packaging. Carbohydr. Polym. 2015, 124, 337–346. [Google Scholar] [CrossRef]
  257. Gorrasi, G.; Vertuccio, L. Evaluation of zein/halloysite nano-containers as reservoirs of active molecules for packaging applications: Preparation and analysis of physical properties. J. Cereal Sci. 2016, 70, 66–71. [Google Scholar] [CrossRef]
  258. Motamedi, E.; Nasiri, J.; Malidarreh, T.R.; Kalantari, S.; Naghavi, M.R.; Safari, M. Performance of carnauba wax-nanoclay emulsion coatings on postharvest quality of ‘Valencia’ orange fruit. Sci. Hortic. 2018, 240, 170–178. [Google Scholar] [CrossRef]
  259. Bott, J.; Störmer, A.; Franz, R. Migration of nanoparticles from plastic packaging materials containing carbon black into foodstuffs. Food Addit. Contam. Part A Chem. Anal. Control Expo. Risk Assess. 2014, 31, 1769–1782. [Google Scholar] [CrossRef]
  260. Avella, M.; De Vlieger, J.J.; Errico, M.E.; Fischer, S.; Vacca, P.; Volpe, M.G. Biodegradable starch/clay nanocomposite films for food packaging applications. Food Chem. 2005, 93, 467–474. [Google Scholar] [CrossRef]
  261. Abreu, A.S.; Oliveira, M.; de Sá, A.; Rodrigues, R.M.; Cerqueira, M.A.; Vicente, A.A.; Machado, A.V. Antimicrobial nanostructured starch based films for packaging. Carbohydr. Polym. 2015, 129, 127–134. [Google Scholar] [CrossRef] [PubMed]
  262. Hu, J.; Kai, D.; Ye, H.; Tian, L.; Ding, X.; Ramakrishna, S.; Loh, X.J. Electrospinning of poly(glycerol sebacate)-based nanofibers for nerve tissue engineering. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 70, 1089–1094. [Google Scholar] [CrossRef]
  263. Naseri-Nosar, M.; Farzamfar, S.; Sahrapeyma, H.; Ghorbani, S.; Bastami, F.; Vaez, A.; Salehi, M. Cerium oxide nanoparticle-containing poly (ε-caprolactone)/gelatin electrospun film as a potential wound dressing material: In vitro and in vivo evaluation. Mater. Sci. Eng. C 2017, 81, 366–372. [Google Scholar] [CrossRef] [PubMed]
  264. Tra Thanh, N.; Ho Hieu, M.; Tran Minh Phuong, N.; Do Bui Thuan, T.; Nguyen Thi Thu, H.; Thai, V.P.; Do Minh, T.; Nguyen Dai, H.; Vo, V.T.; Nguyen Thi, H. Optimization and characterization of electrospun polycaprolactone coated with gelatin-silver nanoparticles for wound healing application. Mater. Sci. Eng. C 2018, 91, 318–329. [Google Scholar] [CrossRef]
  265. Ambrogi, V.; Pietrella, D.; Nocchetti, M.; Casagrande, S.; Moretti, V.; De Marco, S.; Ricci, M. Montmorillonite–chitosan–chlorhexidine composite films with antibiofilm activity and improved cytotoxicity for wound dressing. J. Colloid Interface Sci. 2017, 491, 265–272. [Google Scholar] [CrossRef] [PubMed]
  266. Cacciotti, I.; Ciocci, M.; Di Giovanni, E.; Nanni, F.; Melino, S. Hydrogen Sulfide-Releasing Fibrous Membranes: Potential Patches for Stimulating Human Stem Cells Proliferation and Viability under Oxidative Stress. Int. J. Mol. Sci. 2018, 19. [Google Scholar] [CrossRef]
  267. Wang, X.; Chang, J.; Wu, C. Bioactive inorganic/organic nanocomposites for wound healing. Appl. Mater. Today 2018, 11, 308–319. [Google Scholar] [CrossRef]
  268. Jacob, J.; Haponiuk, J.T.; Thomas, S.; Gopi, S. Biopolymer based nanomaterials in drug delivery systems: A review. Mater. Today Chem. 2018, 9, 43–55. [Google Scholar] [CrossRef]
  269. Javanbakht, S.; Namazi, H. Doxorubicin loaded carboxymethyl cellulose/graphene quantum dot nanocomposite hydrogel films as a potential anticancer drug delivery system. Mater. Sci. Eng. C 2018, 87, 50–59. [Google Scholar] [CrossRef] [PubMed]
  270. Benucci, I.; Liburdi, K.; Cacciotti, I.; Lombardelli, C.; Zappino, M.; Nanni, F.; Esti, M. Chitosan/clay nanocomposite films as supports for enzyme immobilization: An innovative green approach for winemaking applications. Food Hydrocoll. 2018, 74, 124–131. [Google Scholar] [CrossRef]
  271. Stumpf, T.R.; Yang, X.; Zhang, J.; Cao, X. In situ and ex situ modifications of bacterial cellulose for applications in tissue engineering. Mater. Sci. Eng. C 2018, 82, 372–383. [Google Scholar] [CrossRef]
  272. Hasan, A.; Waibhaw, G.; Saxena, V.; Pandey, L.M. Nano-biocomposite scaffolds of chitosan, carboxymethyl cellulose and silver nanoparticle modified cellulose nanowhiskers for bone tissue engineering applications. Int. J. Biol. Macromol. 2018, 111, 923–934. [Google Scholar] [CrossRef]
  273. Kalishwaralal, K.; Jeyabharathi, S.; Sundar, K.; Selvamani, S.; Prasanna, M.; Muthukumaran, A. A novel biocompatible chitosan–Selenium nanoparticles (SeNPs) film with electrical conductivity for cardiac tissue engineering application. Mater. Sci. Eng. C 2018, 92, 151–160. [Google Scholar] [CrossRef]
  274. Kalita, H.; Pal, P.; Dhara, S.; Pathak, A. Fabrication and characterization of polyvinyl alcohol/metal (Ca, Mg, Ti) doped zirconium phosphate nanocomposite films for scaffold-guided tissue engineering application. Mater. Sci. Eng. C 2017, 71, 363–371. [Google Scholar] [CrossRef]
  275. Cacciotti, I. Cationic and Anionic Substitutions in Hydroxyapatite. In Handbook of Bioceramics and Biocomposites; Antoniac, I.V., Ed.; Springer International Publishing: Cham, Switzerland, 2016; pp. 145–211. [Google Scholar]
  276. Nazeer, M.A.; Yilgör, E.; Yilgör, I. Intercalated chitosan/hydroxyapatite nanocomposites: Promising materials for bone tissue engineering applications. Carbohydr. Polym. 2017, 175, 38–46. [Google Scholar] [CrossRef]
  277. Arumugam, R.; Subramanyam, V.; Chinnadurai, R.K.; Srinadhu, E.S.; Subramanian, B.; Nallani, S. Development of novel mechanically stable porous nanocomposite (PVDF-PMMA/HAp/TiO2) film scaffold with nanowhiskers surface morphology for bone repair applications. Mater. Lett. 2019, 236, 694–696. [Google Scholar] [CrossRef]
  278. Bianco, A.; Bozzo, B.M.; Del Gaudio, C.; Cacciotti, I.; Armentano, I.; Dottori, M.; D’Angelo, F.; Martino, S.; Orlacchio, A.; Kenny, J.M. Poly (L-lactic acid)/calcium-deficient nanohydroxyapatite electrospun mats for bone marrow stem cell cultures. J. Bioact. Compat. Polym. 2011, 26, 225–241. [Google Scholar] [CrossRef]
  279. D’Angelo, F.; Armentano, I.; Cacciotti, I.; Tiribuzi, R.; Quattrocelli, M.; Del Gaudio, C.; Fortunati, E.; Saino, E.; Caraffa, A.; Cerulli, G.G.; et al. Tuning Multi/Pluri-Potent Stem Cell Fate by Electrospun Poly(l-lactic acid)-Calcium-Deficient Hydroxyapatite Nanocomposite Mats. Biomacromolecules 2012, 13, 1350–1360. [Google Scholar] [CrossRef]
  280. Ji, M.; Li, H.; Guo, H.; Xie, A.; Wang, S.; Huang, F.; Li, S.; Shen, Y.; He, J. A novel porous aspirin-loaded (GO/CTS-HA)n nanocomposite films: Synthesis and multifunction for bone tissue engineering. Carbohydr. Polym. 2016, 153, 124–132. [Google Scholar] [CrossRef]
  281. You, Y.; Zhang, H.; Liu, Y.; Lei, B. Transparent sunlight conversion film based on carboxymethyl cellulose and carbon dots. Carbohydr. Polym. 2016, 151, 245–250. [Google Scholar] [CrossRef]
  282. Mitta, S.B.; Reddeppa, M.; Vellampatti, S.; Dugasani, S.R.; Yoo, S.; Lee, S.; Kim, M.-D.; Ha Park, S. Gold nanoparticle-embedded DNA thin films for ultraviolet photodetectors. Sens. Actuators B Chem. 2018, 275, 137–144. [Google Scholar] [CrossRef]
Figure 1. Schematic preparation of nanocomposite films and their functional properties.
Figure 1. Schematic preparation of nanocomposite films and their functional properties.
Polymers 11 00675 g001
Figure 2. Potential mechanisms of antimicrobial activities of nanoparticles.
Figure 2. Potential mechanisms of antimicrobial activities of nanoparticles.
Polymers 11 00675 g002
Table 1. Advantages and disadvantages of biopolymer films.
Table 1. Advantages and disadvantages of biopolymer films.
Type of Biopolymer FilmAdvantagesDisadvantages
cellulose-based filmstasteless, odorless, resistant to oil and fat, hydrophilic nature [27]; thermal and chemical stability [39]hardly dissolves or melts due to high crystallinity [40]; non antimicrobial activity [41]
chitin and chitosan-based filmsgood CO2 barrier properties, antimicrobial activity [42]non antioxidant and antifungal activity [43]; limited oxygen and water impediment ability [44]
starch-based filmsodorless, tasteless, good O2 and CO2 barrier properties [45]poor water vapor barrier [46] and tensile properties [47]
pectin-based filmsexcellent oxygen barring capacity [48]high water vapor permeability [49]; poor mechanical performance [48]
pullulan-based filmsheat-sealable [50]; highly impermeable to both oil and oxygen [51]; excellent mechanical properties and a low permeability to oil and oxygen [52]low solubility [50]; hydrophilic nature [52]
alginate-based filmsgood water solubility, gel ability, and film-forming properties [53]insufficient mechanical properties and poor water resistance [54]
gelatin-based filmsgood mechanical and barrier properties [55]low water vapor permeability [56]
whey protein-based filmsexcellent barrier properties to aroma compounds and oils [27]hydrophilic nature so it has limitation to moisture [27]
lipids-based filmsexcellent barriers against moisture migration [27]damage the appearance and gloss of the coated food products [57]
bacterial cellulose-based filmsflexibility and excellent mechanical properties [58]insoluble in water [59]
PCL- based filmshigh mechanical strength, biocompatibility, processability, and permeability [60]highly hydrophobic and crystalline [61]
PLA-based filmsenvironmental friendliness, good transparency, and biological compatibility [62]high hardness, and brittleness, low strength, and poor thermal stability [63]
PGA-based filmshigh mechanical strength [64]high degree of crystallinity, a high melting point, and it is insoluble in common organic solvents [65]
PU-based filmsfavorable processability, versatile structure–property relationships, and excellent elasticity [66]low water resistance and hardness [66]
Table 2. Recent examples of nanofillers properties.
Table 2. Recent examples of nanofillers properties.
Type of NanofillersProperties Added to the FilmReference
Clay Nanofillers
MMT, Hal etc.UV shielding properties[69]
Good mechanical stability
Thermal stability[70]
Organic Nanofillers
NanocelluloseBlood compatibility
Antibacterial effect[71]
Thermal stability[72]
Good mechanical stability
Low cytotoxicity[73]
Chitosan nanoparticlesBiocompatibility
Biodegradability[74]
Low toxicity[75]
Antimicrobial activity[76]
Inorganic Nanofillers
AgNPsAntimicrobial effect[77]
UV shielding properties[78]
Antioxidant activity[79]
Photocatalytic effect[80]
SeNPsAntimicrobial effect[81]
Antioxidant activity[82]
CuNPsAntimicrobial effect[83]
UV shielding properties[84]
SNPsAntimicrobial effect[85]
TiO2 NPsAntifouling effect[86]
Antimicrobial activity[87]
Photocatalytic activity[88]
UV shielding properties[69]
ZnO NPsAntifungal effect
UV shielding properties[89]
Antimicrobial effect[90]
Dielectric properties
Electromagnetic shielding
Thermal conductivity[91]
CeO2Antimicrobial effect
UV shielding properties
Flame retardancy
Wrinkle resistance[92]
Carbon Nanofillers
Graphene, graphene oxide, etc.Lightweight
Processing benefits, flexibility, resistance to corrosion
Extraordinary electrical, mechanical, and thermal properties[93]
Table 3. Recent examples of nanocomposite films.
Table 3. Recent examples of nanocomposite films.
NanofillerPolymerThe Effect of Nanofiller AdditionReference
Metal Nanostructures
AgNPsgelatin• Improvement of antibacterial effect from 0 up to 14 mm of inhibition zone against S. typhimurium, B. cereus, L. monocytogenes, E. coli, and S.aureus
• Reduction of TS and YM by up to ~25 % and ~36%, respectively, depending on AgNPs concentration
• No changes in EAB, WVP, MC, and WCA
[199]
AgNPschitosan-gelatin• Reduction of TS (~27%) and improvement of EAB (~34%)
• Increased the shelf-life of red grapes on which the film was applied
[200]
AgNPschitosan• Improvement of antimicrobial activity against P. aeruginosa (to ~28 mm of inhibition zone), S. aureus (to ~37 mm), and MRSA (to 24.73 mm), depending on AgNPs concentration[201]
AgNPschitosan/cellulose• Improvement of antimicrobial activities against S. aureus (~0.8 mm of inhibition zone) and E. coli (~1.2 mm)[202]
AgNPschitosan/PVA• Improvement of antioxidant activity by up to ~33% (DPPH radical scavenging activity), up to ~37% (ferric reducing ability) and up to ~31% (β-Carotene bleaching), depending on AgNPs concentration
• Low toxicity
• Improvement of antimicrobial activity against S. areus ( increase by ~1 mm of inhibition zone), B. cereus (~8 mm), M. luteus (~2 mm), S.enterica (~3 mm), E. coli (~3 mm), and P. aeruginosa (~2 mm)
[203]
AgNPs
AuNPs
chitosan• Improvement of antimicrobial activity
• AuNPs has better activity against A. niger than AgNPs (from 0 to 25 mm of inhibition zone)
• AgNPs has better activity against C. albicans than AuNPs ( from 6 to 19 mm of inhibition zone)
• No significant difference between antimicrobial activity of AuNPs and AgNPs against S. aureus and P. aeruginosa
[204]
AgNPs
SeNPs
furcellaran• Enhancement of MC (with AgNPs ~11.5% and with SeNPs ~14%), WS, EM (with AgNPs and with SeNPs ~10%), but reduction in SR (with AgNPs ~13% and with SeNPs ~20%)
• AgNPs improved the UV-blocking effect
• No changes in EAB
• SeNPs improved antimicrobial activity against E. coli (SeNPs from 0 up to ~38 mm of inhibition zone; AgNPs from 0 to ~10 mm), S. aureus (SeNPs from 0 up to ~22 mm), and MRSA (SeNPs from 0 up to ~26 mm)
[81]
SNPschitosan• Increment of TS (by up to ~18%), EM (by up to ~18%) and WCA (by up to ~6%)
• Reduction of EAB (by up to ~39%), WVP (by up to 14%)
• Enhancement of thermal stability
• Antimicrobial activity against L. monocytogenes (complete destroy after 12 h) and E. coli (complete destroy after 6 h)
[85]
Lignin capped AgNPsagar• Enhancement of UV screening effect
• Improvement of TS (by up to ~23%) and YM (by up to ~13 %)
• No changes in EAB
• Reduction of WVP (by up to ~22%), WCA (by up to ~9%), WS (by up to ~16%), and SR (by up to ~50%)
• Improvement of antimicrobial activity against E. coli (complete destroy after 6 h) and L. monocytogenes (complete destroy after 12 h)
[150]
Ag-Cu NPsgelatin• Strong UV screening effect
• Improvement in TS (by up to ~49%) but reduction in EAB (by up to ~40%)
• Strong antimicrobial activity against S. thyphimurium (from 3.5 to 7 log) and L. monocytogenes (from 0.5 to 3 log)
• Enhancement of thermal stability
[173]
AgNPs inside gelatin-montmorillonite (AgM)cellulose acetate• Improvement of UV barrier, TS (by up to ~6%) and EM (by up to ~18%), but reduction in EAB (by up to~50%)
• Reduction of OP (by up to ~14%)
• Enhancement of thermal stability
• Improvement of antimicrobial activity against E. coli (from 0 to 36 mm of inhibition zone), S. aureus (from 0 to 34 mm), Salmonella (from 0 to 32 mm), Psuedomonas (from 0 to 35 mm), A. niger, and A. flavus, depending on the AgNPs and thymol concentration
[205]
AgNPs
chitin nanofiber
chitosan• AgNPs reduced TS, YM, and color properties
• CHNF improved WS, SR, WVP, TS, and YM
[144]
AgNPs
nanocellulose
PVA• AgNPs enhanced TS and EAB and reduced WVP
• Enhancement of thermal stability
• Improvement of antimicrobial activity against E.coli and MRSA
[206]
Metal Oxide Nanostructures
Magnetite nanoparticlesgelatin• Enhancement of thermal stability[207]
Magnetic nanoparticleschitosan• Improvement of TS (by up to ~37%) and EAB (by up to ~18%)[208]
TiO2 NPsgelatin• Enhancement of TS (by up to ~60%) and EAB (by up to ~48%)
• Improvement of barrier properties against UVC light
• Irradiation of the film with UV-A light (365 nm) resulted in the most effective antibacterial activity against E. coli
[209]
TiO2 NPsgelatin–agar• Improvement of TS (by up to ~29%) and reduction in EAB (by up to ~22%)
• Increment of WS (by up to ~4%) and MC (by up to ~10%)
• Reduction of WVP (by up to ~32%)
• UV blocking effect
[146]
TiO2 NPschitosan• Improvement of antimicrobial activity against bacteria (S. aureus, E. coli, S. typhimurium, and P. aeruginosa) and fungi (Aspergillus spp. and Penicillium spp.)
• Enhancement of TS (by up to ~56%), WVP (by up to ~22%) and ethylene photocatalytic degradation properties
• Reduction of EAB (by up to ~10%)
[210]
TiO2 NPspotato starch• Reduction of WS (by up to ~9%), MC (by up to ~11%) and WVP (by up to ~35%)
• Improvement of TS (by up to ~45%) and reduction of EAB (by up to ~28%)
• Improvement of UV-blocking effect
[121]
CuO NPsPVA–gelatin• Improvement of UV screening effect[123]
ZnO NPsgelatin• Improvement of UV screening effect and thermal stability
• Reduction of EM (by up to ~82%) and TS (by up to ~72%)
• Increment of WVP (by up to ~99%), MC (by up to ~29%), WCA (by up to ~20%)
• Improvement of antimicrobial activity against E. coli (from 9 to 5 log) and L. monocytogenes (from 9 to 1 log)
[160]
ZnO NPschitosan/CMC• Improvement of shelf life of white soft cheese on which the film was applied
• Improvement of antibacterial activity against bacteria S. aureus (from 5 to 11 mm of inhibition zone) P. aeruginosa (from 3 to 11 mm), E. coli (from 3 to 9 mm), and fungi C. albicans (from 3 to 15 mm)
• Improvement of TS (by up to ~85%)
[211]
ZnO NPschitosan• Improvement of TS but reduction of EAB
• Improvement of antimicrobial activity against E. coli (3.4 log reduction after 0.5 h) and S. aureus (4 log reduction after 0.5 h)
• Biocompatibility and nontoxicity
[212]
ZnO NPsmahua oil-based polyurethane/chitosan• Improvement of TS (by up to ~56%) but reduction of EAB (by up to ~20%)
• Reduction of OP (by up to ~3%) and WVP (by up to ~37%)
• Improvement of antimicrobial activity against E. coli (~25 mm) and S. aureus (~20 mm)
• UV-screening ability and biodegradation
• Non-cytotoxic
[213]
ZnO NPsPLA• Improvement of TS (by up to ~37%), WVP (by up to ~31%) and UV-light barrier properties
• Reduction of EAB (by up to ~10%)
• Improvement of antibacterial activity against E. coli (from 10 to 3.5 log after 12 h) and L. monocytogenes (from 12 to 8 log after 12 h)
[31]
ZnO NPschitosan/PVA• Improvement in photoluminescent properties and thermal stability[214]
SnO2 NPsCMC• The choice of nanocomposite preparation procedure caused four different morphologies of SnO2 NPs (microcube, nanosphere, olive-like and nano-flower) which had different effects on thermal stability of CMC[125]
ZnO NPs
CuO NPs
carrageenan• ZnONPs strongly improved antimicrobial activity against E. coli and L. monocytogenes, UV-blocking effect and thermal stability
• Reduction of TS (by up to ~55%) and EM (by up to ~26%) depending on the CuO NPs and ZnO NPs concentration and the ratio of nanoparticles
• Improvement of EAB
[215]
Fe2O3 NPscellulose• Improvement of TS (by up to ~10%) and YM (by up to ~15%) and thermal stability[216]
MgO NPsPLA/polyethylene glycol• Improvement of EAB (by up to ~86%) but reduction of TS (by up to ~64%)
• Enhancement of optical properties
[217]
α- Fe2O3 NPs
FeNPs
chitosan/PVA• Improvement of magnetic properties
• FeNPs decreased TS and EM depending on the composition of chitosan and PVA
• Fe2O3 NPs increased TS and EM depending on the composition of chitosan and PVA
• FeNPs and Fe2O3 NPs caused reduction in EAB
[33]
ZnO nanorod
nano-kaolin
semolina• Reduction of OP (by up to ~34%), MC (by up to ~64%) and WS (by up to ~56%), depending on the ratio ZnO nanorods/nano-koalin
• Improvement of UV barrier properties and antimicrobial activity against E. coli (from 0 to ~3 mm)
[218]
ZnO nanorodsstarch/gelatin• Reduction of OP (by up to ~61%)
• Improvement of TS (by up to ~30%) but decrement of EAB (by up to ~44%)
• Enhancement of UV barrier properties
[174]
ZnO nanorodsgelatin/clove essential oil• Reduction of TS (by up to ~61%) and increment of EAB (by up to ~155%) and OP (by up to 98%)
• Increment of oxygen and UV barrier property
• Improvement of antimicrobial activity against L. monocytogenes (from 10 to 0 log after 7 days) and S. typhimurium (from 10 to 0 log after 7 days)
• Improvement of shelf-life of peeled shrimps
[219]
ZnO nanorodsgelatin• Decrement of hydrophobicity and moisture contents
• Reduction of WVP (by up to ~80%)
• Improvement of UV-blocking effect and antimicrobial activity against S. aureus (from 0 to 80 mm2 of inhibition zone)
[185]
ZnO nanorodssoybean polysaccharide• Reduction of WVP (by up to ~36%) and OP (by up to ~43%)
• Decrement of TS (by up to ~18%) and increment of EAB (by up to ~41%) and heat seal strength (by up to ~29%)
• Improvement of antimicrobial activity against E. coli (from 7 to 5 log after 12 h) and S. aureus (from 6 to 1 log after 12 h) and UV-blocking effect
[220]
ZnO nanorodsPVA/CMC• Improvement of dielectric properties[221]
Cellulose Nanostructures
cellulose nanocrystalscarrageenan• Improvement of TS (by up to ~70%) and toughness (by up to ~10%) parameters
• Reduction of EAB (by up to ~25%)
• Enhancement of thermal stability
[179]
rice cellulose nanocrystalschitosan/PVA• Improvement of TS (by up to ~75%) and EM (by up to ~98%)
• Reduction of EAB (by up to ~43%)
• Enhancement of thermal stability
• No changes in antifungal against C. gloeosporioides and L. theobromae and antimicrobial against S. mutans, S. aureus, E. coli, and P. aeruginosa activities
[222]
cellulose nanocrystalschitosan• Improvement of mechanical properties (by up to ~44%) and thermal stability[106]
flax cellulose nanocrystalschitosan• Improvement of TS (by up to ~24%), EAB (by up to ~22%) and YM (by up to ~140%)
• Reduction of WVTR (by up to ~11%) and increment of WVP (by up to ~85%)
• Enhancement of antimicrobial activity against P. aeruginosa, E. faecalis, L. monocytogenes, E. coli, and S. aureus (from 6.31 to 16.05 mm of inhibition zone)
[223]
bacterial cellulose nanocrystalPVA• Improvement of TS, YM, and toughness depending on the presence of glycerol, boric acid, and BCNC
• No changes in EAB
[105]
cellulose nanowhiskerschitosan• Increment of YM but reduction of TS and EAB
• Enhancement of thermal stability
[224]
nanocrystalline cellulosechitosan/guar gum• Improvement of the shear viscosity of the suspensions
• Reduction of air permeability (by up to ~53%) and EAB (by up to ~53%)
• Improvement of YM (by up to ~86%) and TS (by up to ~75%)
[225]
sugar palm nanocrystalline cellulosesugar palm fibre• Reduction of MC (by up to ~19%), WS (by up to ~56%)
• Improvement of WVP (by up to ~18%) and thermal stability
[153]
bacterial cellulose nanocrystals and AgNPschitosan• Improvement of UV barrier properties
• Enhancement of physical, mechanical properties and in thermal stability, depending on the concentration and ratio of AgNPs and BCNC
• Improvement of antimicrobial and antifungal activity (from 0 to 96 mm2 of inhibition zone depending on the concentration and ratio of AgNPs and BCNC)
[184]
cellulose nanocrystalschitosan
alginate
κ-carrageenan
• Improvement of YM, TS, and toughness in every type of film
• Reduction in parameter of EAB
[178]
cotton linter cellulose nanofibrilCMC• Improvement of TS (by up to ~23%) and EM (by up to ~28%) but reduction of EAB (by up to ~26%)
• Enhancement of thermal stability
• Reduction of WCA (by up to ~39%)
• No changes in WVP
[226]
celullose nanocrystalsCMC• Enhancement of thermal stability
• Improvement of TS (by up to ~74%) and EM (by up to ~129%) and reduction of EAB (by up to ~47%)
[227]
cellulose nanofiberssoy protein• Improvement of TS (by up to ~400%) and YM (by up to ~767%)
• Reduction of EAB (by up to ~56%)
• No effects of WVP
[228]
cellulose nanocrystalscassava starch• Reduction of WVP (by up to ~43%), oil permeability (by up to ~42%) and MC
• Improvement of TS (5.6 times higher than cassava starch films)
• Increment of WS
[229]
cellulose nanocrystalsPVA/CMC• Enhancement of TS (by up to ~83%) and EM (by up to ~147%)
• Reduction of WVP (by up to ~82%)
[2]
crystalline nanocelluloseCMC/chitosan• Enhancement of barrier against grease and oil
• Improvement of TS (52% higher than CMC/chitosan films), YM, and WVP (by up to ~38%)
• Reduction of strain at break
[230]
Cellulose nanofibers
TiO2 NPs
whey protein• TiO2 increased the water resistance
• TiO2 and CNFs improve mechanical properties and the values depended on the concentration of nanofillers
• TiO2 enhances antimicrobial activity against L. monocytogenes and S. aureus and antioxidant properties
[87]
licorice residue nanocellulosesoy protein isolate• Improvement of TS but reduction of EAB
• Reduction in WVP (by up to ~27%) and OP (by up to ~55%)
• UV-blocking effect
[152]
Nanoclays
betonine nanoclayschitosan/PVA• Reduction of WVP (by up to ~69%) and TS (by up to ~30%)
• Improvement of thermal stability
• Enhancement of antibacterial properties against E. coli (efficiency 48.50 %), P. aeruginosa (efficiency 40%), S. aureus (efficiency 8%)
[231]
cloisite Na+ nanoclaysagar• Improvement of TS (by up to ~31%)
• Reduction in WVP (by up to ~50%), WCA (by up to ~10%), and WS (by up to ~23%)
[155]
halloysite nanotubeschitosan/starch• Reduction of WS
• Improvement of water absorption capacity, porosity, folding strength, and WVTR
• Improvement of impermeability to bacteria (the microbial penetration test)
[145]
hallosite nanotubes with metal ions (Ag, Zn, Cu)CMC• Improvement of WVP and thermal stability
• Enhancement of antimicrobial activity against L. monocytogenes and E. coli
• Improvement in TS and EM but reduction of EAB (all results were depending on the type of metal ions in hallosite nanutubes)
[232]
nanoclays:
Na+ montmorillonite
halloysite
Nanomer® I.44 P
fenugreek seed gum• Enhancement of thermal stability
• No influence on antimicrobial activity against E. coli, S. aureus, B. cereus
• Antimicrobial activity against L. monocytogenes (all results were depending on the type of nanoclays)
[168]
halloysite nanotubes loaded with the essential oilpectin• Reduction in thermal stability and EAB
• Improvement in TS and EM and surface hydrophobicity
[233]
halloysite nanotubes
ZnO NPs
alginate• Improvement of TS (by up to ~12%), WVP (by up to ~27%), and WCA (by up to ~28%)
• No changes in EAB (by up to ~6%) and EM (by up to ~7%)
• Improvement in UV-blocking effect and antimicrobial activity against E. coli (from 7 to 0 log after 3 h) and L. monocytogenes (from 6 to 0 after 9 h)
[95]
Cloisite 30B
AgNPs
gelatin• Improvement of optical properties
• AgNPs improved TS while nanoclay increased EAB
• Enhancement of antimicrobial activity against E. coli and L. monocytogenes
[234]
bismuth tungstate/ TiO2 NPs (Bi2WO6-TiO2)starch• Improvement of TS (by up to ~233%) and photocatalytic activity
• Reduction of EAB (by up to ~15%)
[235]
halloysite
nisin
starch• HTN improved TS and YM
• Nisin reduced TS and YM but increased antimicrobial properties against L. monocytogenes, C. perfringens, and S. aureus
• No changes in WS
[236]
calcium montmorillonitecarboxymethyl starch• Improvement of TS (by up to ~500%), YM (by up to ~1733%), and WCA (by up to ~53%)
• Reduction of EAB (by up to ~49%), WS (by up to ~4%)
[237]
montmorillonite
cellulose NPs
alginate• Reduction of WVP
• CNC increased the TS and EAB, while MMT in high concentration reduced the TS
[238]
montmorillonite − CuO nanocompositeschitosan• Enhancement of TS, EAB, WVP, and OP
• Reduction of WS and UV transition
• Improvement of antimicrobial activity against E. coli, P. aeruginosa, S. aureus, B. cereus (all results were depending on the concentration and ratio of MMT and CuONPs
[169]
montmorillonite ZnO nanopowderscationic starch• MMT reduced of WVP and UV light transmittance
• ZnO improved of WVP and UV light transmittance
• Improvement of TS and optical properties, but reduction of EAB
[166]
sodium montmorillonite nanoclay
ZnO
CMC• Reduction in WVP
• Addition of ZnO NPs increase UV-blocking effect
• Addition of ZnO NPs enhance antimicrobial activity against E. coli and S. aureus
[154]
Carbon Nanostructures
multi-walled carbon nanotube-Valinechitosan/PVA• Improvement of thermal stability[239]
carbon nanotubes
AgNPs
chitosan• Enhancement of TS (by up to ~131%), EAB (by up to ~18%) and toughness (by up to 125%)
• Reduction of dielectric properties
[240]
graphene oxidecellulose carbamate• Improvement of thermal stability
• Improvement of TS (by up to ~100%) and reduction of EAB (by up to ~64%)
[129]
graphene nanoplateletsCMC• Reduction of ultimate tensile strength (by up to ~50%) but increment of strain to break (by up to ~66%)
• Enhancement of UV-blocking effect
• Improvement of water repelling nature
[241]
reduced graphene oxidesodium CMC• Improvement of TS (by up to ~73%) and YM (by up to ~132%)[177]
graphene oxideamylose• Enhancement of stability in acidic and alkaline solutions
• Improvement of TS (16.5 times higher than amylose films)
• Reduction of the MC and UV transmittance
[242]
graphene oxide
reduced graphene oxide
sodium CMC/silk fibroin• Improvement of thermal stability[131]
Other Nanostructures
melanin nanoparticlescarrageenan• Increment of TS (by up to ~27%), EAB (by up to ~25%), WVP (by up to ~25%) and WCA (by up to ~25%)
• Reduction of YM (by up to ~38%)
• Enhancement of antioxidant activity (by up to ~962%-DPPH method and by up to ~559%-ABTS method) and thermal stability
• Improvement of antimicrobial activity against L. monocytogenes and E. coli
[243]
ZnS NPschitosan/PVA• Reduction of WS and SR
• Improvement of TS and thermal stability and reduction of EAB (all results were correlating with the type of plasticizer)
[244]
guar gum benzoate NPsgelatin• Improvement of antimicrobial activity against E. coli and S. aureus
• Enhancement of thermal stability
• Improvement of TS (by up to 67%) and YM (by up to ~550%)
• Reduction in EAB (by up to ~63%)
[245]
chitosan NPsrice straw nanofibrillated cellulose• Improvement of TS (by up to ~40%) and YM (by up to ~42%) but reduction of EAB (by up to ~94%)
• Enhancement of antimicrobial activity against bacteria (S. aureus, E. coli), and yeast (S. cervisiae)
• Reduction in porosity
• No changes in WVP
[110]
chitosan NPstara gum• Improvement of TS and reduction of EAB
• Reduction of WS (by up to ~74%), WVP (by up to ~23%) and MC (by up to ~24%)
• Antimicrobial activity against E. coli (from 0 to 87.32 mm2 of inhibiton zone) and S. aureus (from 0 to 111.71 mm2)
[246]
chitosan NPsPVA/mulberry extract• Improvement of TS but reduction of EAB[111]
lignin NPschitosan
PVA
chitosan/PVA
• Improvement of TS and YM of PVA films
• Enhancement of thermal stability of chitosan, PVA, and chitosan/PVA films
• Improvement of UV barrier properties of tested films
• Enhancement of antioxidant activity of chitosan films with LNP
• Improvement of antimicrobial activity against Erwinia carotovora subsp. carotovora and Xanthomonas arboricola pv. Pruni
[247]
chitosan/gallic acid NPskonjac glucomannan• Improvement of UV barrier properties
• Enhancement of TS (by up to ~43%) but reduction of EAB (by up to ~16%) and WVP (by up to ~33%)
• Improvement of antimicrobial activity against S. aureus (from 0 to 20 mm of inhibition zone) and E. coli (from 0 to 12 mm)
[167]
chitin nanofibergelatin/CMC• Reduction of WS, SR and WVP
• Improvement of TS
[109]
chitin nanowhiskersmaize starch• Improvement of TS (by up to ~125%) and thermal stability but reduction of EAB (by up to ~37%) and WVP (by up to ~58%)
• Enhancement of antimicrobial activity against E. coli and L. monocytogenes
[143]
oxidized chitin nanocrystalsCMC• Improvement of TS (by up to ~88%) and EM (by up to ~244%)
• Reduction of WVP (by up to ~10%), WCA (by up to ~14%) and EAB (by up to ~65%)
[248]
chitin nanowhiskers /hybrid ZnO-Ag NPsCMC• Enhancement of thermal stability and UV-barrier property
• Improvement of TS (by up to ~32%) and YM (by up to ~101%) but reduction of EAB (by up to ~34%)
• Enhancement of antimicrobial activity against E.coli (from 6 to 0 log after 6h) and L, monocytogenes (from 7 to 4 log after 9 h)
• Reduction of WVP (by up to ~23%)
[249]
pullulanlysozyme nanofibers• Improvement of YM (by up to ~48%), TS (by up to ~7%) but reduction of EAB (by up to ~80%)
• Enhancement of antioxidant activity (from 0 to ~80% DPPH method) and antimicrobial activity against S. aureus
[250]
chitosannanocrystalline erbium doped hydroxyapatite• Improvement in antimicrobial activity against E. coli and S. aureus[251]
potato starch
tapioca starch
chitosan
turmeric nanofiber• Improvement of TS, YM, and thermal stability
• Reduction of EAB
• Antimicrobial activity against B. cereus, E. coli, S. aureus, and S. typhimurium (the values were depending in the type of biopolymer)
[252]
maltodextrinpolyvinyl acetate NPs• Improvement of TS (by up to ~106%)[165]
Abbreviations: NPs—nanoparticles; CMC—carboxymethyl cellulose; PVA—poly(vinyl) alcohol; PLA—poly(lactic acid); TS—tensile strength; EAB—elongation at break; YM—Young’s modulus; EM—elastic modulus; WVP—water vapor permeability; OP—oxygen permeability; WS—water solubility; SR—swelling ratio; MC—moisture content; WCA—water contact angle.

Share and Cite

MDPI and ACS Style

Jamróz, E.; Kulawik, P.; Kopel, P. The Effect of Nanofillers on the Functional Properties of Biopolymer-Based Films: A Review. Polymers 2019, 11, 675. https://doi.org/10.3390/polym11040675

AMA Style

Jamróz E, Kulawik P, Kopel P. The Effect of Nanofillers on the Functional Properties of Biopolymer-Based Films: A Review. Polymers. 2019; 11(4):675. https://doi.org/10.3390/polym11040675

Chicago/Turabian Style

Jamróz, Ewelina, Piotr Kulawik, and Pavel Kopel. 2019. "The Effect of Nanofillers on the Functional Properties of Biopolymer-Based Films: A Review" Polymers 11, no. 4: 675. https://doi.org/10.3390/polym11040675

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