Next Article in Journal
Non-Absorbing Dielectric Materials for Surface-Enhanced Spectroscopies and Chiral Sensing in the UV
Next Article in Special Issue
A Multifactorial Approach to Untangle Graphene Oxide (GO) Nanosheets Effects on Plants: Plant Growth-Promoting Bacteria Inoculation, Bacterial Survival, and Drought
Previous Article in Journal
A Review of Microscale, Rheological, Mechanical, Thermoelectrical and Piezoresistive Properties of Graphene Based Cement Composite
Previous Article in Special Issue
Preparation and Performance of Porous Carbon Nanocomposite from Renewable Phenolic Resin and Halloysite Nanotube
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphene Derivatives in Biopolymer-Based Composites for Food Packaging Applications

by
Ana Barra
1,2,3,†,
Jéssica D. C. Santos
1,4,†,
Mariana R. F. Silva
1,†,
Cláudia Nunes
2,
Eduardo Ruiz-Hitzky
3,
Idalina Gonçalves
2,
Selçuk Yildirim
4,
Paula Ferreira
1,* and
Paula A. A. P. Marques
5,*
1
Department of Materials and Ceramic Engineering, CICECO—Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
2
Department of Chemistry, CICECO—Aveiro Institute of Materials, University of Aveiro, 3810-193 Aveiro, Portugal
3
Materials Science Institute of Madrid, CSIC, c/Sor Juana Inés de la Cruz 3, 28049 Madrid, Spain
4
Institute of Food and Beverage Innovation, Zurich University of Applied Sciences, 8820 Wädenswil, Switzerland
5
Department of Mechanical Engineering, TEMA—Centre for Mechanical Technology and Automation, University of Aveiro, 3810-193 Aveiro, Portugal
*
Authors to whom correspondence should be addressed.
These authors contributed equally to the work.
Nanomaterials 2020, 10(10), 2077; https://doi.org/10.3390/nano10102077
Submission received: 23 September 2020 / Revised: 12 October 2020 / Accepted: 16 October 2020 / Published: 21 October 2020
(This article belongs to the Special Issue Investigation and Development of Graphene Oxide-Based Materials)

Abstract

:
This review aims to showcase the current use of graphene derivatives, graphene-based nanomaterials in particular, in biopolymer-based composites for food packaging applications. A brief introduction regarding the valuable attributes of available and emergent bioplastic materials is made so that their contributions to the packaging field can be understood. Furthermore, their drawbacks are also disclosed to highlight the benefits that graphene derivatives can bring to bio-based formulations, from physicochemical to mechanical, barrier, and functional properties as antioxidant activity or electrical conductivity. The reported improvements in biopolymer-based composites carried out by graphene derivatives in the last three years are discussed, pointing to their potential for innovative food packaging applications such as electrically conductive food packaging.

Graphical Abstract

1. Introduction

Non-sustainable food production and consumption leads to many serious health and environmental problems. It is estimated that one third of the total foodstuffs produced worldwide, 1.3 billion tons, equivalent to one trillion dollars, is wasted. The United Nations (UN) 2030 sustainable development agenda addresses this problem with Goal 12—Ensure sustainable consumption and production [1]. To mitigate this global problem, the development of food processing technologies and packaging materials is of the utmost importance. Packaging offers mechanical and barrier protection against light, dust, pests, gases, moisture, volatiles, and both chemical and microbiological contamination, extending the foodstuffs’ shelf lives while ensuring their quality, safety, and authenticity without requiring additives. However, the sustainability of conventional packaging has become a major issue. The use and disposal of non-biodegradable polymers led to environmental issues across numerous ecosystems [2]. In this context, biopolymers with biodegradable properties have been explored as an alternative to design novel packaging materials that can extend the foodstuffs’ shelf lives and, concomitantly, contribute to minimize the packaging negative environmental impact [3]. Although packaging is a top priority in research, with the development of bio-based and biodegradable plastics representing 53% share of this global demand, their introduction in the food packaging industry is well below expectations [4]. See, for instance, that sustainable packaging market alone is expected to grow from ca. US$ 225 billion in 2018 to over US$ 310 billion by 2024 [5]. The high cost and difficulty to overcome specific technical issues (e.g., viscosity, thermal stability, hydrophilicity, low mechanical strength, and poor gas barrier properties), essential to preserve food quality and safety, are the main obstacles for their large-scale use [6]. Investments in research have been enforced to develop biodegradable food packaging materials, fitting practical criteria. Recently, the incorporation of graphene derivatives into biopolymer has garnered attention due to their potential to provide enhanced mechanical and barrier properties [7,8,9]. This review examines the advances made in the last three years in the field of biopolymer-based biocomposites containing graphene derivatives. Additionally, the most used biopolymers for packaging materials were reviewed. Nowadays, the current design of biocomposites containing graphene derivatives goes far beyond the reinforcement of barrier and mechanical properties. Some graphene derivatives, such as reduced graphene oxide (rGO) with radical scavenging capacity, may assign an antioxidant activity to the packaging materials, an extremely important property to enhance foodstuffs’ shelf lives [10]. An interesting breakthrough introduced through use of graphene derivatives is electrical conductivity, which is advantageous for the development of food packaging materials suitable for their application in pulsed electric field technology (PEF) [10,11,12].

2. Biopolymers as Food Packaging Raw Materials

Biodegradable biopolymers have been gaining societal, scientific, industrial, and economic importance to develop sustainable food packaging bioplastics, alternatives to conventional non-biodegradable fossil-fuel-based plastics [6,13]. In recent years, several approaches to develop biodegradable plastic materials from biomass feedstock by biocatalytic transformation, chemical synthesis, or simple feedstock polymers’ extraction have received much attention, taking advantage of using renewable resources as raw materials during polymer production/purification. This strategy contributes to a sustainable essence of the final packaging while advancing toward a circular economy [6]. According to nova-Institute data [4], simultaneous bio-based and biodegradable polymers had a 33% increase in applications within the plastic packaging industry between 2017 and 2019, particularly dominated by starch-based blends, polylactic acid (PLA), and polyhydroxyalkanoates (PHA) polymers (Figure 1). However, the number of scientific research articles focused on the development of bio-based and biodegradable formulations for food packaging application has had an even more notable increase, 78%, during the last three years. In this section, the available and emergent bio-based and biodegradable polymers under research, namely polysaccharides, proteins, and polyesters, presented in Figure 1, will be discussed as well as their production sources, technical properties, and challenges within the food packaging industry.

2.1. Polysaccharides

2.1.1. Starch

Starch is constituted by two distinct polymers, the mostly linear amylose, α-d-(1,4)-linked glucopyranosyl units, and the highly branched amylopectin, α-d-(1,4)-linked glucopyranosyl units partially substituted by α-d-(1,6) linkages units (Figure 2), with a proportion of 20–30% and 70–80%, respectively, depending on its botanical source [13,14,15]. It is considered one of the most promising natural biopolymers since it is easy to acquire, it can be extracted from foodstuffs byproducts (has four conventional sources such as wheat, corn, potato, cassava, and other sources including fruit waste [16]), and it is biodegradable and low-cost [17]. It is important to note that more than 310 tons of starch are industrially produced from corn, potato, or wheat for packaging purposes, accounting for approximately 25% of all used bio-based and/or biodegradable plastics (Figure 1) [4].
Starch has been used to form edible or biodegradable films and granulates through casting from aqueous solutions, film blowing, forming, injection molding, blow molding, and extrusion [18,19,20,21,22].
Plasticizers, along with temperature (for gelatinization) [16], allow the increase of the starch-based films flexibility, improving processability. Starch-based films have reasonable gas barrier properties; however, they still present some sensitivity to humidity conditions. Moreover, they also exhibit poor moisture barrier and mechanical properties, limiting their applications as packaging materials [23,24]. Furthermore, starch retrogradation naturally increases the matrix crystallinity over time, promoting brittle films [25,26], another factor that prevents the starch-based materials’ widespread adoption. Nevertheless, starch has been used as a filler in plastics, allowing to reinforce their structure [27,28]. Within commercial perspectives, plasticized starch is often blended with a range of other polymers, like polyethylene (PE), polypropylene (PP), polystyrene (PS), Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), or PLA in a proportion range between 30% and 70% [29], where biodegradability of the final product is dependent on the degradation abilities of the polymeric compounds [13]. There is a huge interest and new research focusing on the trade-off between biodegradability and functionality of starch-based plastics, promoting their efficiency and versatility. These blended materials have been considered as an alternative to pristine PE, PP, or PS-based plastics [4,30,31].
In a laboratorial environment, other interesting future scalable approaches have been explored to improve starch-based materials. Gonçalves et al. [16] used oil and waxes recovered from potato frying residues and potato peels, respectively, to tailor the surface properties (roughness and wettability) and flexibility of starch-based films. Chollakup et al. [32] found out that the incorporation of compounds such as cinnamon oil and fruit peel extract can provide antibacterial and antioxidant properties. Still, in both examples, the final product is not suitable for applications as food packaging at industrial scale. Currently, the best option is, as mentioned, to use starch blended with other more resistant materials (PE, PP, PLA, etc.).

2.1.2. Cellulose

Cellulose is a linear polymer of several hundred to many thousands of (β1→4) linked D-glucose units (Figure 2); it is the most abundant natural biopolymer on Earth [33], it is biodegradable, and it is commercially derived by a delignification process from wood pulp, which contains 40–50% of cellulose by weight [34]. This biopolymer, by itself, is unsuitable for film production due to its high crystallinity and long fibers [33,35], which make it have a non-thermoplastic nature (it cannot be softened or melted by the application of heat, nor can it be processed [36]) due to its strong intra- and inter-chain hydrogen bonds [37]. Chemically modified forms of cellulose fare a better chance against water, while exhibiting good mechanical properties, such as cellophane, cellulose esters (cellulose nitrate and cellulose acetate), and cellulose ethers (carboxymethyl cellulose and hydroxyethyl cellulose) [38].
A notable material derived from acid hydrolysis of cellulose is the nanocrystalline cellulose (CNC). It has been widely explored in numerous areas in food packaging (Figure 1) [39,40,41]. CNC has high elastic modulus, optical transparency, low thermal expansion coefficient, good gas barrier properties, and low toxicity, thus being biocompatible and biodegradable [42]. In packaging, CNC has been used as a gas barrier and filler of nanocomposites [39,40,43]. Microcrystalline cellulose (MCC) and nanofibrillated cellulose (NFC) have also used been for similar purposes [44,45]. Nanocellulose-based films are usually combined with other biopolymers, as chitosan and PHA, or plasticizers (e.g., glycerol, sorbitol, methoxypolyethylene glycol (MPEG) [46]) to improve or modify their physicochemical properties and extend their application range [47,48,49,50]. These biocomposites have shown excellent mechanical and oxygen barrier properties; however, their performances rapidly decline in the presence of moisture [49,50,51,52]. To overcome this drawback, nanocellulose grafting or blending with hydrophobic compounds as tannins, cholesterol, lignin, and fatty acids have been investigated [51,53,54]. However, these solutions often showed insufficient improvements on the hydrophobicity performance and thus are still unsuitable for using as food packaging materials.

2.1.3. Chitosan

Chitosan (CS) is a linear polysaccharide of 2-amido-2-deoxy-β-d-glucoses attached by (β1→4) linkages (Figure 2), obtained by the deacetylation of more than 50% of chitin [55,56]. Chitosan’s molecular weight depends on its source and, on average, can vary from 50 to 1000 kDa [57]. Manipulation of deacetylation percentages allows tailoring of its physicochemical properties and degradation profile [58]. Studies comparing low, medium, and high CS molecular weights found the latter molecular weight (>300 KDa) achieved better results in packaging [59,60]. CS is only soluble in acidic media, and its chains become positively charged when the pH level is under their pKa ≈ 6.5 [61,62]. CS also contains a primary amino group (NH2) that can be protonated to NH3+ and readily form electrostatic interactions with anionic groups in an acid environment. This characteristic allows the incorporation of a variety of chemical groups [63] (e.g., grafting phthalic anhydride, which increases antibacterial properties against Gram-positive and Gram-negative bacteria [64]) and letting prepared systems react to external stimuli such as temperature [65].
This biopolymer has been widely studied due to its potential in areas such as material science [66] and pharmacology [67], since it can be obtained at low cost, has large scale availability, is nontoxic, and is biodegradable [58]. Moreover, the antioxidant, antifungal, and antibacterial activities of CS has been attracting special attention in the food packaging sector as demonstrated in Figure 1 [68,69]. Additionally, because CS also has selective permeability to CO2 and O2, it has been investigated as edible coatings for fruit packaging applications. For this particular use, it delays the rate of respiration, decreases both weight loss and ethylene production, inhibits postharvest diseases, and increases the antioxidant process, allowing to extend the product shelf life while preserving the fruits overall quality. These results were achieved just by using CS as a packaging solution [17]. When applied to films formation, CS gives rise to transparent, flexible, and good oxygen barriers [17,70]. However, CS-based materials have some limitations such as low water vapor barrier characteristics and low mechanical strength; thus, they still require additives as fillers or plasticizers (e.g., glycerol [71]) to overcome these fragilities for food packaging applications [72,73].

2.1.4. Alginate

Alginate is the general name given to the family of linear polysaccharides consisting of binary copolymers, made up of (1–4) linked β-d-mannuronic acid (M) and α-l-guluronic acid (G) monomers (Figure 2), occurring in different proportions and distributions across the chain, depending on the source [74]. Thus, the molecular weight of alginate can vary between 32 and 400 kDa [75]. Alginate is usually extracted from brown algae, mainly Laminaria hyperborean, Macrocystis pyrifera, and Ascophyllum nodosum. In addition, microbial alginate can be produced by Azotobacter vinelandii and Pseudomonas aeruginosa [76]. The alginate ability to form strong gels/low-soluble polymeric materials in the presence of divalent cations, commonly Ca2+, due to the formation of a three-dimensional (3D) structure designed as an “eggbox” model, has been explored for the production of biodegradable and edible films [76]. The main advantages of alginate are: chemical stability, controllable swelling properties, and low content of toxic, pyrogenic, and immunogenic contaminants [74]. Alginate has been studied for new packaging options in the form of casted films [71,77] and as food-grade edible coatings [78]. Modified alginate such as propylene glycol alginate [79], sodium alginate [80,81], or calcium alginate [82] are frequently applied. The brittleness and low moisture barrier properties are the main disadvantages of alginate materials, which can be optimized with the use of diverse additives. Plasticizers such as glycerol and sorbitol improved the alginate films’ flexibility [71]. The combination of alginate with other biopolymers such as PLA can increase the tensile strength and the oxygen barrier properties [80]. In the same way, blends with biopolymers, e.g., soy protein, increased the tensile strength and decreased the water vapor permeability and water solubility [79]. Inorganic additives are also employed to enhance the performance of alginate materials. Sulfur nanoparticles with antimicrobial activity improved the mechanical properties of alginate films while increasing the UV barrier properties and hydrophobicity [83]. Similarly, alginate composites containing antimicrobial silver nanoparticles improved the shelf lives of fruits and vegetables [81].

2.1.5. Pullulans

Currently, the term ‘‘pullulan” is used in the literature to mean not only the ‘‘polymaltotriose”, the maltotriose -(α1à4)Glcp-(α1à4)Glcp-(α1à6)Glcp trimer (Figure 2), produced by different strains of fungus-like yeast Aureobasidium spp.- Aureobasidium pullulans, Aureobasidium melanogenum, and Aureobasidium mousonni, [84,85,86] but also other slightly different polysaccharide varieties (e.g., aubasidan-like, or pullulan-like), still similar to the pullulan, produced by distinct Aureobasidium pullulans varieties (e.g., Aureobasidium pullulans var. aubasidani and Aureobasidium pullulans var. pullulans, each with numerous distinct strains) [87].
This microbial extracellular polysaccharide is a biodegradable, biocompatible, non-mutagenic, nontoxic, hygroscopic (meaning that depending on the relative humidity at which it stored it can absorb water), non-carcinogenic, and edible polymer [88,89]. To create pullulan films or coatings, numerous techniques such as solvent casting [90,91], extrusion [92,93], coating by dipping or spraying [94,95], layer-by-layer assembly [96,97] or electrospinning [98,99,100] can be used. The resulting films are devoid of color, opacity, taste, and odor, and are also heat-stable and impermeable to both oil and oxygen [88,89]. Still, these films present drawbacks such as poor mechanical properties, namely their brittleness and their inability to resist water due to their hydrophilic nature and lack of active functions [88]. Thus, it is necessary to combine pullulans with other polymers and/or plasticizers [98,101,102] or even (nano)particles, which act as fillers [88,103,104], to improve the materials barrier properties (water and oxygen) and improve mechanical properties (e.g., tensile strength). The use of other materials such as silver particles can also impart new properties such as antimicrobial activity [103,105].

2.2. Proteins

2.2.1. Gelatin

Gelatin is derived from the fibrous insoluble protein called collagen (by thermal denaturation of collagen in the presence of diluted acid [106]) and is typically obtained from bones, skin, and connective tissue generated as waste during animal slaughtering and processing [107]. It is a heterogeneous mixture of single- or multi-stranded polypeptides, each with extended left-handed proline helix conformations and containing between 300 and 4000 amino acids most of which make up glycine, proline, and 4-hydroxyproline residues. Its structure is a mixture of α-chains (one polymer/single chain), β-chains (two α-chains covalently crosslinked), and γ-chains (three covalently crosslinked α-chains) [108]. Gelatin’s typical amino acid composition is Ala-Gly-Pro-Arg-Gy-Glu-4Hyp-Gly-Pro- (Figure 2) [107]. There are two types of gelatin of animal origin: Type A, with an isoelectronic point at pH ~8–9, obtained from acid treated collagen; and Type B, with an isoelectronic point at pH ~4–5, derived from an alkali-treated precursor which converts asparagine and glutamine residues into their respective acids, resulting in higher viscosity. The gelatin derived from pig skin is normally type A and the one from beef skin or pig cattle hides and bones is type B [109].
Gelatin is abundant, low-cost, and biodegradable [106]. It has been broadly studied in food packaging (Figure 1); however, it is hygroscopic, and consequently, it tends to swell or to be dissolved when put in contact with foodstuffs with high moisture content [106]. Another major disadvantage is its poor mechanical properties, especially when wet (poor water vapor permeability).
To overcome these issues, several approaches have been explored, namely addition of different molecules such as crosslinkers (to reduce solubility), plasticizers, and antimicrobial or antioxidant compounds [110,111]; combining gelatin with other biopolymers such as starch [112,113], chitosan [114,115], and whey protein [116], among others, to improve mechanical and water resistance properties; and adding nanoparticles such as silver, for instance, which can impart antimicrobial properties [114,116]. Recently, gelatin has been explored to create coatings and films for fruit packaging purposes due to its excellent ability to form films, its good oxygen barrier capacity, and UV-absorbing properties mediated by the presence of aromatic amino acids in its structure [117,118,119,120,121].

2.2.2. Zein

Zein, seen in Figure 2, is a natural prolamin which can be separated by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) into α-, β-, γ-, and δ-zein [122]. It is the main storage protein of corn at 45–50%, it is water insoluble (due to the hydrophobic character of its apolar amino acids, proline and glutamine [123]), and it is nontoxic and biodegradable [124,125].
Zein is an option for food packaging because it has excellent oxygen and carbon dioxide barrier properties and high thermal resistance [123,126]. Furthermore, in comparison with films produced from other proteins, it presents higher tensile strength and lower water vapor permeability [123]. However, films produced solely from zein present fragilities specifically related with brittleness, low surface functionality, and poor mechanical properties (such as lack of flexibility [127]) needed for industrial processing. Additionally, they cannot resist a high relative humidity condition. All these factors limit their use in food packaging applications [127].
To improve zein films, several different options have been explored: adding plasticizers (e.g., oleic acid, polyethylene glycol, glycerol, etc.) [128,129], adding micro- or nanoparticles (e.g., zinc or silver) [129,130], using other biopolymers to create better biopolymer blends (e.g., PP, PLA, etc.), [123,125,129,131,132] or adding other compounds (e.g., tannic acid, gamma-Cyclodextrin, etc.) which can provide different supplementary abilities (e.g., antioxidant, antibacterial properties) [130,132,133].

2.3. Polyesters

2.3.1. Polylactic Acid

Lactic acid is a monomer produced by fermentation of carbohydrates by bacteria, mainly Lactobacillus which by polymerization originates the polylactic acid (PLA—Figure 2), a thermoplastic aliphatic polyester [134]. PLA has good mechanical and thermal properties [17] which are strongly related to the ratio between the two mesoforms D and L [17]. PLA has high stiffness, strength, and water and oxygen permeability levels comparable to polystyrene [17,135]. PLA has excellent properties, namely high transparency, rigidity, and biodegradability, and it can be produced from renewable sources [37,136]. It can also tolerate various types of processing technologies, namely injection molding, extrusion, blow molding, and thermoforming [137]. However, PLA also has several shortcomings such as poor heat resistance, brittleness, poor melting strength, low degradation rate, and a narrow processing window [138,139,140]. To overcome these drawbacks, PLA-based products are being designed not only using PLA but also by mixing it with other biodegradable (bio)polymers and non-biodegradable resins and/or by compounding PLA with fillers such as fibers or micro- and nanoparticles [141,142,143]. Currently, PLA is being used in packaging applications as films, as thermoformed blow-molded containers and as short shelf life bottles [139,140,142,143]. Commercially, it is a good candidate to replace PE, PS, and polyethylene terephthalate (PET), which are non-biodegradable polymers commonly used for food packaging [37]. More than 160 tons of PLA are annually produced at the commercial level for plastic packaging purposes, which represent about 13% of all bio-based and/or biodegradable materials, making it the second most used biodegradable polymer (after starch-based blends) within this sector [4].

2.3.2. Poly(3-hydroxybutyrate-co-3-hydroxyvalerate)

Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), seen in Figure 2, is an aliphatic polyester from the PHA family, known by its biodegradability, nontoxicity, and biocompatibility [144]. PHA are linear polyesters that naturally occur in a variety of microorganisms and are accumulated intracellularly as carbon and energy reserves [13]. Currently, this natural ability of some microorganisms has been explored in the industrial synthesis of these biodegradable polymers by fermentation using renewable matrices as carbon sources [13,38]. Within the packaging sector, 96 tons of PHA were produced in 2019 from sugarcane pulp, sugar beet pulp, corn, potato, and wheat [29]. The majority of PHA are composed of 3-hydroxy fatty acid monomers [13]. A huge variety of PHA exist, and this diversity allows the production of PHA with a wide range of properties to compete with conventional plastic materials applied for food packaging [38]. Some PHA exhibit similar thermal and mechanical properties to PE, PS, and PP, which make them promising for food packaging applications [13,37,38]. However, this process is still not very cost effective and so the use of PHBV is limited due to its high production cost [144]. PHBV consists of a poly(3-hydroxybutyrate) or PHB with a few added 3-hydroxyvalerate (HV) units [145]. Moreover, PHBV has great oxygen barrier properties, chemical inactivity, high viscosity in a liquid state (good for extrusion processes), and has better mechanical properties, namely increased surface tension and more flexibility compared to PHB [144,145]. Nonetheless, PHBV still has some considerable deficits, namely brittleness, a narrow processing window, poor thermal and mechanical properties, no antimicrobial properties, and low resistance to water vapor permeability [145,146]. To improve these properties, fillers as clays, cellulose nanocrystals, and metal oxides [146,147,148,149] have been successfully used, enhancing thermal stability, mechanical properties, and barrier properties [146,147,148,149].

3. Graphene Derivatives-Based Biocomposites as Food Packaging Materials

As mentioned in the previous section, biopolymers by themselves do not present all the necessary requirements intended for food packaging, namely in terms of mechanical resistance and water/gas barrier properties. Furthermore, there is an ongoing trend towards active and intelligent packaging, i.e., approaching packaging that is able to interact with food to actively extend its shelf life, and to inform the consumer about the safety state of the food products, respectively [150]. In this context, graphene derivatives may play a role on assigning activity and intelligence to the packaging, in addition to other properties.

3.1. Graphene Derivatives

Graphene is one of the most promising nanomaterials which is constituted by a flat monolayer of carbon atoms in a two-dimensional (2D) hexagonal lattice, held together by a backbone of overlapping sp2 hybrids bonds [151]. This structure confers to graphene many remarkable properties, such as the strongest mechanical robustness with a modulus of over 1060 GPa [152], without losing its molding properties, large surface area [153], impermeability to gases [154], and optical transparency [155], among others. Graphene was considered by the scientists A.K. Geim and K.S. Novoselov, who isolated it for the first time, to be the mother of all graphitic forms because it is the 2D building material for carbon structures of all other dimensionalities, and because it can be encased into 0D fullerenes, folded into 1D nanotubes, or stacked into 3D graphite [151]. When short stacks of graphene sheets are packed, having a platelet shape, they receive the name of graphene nanoplatelets (GNP) [156]. Graphene oxide (GO) is a derivative of graphene obtained from graphite in two steps. The first step is the chemical oxidation (with strong acids) of graphite into graphite oxide, followed by simple stirring or mild sonication (mechanical process) to exfoliate graphite oxide into single layers [157]. Thus, GO is a carbon layer with several oxygen functional groups (carboxylic, hydroxyl, and epoxy) on its basal planes and at its edges, resulting in a hybrid structure of sp2 and sp3 configurations [158]. The high oxygen content has been demonstrated to be very useful for the chemical modification/functionalization with other molecules, thus allowing its dispersion in different matrices enabling the preparation of nanocomposites with interesting properties [159,160]. Globally, GO retains plenty of the properties of the graphene, but it is much easier and cheaper to prepare in bulk quantities and to process (better dispersion in different solvents). In fact, the chemical exfoliation of graphite in oxidative medium originates stable aqueous suspensions of GO. Depending on the practical application of GO, a reduced form can be more suitable—for example, to be incorporated in a polymeric matrix of hydrophobic nature [161]. Thus, GO can be reduced (rGO) by chemical, thermal, microwave, photo-chemical, and photo-thermal or microbial/bacterial methods, where the material can recover partial or complete hybridized sp2 configuration, thus approaching the graphene configuration [162,163,164].
Interestingly, graphene derivatives have been widely used as polymer reinforcement and are well known to impact on several properties of the final nanocomposite, namely on its mechanical, thermal, electrical, conductive, and fire retardancy properties, to name just a few [150,151,152,153,154,155,156]. Due to the already described variety of graphene derivatives, the term “graphene” is often used in a generic manner to describe indifferently each of these nanostructures, thus creating misinterpretations about its properties [165]. Owing to the chemical and structural features presented by each graphene derivative, they have different properties that make them proper additives for specific applications. The most fundamental properties of graphene derivatives to consider are: (i) number of graphene layers, (ii) average lateral size, and (iii) oxygen content (with a variable carbon-to-oxygen (C/O) atomic ratio). The scheme presented in Figure 3 helps to visualize the categorization of different graphene derivatives types according to the three fundamental properties mentioned previously [165].
The use of graphene derivatives in biopolymer-based composites for food packaging applications is still lightly explored. The Web of Science search of the topic “biopolymers food packaging composites” displays 234 results, but when refined including the word “graphene”, only nine articles are displayed (accessed on 22 September 2020). However, given the potential of these nanomaterials to improve determining properties of biopolymeric composites for food packaging application, this will certainly be an expanding area with high impact in the future. Hereafter, some of the most relevant properties obtained for biopolymer-based composites including graphene derivatives will be reviewed.

3.2. Properties of Biopolymer-Based Composites with Graphene Derivatives

The first step in the development of graphene-based biopolymer composites is to guarantee the uniform dispersion of graphene materials into a polymer matrix, which might not be easy given the high propensity for self-agglomeration of graphene-based materials, as a result of the strong van der Waals forces and ππ electrostatic interactions between nanosheets or nanotubes, and/or their common weak dissolvability in water and organic solvents during biocomposite fabrication [152]. The solution mixing is considered the most common and simplest method to prepare polymer composites. In this method, fillers are mixed with the polymer solution, homogenized by physical stirring, and the composites recovered after solvent evaporation [166]. However, achieving a good dispersion is still a challenge. The low dispersion of these nanofillers and the consequent weak interfacial adhesion between the graphene-based filler and the host biopolymer can compromise the final morphological, mechanical, barrier, and physicochemical properties of the films [167]. Sonication, ultrasonication, high-speed blending, and melt blending are some strategies used to achieve a proper dispersion of these nanofillers in the bio-based matrices [168,169,170]. When these methodologies are not enough to achieve a good dispersion, the wrapping of surfactants using noncovalent interaction methods [171] or the chemical modification of graphene materials by covalent bonding [145,167] are some of the approaches used. Recently, as an alternative to chemical modification, oil-in-water Pickering emulsions were used to disperse carbon nanotubes (CNT) into PLA matrix using cellulose nanocrystals as stabilizer and as dispersant due to its amphipathic character [172]. After the dispersion, films are commonly obtained by solution casting method [7,170] or by melt processing, which is a more convenient method for the potential industrial application [173,174,175]. The final structural organization of these composites containing graphene derivative-based fillers (originally suggested for layered silicates) is categorized in three types, namely phase separated, intercalated, and exfoliated (Figure 4). The performance of composites is maximized in their exfoliated form, corresponding to the carbon nanostructures well dispersed within the matrix. This condition provides a tortuous diffusion barrier and creates a good percolation network, which greatly enhances the overall properties [176].

3.2.1. Mechanical and Thermal Stability Properties

The preparation of biopolymers nanocomposites with GNP [152,173,178,179], chemically modified rGO [145,167,180,181], or CNT [172,174,175,182] reinforcement has exhibited significant enhancements in the mechanical properties of chitosan, PLA, alginate, PHBV, cellulose nanofibers, and starch even at very low concentrations (<5%), as described in Table 1. The Young’s modulus, tensile strength, and elongation at break are increased with the incorporation of the diverse graphene derivatives as fillers. For example, compared with the neat biopolymer, the incorporation of only 0.7 wt% rGO into a PHBV-based formulation increased these properties by 100%, 119% and 24%, respectively [145]. In turn, the incorporation of 2 wt% of CNT into a PLA matrix also provides an improvement of 52% and 36% on the materials’ tensile strength and elongation at break, respectively [7]. There is, therefore, a reinforcement of mechanical properties by the incorporation of graphene derivatives, and the different morphology of these nanostructures does not seem to have a predominant or limiting role for this mechanical improvement [183]. In some cases, the incorporation of graphene derivatives into biopolymer-based formulations could result in a slight decrease of elongation at break, meaning that these composites can present lower flexibility [152,170,178]. However, this drawback can be overcome or limited through the synergism of graphene derivatives with other nanofillers, such as cellulose nanocrystals (CNC) or ZnO [173,178]. The simultaneous use of different graphene derivatives as fillers could also be a strategy to enhance the mechanical performance of final composites since the reinforcement of PLA-based films with a 0.4 wt% of a hybrid co-filler with added single-walled carbon nanotubes (SWCNT) and GO resulted in an increased tensile strength and Young’s modulus up to 75% and 130%, respectively [7]. Mechanical properties of biocomposites are directly related to the biopolymer(s) interaction type established with the carbon filler(s), such as electrostatic interactions, van der Waals forces, or hydrogen bonding between the hydrophilic groups of the biopolymer(s) and the oxygenated functional groups on carbon nanostructures, or chemical bonding when grafting processes are applied as illustrated on Figure 4. This interaction can improve the stress transfer mechanism between graphene derivatives and the biopolymer matrix, thus increasing the material rigidity and tensile strength of the nanocomposite [145,167]. In the case of GO or chemically modified GO, mechanical properties are, therefore, strongly influenced by the degree oxidation [152], since the incorporation of the same amount of GO (0.6 wt%) prepared with growing ratios of KnMO4/graphite (from 2:1 to 8:1) into a CS matrix generated variable mechanical properties: Young’s modulus decreasing and increasing by −20% to 307%, tensile strength increasing by 15% to 327% and elongation at break increasing and decreasing by +1% to −29% with the oxidation degree increase when compared with this neat biopolymer. The increase of the oxidation degree led to an increase of the rigidity and strength, accompanied by a stretchability decrease, which is linked to the chain mobility restriction as a result of more hydrogen bonds forming. The oxidation degree of rGO is, thus, a factor that should be taken into account on the optimization of biocomposites processability. In addition, the grafting of rGO with cellulose nanocrystals (rGO-CNC) potentiates the formation of hydrogen bonds between biopolymer/graphene derivative, and thus better mechanical results are achieved even when compared with the simultaneous incorporation of non-grafted rGO and CNC [145]. The combined use of rGO and CNT seems to potentiate the Young’s modulus and tensile strength increase [184]. Generally, the upper reinforcement degree of rGO or CNT and their chemically modified forms is usually reflected on the improvement of tensile strength and Young’s modulus. However, at high concentrations (>20 wt%), graphene derivatives can introduce defects in the biocomposite microstructure, compromising the further improvement of mechanical properties [179,183]. In addition, it is still necessary to assess the mechanical performance of graphene-based biocomposites as food packaging relative to shelf life. Food products emit compounds such as CO2 which, in the presence of water, form carbonic acid that can interact with the biocomposites and influence their resistance [185].
The thermal stability of the biocomposites can also be improved with the inclusion of graphene derivatives into biopolymer-based formulations. The temperature of degradation is delayed by 5 °C up to 20 °C, when compared with the biopolymer film without these nanofillers [145,167]. The high heat resistance of graphene derivatives seems to improve the thermal stability of biocomposites since the introduction of these nanofillers into biopolymer matrix creates an effective heat and gas barrier which prevents the diffusion of volatile degradation products and radicals, slowing the biocomposite thermal degradation [145,167].

3.2.2. Barrier Properties

Graphene-based biocomposites exhibit great barrier properties against gases, water vapor, and UV light [7,145,152,167,173,179]. The gas barrier properties have a special importance on modified atmosphere in food packaging. Water vapor barrier properties should avoid the water penetration on packaging atmosphere or the dehydration of foodstuffs, and UV light barrier should avoid the transmittance of UV light across packaging materials and the subsequent degradation of organic compounds present in the food matrices, thus preventing its deterioration and quality lost. A hybrid co-filler made with SWCNT and GO fillers with a loading of only 0.4% into PLA films decreased the oxygen transmission rate by 67% and diminished the transmission of ultraviolet-visible light by 30%. Manikandan et al. [8] proved that the use of polyhydroxybutyrate biocomposite containing 0.7 wt% GNPs had improved water, oxygen, and UV barrier properties and was also able to promote a fourfold increase of potato chips and milk products shelf life. The improvement in gas and water barrier of graphene-based biocomposites is explained by the introduction of impermeable graphene-based nanofillers with large surface area which restricts the motions of the biopolymer chains, and the creation of a tortuous pathway into the film matrix that complicates gas molecules diffusion.

3.2.3. Surface Hydrophobicity Properties

Surface hydrophobicity of biopolymer-based materials is also enhanced by the incorporation of graphene derivatives. As described in the Section 2, the hydrophilic character of the bio-based and biodegradable polymers is the main limiting factor for their application. In contact with food, these biopolymers can absorb water or even solubilize and, concomitantly, compromise their protective function as packaging films. The integration of graphene derivatives into biocomposites increases their surface hydrophobicity and decreases their water solubility. Recently, the use of 0.6 wt% of rGO grafted with maleic anhydride and subsequently with dodecyl amine increased the contact angle of PLA/starch composite from 67° to 81° [167]. Furthermore, solubility of CS in water decreased from 38% to 22% with rGO at 0.5 wt% [152]. These values are promising, but they are not enough for these composites’ application as food packaging films.

3.2.4. Biodegradability

The development and optimization of graphene-based biocomposites is ongoing and it is important to prove that the biodegradable properties of natural polymers are not compromised. The biodegradation of polymers occurs as a result of the activity of specific microorganisms to hydrolyze and oxidize these molecules in a short time after their disposal. The strong interaction between graphene fillers and the biopolymer can improve the mechanical stability, as previously mentioned; however, the degradation rate can be slightly reduced by delaying the diffusion of water into the polymeric matrix [189]. Lyn et al. [152] supported this hypothesis by demonstrating that solvent casting CS/rGO composites biodegrade completely after 20 days of composting. On the contrary, melt processing films of PLA/starch blends with 5 wt% of rGO have 17–26% higher weight loss than the blends without the rGO in 183 days of controlled aerobic composting process [190]. Nonetheless, despite the degradation rate decreasing, the biocomposites biodegradability is not compromised [152,189,190].

3.2.5. Active Properties

Nowadays, in addition to the desirable biodegradable packaging, there are global research interests for the development of active food packaging to extend shelf life, enhance safety, and maintain the organoleptic properties. Biocomposites with graphene derivatives can exhibit antioxidant, antimicrobial, and antifungal activity as extensively reviewed by Carvalho et al. [191]. The radical scavenging capacity of rGO can provide antioxidant activity to biocomposites. CS-based films with 20–33 wt% of rGO showed an increase of inhibition in the range of 54% to 82% after 8 h of incubation [170] and this activity can avoid the oxidation of packaged foodstuffs. The antimicrobial and antifungal properties of graphene-based nanostructures are based on their capacity to induce cell membrane disruption and oxidative stress that compromise bacterial proliferation and sporulation [191]. Biocomposites with graphene derivatives revealed in vitro antibacterial properties against a broad spectrum of pathogenic microorganisms (such as Enterococcus faecalis, Staphylococcus epidermidis, Escherichia coli, Staphylococcus aureus, Staphylococcus haemolyticus, and Bacillus subtilis). The surface modification of graphene derivatives with essential oils or other metal compounds such as Ag, ZnO, or TiO2 has been adopted as strategy to enhance the inherent antimicrobial potential of these nanostructures [191]. Recently, a PLA/CNT/cinnamaldehyde film revealed great potential application as a controlled-release antibacterial active food packaging film with an active effect proven up to 21 days [192]. Antifungal activities of biocomposites containing carbon nanostructures against Aspergillus niger, Cryptococcus neoformans, Candida tropicalis, Candida albicans, Botrytis cinereas, and Rhizopus spp. have also been described [191].
Alternatively, the delocalized conjugated electron structure of graphene derivatives has been recently explored for production of photocatalytic hybrid systems with application in organic compounds’ degradation [186,193]. Organic compounds’ degradation might have special importance for controlling the composition inside of the modified atmosphere packaging in order to extend the shelf life of food products. The organic compounds’ degradation ability of visible light-responsive GO/Bi2WO6 hybrid system was shown to be 4.4 times more effective on ethylene degradation than pure Bi2WO6 when incorporated into starch-based films [193]. Ethylene is a phytohormone responsible for inducing the fruit ripening process and its accumulation inside packaging leads to senescence of fruit. The application of starch/GO/Bi2WO6 composites as active packaging materials of highly perishable fruit can be an opportunity to decrease fruit losses and wastes [193].
Therefore, the use of graphene derivatives into biopolymer-based composites reveals to be a promising approach to develop active food packaging, enhancing the preservation of food products.

3.2.6. Clay–Graphene Bionanocomposites

Nano- and micro-particulated clay minerals could be incorporated as fillers leading to biopolymer–clay nanocomposites (bionanocomposites) based on the assembly of polysaccharides, proteins, polyesters, etc., with layered silicates, such as smectites and vermiculites, as well as fibrous clays like sepiolite and palygorskite [194,195,196,197,198,199,200,201]. These clay-based biocomposites are of great interest for diverse applications, including films and coatings for use in food packaging, as they can improve barrier and mechanical properties, in addition to conferring water resistance and other characteristics very attractive for these types of uses [200].
Clays could be assembled to carbon particles producing graphene-like materials of variable composition and properties, including adsorption ability and electrical conductivity, mainly afforded by the clay component and the graphene derivative component, respectively. These carbon–clay materials could be synthetized following two main processing approaches: (i) top-down, by crushing together both components, and (ii) bottom-up, by growing carbon on clay minerals used as supports [202,203]. Another interesting preparative way to develop carbon–clay composites is based in the use of the fibrous clay known as sepiolite, which is processed and commercialized as a rheological grade product (Pangel®). It is well known that the dispersion of carbon NP—for instance, carbon nanotubes—in aqueous medium [204,205,206,207,208] is of paramount importance to homogeneously disperse carbon NP as fillers in water soluble biopolymers. An efficient approach for this purpose is based on the use of the rheological grade sepiolite assisted by ultrasonic irradiation in the presence of those carbon NP leading to dispersions stable for several months without syneresis effects [209]. For instance, the stabilization of MWCNT and GNP was initially interpreted in terms of steric stabilization where sepiolite nanofibers act as interposed species in between the carbon NP avoiding its reassembly [209]. Based on this behavior, diverse films of carbon–clay bionanocomposites containing gelatin, polyvinyl alcohol and alginate provided with electrical conductivity have been reported [210].
Clay–graphene materials dispersed in diverse biopolymers are fillers of great interest because they can introduce significant improvement on the resulting biocomposites, increasing their barrier properties and conferring electrical conductivity. In fact, synergistic effects of clay minerals and graphene-based materials are observed for the use of clay–graphene systems as fillers of polymers. For instance, GNP/sepiolite can be used as fillers of alginate leading to films showing simultaneously reinforcing properties together marked in-plane conductivity (ca. 500 S m−1), which increases by the additional incorporation of MWCNT favoring percolation reaching values up to 2500 S m−1 [202,210]. These alginate-based biocomposite films show variable mechanical properties (e.g., elongation at break and Young’s modulus) depending on the proportion of the starting components, i.e., sepiolite, GNP, and alginate [210]. A clear disadvantage of the use of sepiolite in the preparation of these carbon–clay fillers can be the possible loss of mechanical properties. Moreover, high clay content significantly reduces the electrical conductivity in detriment of its use as films for pulsed electric field (PEF) application.

3.3. Emerging Application for Biopolymer-Based with Graphene Derivatives

The use of electrically conductive graphene derivatives in food packaging can be an important feature for the development of food processing technologies. Food processing by pulsed electric field (PEF) is one of the most promising emerging non-thermal technologies. The PEF treatment consists in the application of short high voltage pulses to food products to inactivate enzymes and microorganisms, while maintaining their sensorial and nutritional properties [211,212,213]. Currently, foodstuffs are processed before packaging in direct contact with the chamber electrodes [214]. This methodology compromises the food safety due to metals releasing from the stainless steel electrodes and post-sterilization recontamination events (before food packaging) [215]. The release of metals from electrodes could be avoided by covering the electrodes with an electrically conductive material, while the recontamination could be prevented if the PEF treatment occurred after packaging (Figure 5). Roodenburg et al. [216] used a commercially available electrically conductive plastic, commonly used to pack electronic components, to cover the chamber electrodes. The use of this composite constituted by ethylene vinyl acetate and 30 wt% carbon black particles, with an electrical conductivity of 0.75 S m−1, led to a bacterial inactivation of 2.1 log10. The same authors also studied the effectiveness of PEF in-pack, using this composite material as a food packaging pouch. The results showed a bacterial inactivation of 5.9 log10, which reaches the pasteurization level [217].
After the PEF proof of concept, it is necessary to develop electrically conductive materials suitable for food packaging. Biopolymer-based composites containing graphene derivatives can be a solution. In this regard, CS-based biocomposites containing rGO were recently suggested for electrically conductive food packaging [170]. In this work, rGO was prepared using caffeic acid as reducing agent, representing a green alternative to toxic reducing agents that should be avoided in food packaging applications. The bionanocomposite films containing 50 wt% rGO showed an electrical conductivity of 0.7 S m−1 and 2.1 × 10−5 S m−1, in-plane and through-plane, respectively. CS-based flexible films containing rGO-Fe3−xO4 were also suggested for electrically conductive food packaging [12]. These sustainable bionanocomposites, produced in absence of toxic chemicals, achieved an in-plane electrical conductivity of ~0.016 S m−1 with 50 wt% rGO-Fe3−xO4. In both works [12,170], the electrically conductive films also display mechanical properties and antioxidant activity attractive for food packaging applications. However, to the best of the authors knowledge, there are no other reports specifically addressing this issue. Therefore, Table 2 lists reports from the last three years regarding electrically conductive biocomposites suggested for other applications that are considered to have potential for electrically conductive food packaging. The graphene derivatives found as fillers in these biocomposites were MWCNT, SWCNT, rGO, GNP, and graphene. PLA, cellulose, and CS are the most used biopolymer matrices. Polymer blends are a strategy to improve the mechanical properties of the matrix and the dispersion of carbonaceous fillers [190,218]. Films containing GNP, as conductive filler, and a matrix of a blend of PLA biopolymer and poly (butylene adipate-co-butylene terephthalate) (PBAT) synthetic biodegradable polymer were successfully prepared [219]. PLA has a weak affinity with GNP, while PBAT has a good affinity thus enabling to hold high GNP loadings. This strategy created conductive channels by confining GNP into the PBAT continuous phase and constructed good percolation networks leading to an electrical conductivity of 338 S m−1. The MWCNT were the most used fillers due to their high electrical conductivity capacity, typically required for electronic applications [172,174,175,201,218,220]. MWCNT were also used is small amounts as doping materials to enhance the electrical conductivity of GNP in the CS matrix [221]. GNP and rGO are cost-effective alternatives given the large-scale production of food packaging [222]. However, the rGO with high electrical conductivity is usually prepared using toxic reducing agents that prevent its use for food packaging. On the other hand, the use of nontoxic reducing agents produces rGO with lower electrical conductivity [12,170,190,223,224]. The in-situ reduction of GO into NFC with hydroiodic acid (HI) lead to an in-plane electrical conductivity of 22.22 S m−1, while the same procedure using ascorbic acid lead to an electrical conductivity of 0.83 S m−1. The films thermally reduced (TR) at 450°C achieved the highest conductivity of 23.42 S m−1, but this method weakened the mechanical properties [225].
The food packaging materials for PEF treatment in-pack should have a through-plane electrical conductivity close to the electrical conductivity of the packaged food, which is typically between 0.1–2 S m−1 [216]. Many of the biocomposites listed in Table 2 present superior electrical conductivity to the food conductivity. However, these values refer to the in-plane conductivity, since the through-plane conductivity is not reported in most cases. These fillers have a preferential alignment in the plane direction, leading to composites with low through-plane conductivity, even in cases of high in-plane conductivity [170,210]. Thus, the through-plane electrical conductivity of graphene-based nanomaterials is very poor [226]. Therefore, the development of novel strategies to improve the through-plane conductivity are paramount to this emerging application.
Table 2. Biopolymer-based composites containing graphene derivatives suitable for electrically conductive food packaging recently described in literature.
Table 2. Biopolymer-based composites containing graphene derivatives suitable for electrically conductive food packaging recently described in literature.
BiopolymerGraphene Derivativeσ (S m−1)ApplicationsRef.
PLA4.3 wt% MWCNT59.30EMI shielding[172]
15 wt% GNP0.35-[227]
3 wt% MWCNT6.42EMI shielding[175]
2 wt% MWCNT19.70EMI shielding[174]
5 wt% SWCNT1010Organic devices[228]
15 wt% GNP0.36Electronics[229]
PLA/Starch
PLA/PBAT
Cellulose derivatives
5 wt% rGO0.001Packaging[190]
40 wt% GNP338Electronic devices[219]
5 wt% rGO15,200Electronic devices[230]
4.5 wt% MWCNT10Electrochemical devices[201]
10 wt% MWCNT37.6Electronics[231]
9 wt% graphene2.4Diverse[223]
9 wt% rGO1.4Diverse[223]
50 wt% rGO (AC)0.83Electronic devices[225]
50 wt% rGO (HI)22.22Electronic devices[225]
50 wt% rGO (TR)23.42Electronic devices[225]
Cellulose/SPI
CS
0.25 wt% *MWCNT0.82-[218]
50 wt% rGO0.7Food packaging[170]
50 wt% rGO-Fe3-xO40.016Biomedical[12]
55 wt% GNP/5 wt% MWCNT2900Bioelectrocatalysis[221]
2.5 wt% rGO0.08Biomedical[224]
PLA: polylactic acid. MWCNT: multi-walled carbon nanotubes. EMI: electromagnetic interference. GNP: graphene nanoplatelets. SWCNT: single-walled carbon nanotubes. rGO: reduced graphene oxide. PBAT: poly (butylene adipate-co-butylene terephthalate). AC: ascorbic acid. HI: hydroiodic acid. TR: thermally reduced. SPI: soy protein isolate. CS: chitosan.

4. Future Perspectives

The returns of the combination of graphene derivatives with biopolymers were reviewed. The electrical conductivity provided by the graphene derivatives opens the way for the development of novel and auspicious materials and technologies, namely electrically conductive packaging materials for food processing by PEF.
The electrically conductive biocomposites show great potential to be used in intelligent packaging as flexible platforms to incorporate specific sensing molecules since the target detection would also promote changes in the electrical conductivity of the biocomposites. In this context, a PLA-based sensor containing a layer of paper coated with GO was recently described [232]. This sensor monitors the food quality through the detection and quantification of biogenic amines. The PLA biopolymer was used as a matrix to entrap colorimetric sensing compounds that detect the biogenic amines, while the ability of GO to adsorb and desorb organic molecules allowed its quantification by laser desorption–ionization mass spectrometry. This research area is expected to expand in the near future.
The presence of clays assembled to graphene derivatives is very attractive for the preparation of multifunctional nanoplatforms with potential applications in food packaging. As reported above (Section 3.2.6), the barrier properties regarding gases and water of layered graphene derivatives such as GO and GNP could be enhanced by the simultaneous presence of clays—in particular, for smectite silicates like montmorillonite. Carbon–clay materials could be functionalized by the immobilization of semiconducting NP such as TiO2 or ZnO on the clay surface [233]. This strategy appears as a promising way to prepare improved UV-shielding films for packaging food, taking into account the high performance of this type of semiconductor as a UV barrier in biopolymer films [234], given that a synergistic effect with the graphene derivatives component present in these composites was expected in our case. Remarkably, these semiconducting NP are also provided with bactericidal activity. In addition, it may be considered that clay minerals are efficient adsorbents for many diverse organic compounds as well as support of metal and metal oxide nanoparticles [235,236,237]. Therefore, the presence of clays in the carbon–clay materials could be of interest to modulate the properties of the resulting films based on clay–graphene biocomposites. For instance, bioactive compounds such as bactericide or antifungal drugs might be easily incorporated into the pores of sepiolite and other clay silicates acting as carriers of these bioactive agents. Moreover, characteristics as scavengers of oxygen can be introduced in the fillers by anchorage of metal oxide NP (e.g., iron-oxides) on the external surfaces of clays.
Prior to the application of these materials, it is important to understand the risks of their migration into food. Until now, there are few literature evidences, and those studies indicate that there is no migration of graphene derivatives into food, or that the migration occurs far below the international migration limits [145,169]. Additionally, these materials also constrain the diffusion and migration of plasticizer molecules for food systems [167]. Still, there is a need to further investigate the risks involved in the use of graphene derivatives in biopolymers. Notably, recent studies sustained by sound science-based assessment of the potential impact on health and environment are being developed with the aim to understand the graphene derivatives properties that control their biological effects with promising results towards their safety application under controlled safe-by-design approaches [238,239,240].

Author Contributions

P.A.A.P.M. and P.F.: conceptualization; A.B., J.D.C.S., and M.R.F.S.: investigation, writing—original draft preparation, review, and editing; C.N., E.R.-H., I.G., S.Y., P.F., and P.A.A.P.M.: supervision, writing—review, and funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the projects UIDB/00481/2020 and UIDP/00481/2020—FCT—Fundação para a Ciencia e a Tecnologia; and CENTRO-01-0145-FEDER-022083—Centro Portugal Regional Operational Programme (Centro2020), under the PORTUGAL 2020 Partnership Agreement, through the European Regional Development Fund. Also, under the scope of the projects CICECO-Aveiro Institute of Materials (UIDB/50011/2020 & UIDP/50011/2020), M-ERA-NET2/0021/2016—BIOFOODPACK—Biocomposite Packaging for Active Preservation of Food, NANOTRONICS (IF/300/2015). These projects are financed through the FCT/MCTES under the Competitiveness and Internationalization Operational Program (COMPETE), by the European Regional Development Fund (FEDER) and Portugal 2020. This research was funded in part by MINECO (Spain) and FEDER (EU) funds (project MAT2015-71117-R). AB, JS, MS, IG and PF thank FCT for the grants (SFRH/BD/148856/2019, SFRH/BD/145660/2019, SFRH/BD/145661/2019, CEECIND/00430/2017 and IF/00300/2015, respectively). This work was also funded by national funds (OE), through FCT, I.P., in the scope of the framework contract foreseen in the numbers 4, 5 and 6 of the article 23, of the Decree-Law 57/2016, of 29 August, changed by Law 57/2017, of 19 July.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. UN Sustainable Development Goals. Available online: https://www.un.org/sustainabledevelopment/sustainable-development-goals (accessed on 20 September 2020).
  2. Yildirim, S.; Röcker, B.; Pettersen, M.K.; Nilsen-Nygaard, J.; Ayhan, Z.; Rutkaite, R.; Radusin, T.; Suminska, P.; Marcos, B.; Coma, V. Active Packaging Applications for Food. Compr. Rev. Food Sci. Food Saf. 2018, 17, 165–199. [Google Scholar] [CrossRef] [Green Version]
  3. Shapi’i, R.A.; Othman, S.H.; Nordin, N.; Kadir Basha, R.; Nazli Naim, M. Antimicrobial properties of starch films incorporated with chitosan nanoparticles: In vitro and in vivo evaluation. Carbohydr. Polym. 2020, 230, 1–6. [Google Scholar] [CrossRef]
  4. Institute for Bioplastics and Biocomposites European Bioplastics. Available online: https://www.european-bioplastics.org/market/ (accessed on 19 August 2020).
  5. Report Linker. Available online: https://www.reportlinker.com/market-report/Packaging/6286/Packaging?gclid=CjwKCAjwnK36BRBVEiwAsMT8WNhZAbfnSMF48UwrjnqHM4QD24JfVUjqiN5HmTXzuMhLAC0s-LLqkBoCYzYQAvD_BwE&fbclid=IwAR120hj1aHAXE4k4qhK_o032LoO6XCLeqMx-vhe3mSBXIbnNx-A-pc-dyQU (accessed on 6 September 2020).
  6. RameshKumar, S.; Shaiju, P.; O’Connor, K.E.; P, R.B. Bio-based and biodegradable polymers - State-of-the-art, challenges and emerging trends. Curr. Opin. Green Sustain. Chem. 2020, 21, 75–81. [Google Scholar] [CrossRef]
  7. Kim, Y.; Kim, J.S.; Lee, S.-Y.; Mahajan, R.L.; Kim, Y.-T. Exploration of hybrid nanocarbon composite with polylactic acid for packaging applications. Int. J. Biol. Macromol. 2020, 144, 135–142. [Google Scholar] [CrossRef]
  8. Manikandan, N.A.; Pakshirajan, K.; Pugazhenthi, G. Preparation and characterization of environmentally safe and highly biodegradable microbial polyhydroxybutyrate (PHB) based graphene nanocomposites for potential food packaging applications. Int. J. Biol. Macromol. 2020, 154, 866–877. [Google Scholar] [CrossRef]
  9. Chen, P.; Xie, F.; Tang, F.; McNally, T. Structure and properties of thermomechanically processed chitosan/carboxymethyl cellulose/graphene oxide polyelectrolyte complexed bionanocomposites. Int. J. Biol. Macromol. 2020, 158, 420–429. [Google Scholar] [CrossRef]
  10. Qiu, Y.; Wang, Z.; Owens, A.C.E.; Kulaots, I.; Chen, Y.; Kane, A.B.; Hurt, R.H. Antioxidant chemistry of graphene-based materials and its role in oxidation protection technology. Nanoscale 2014, 6, 11744–11755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Nunes, C.; Coimbra, M.A.; Ferreira, P. Tailoring Functional Chitosan-Based Composites for Food Applications. Chem. Rec. 2018, 18, 1138–1149. [Google Scholar] [CrossRef]
  12. Barra, A.; Alves, Z.; Ferreira, N.M.; Martins, M.A.; Oliveira, H.; Ferreira, L.P.; Cruz, M.M.; de Deus Carvalho, M.; Neumayer, S.M.; Rodriguez, B.J.; et al. Biocompatible chitosan-based composites with properties suitable for hyperthermia therapy. J. Mater. Chem. B 2020, 8, 1256–1265. [Google Scholar] [CrossRef] [PubMed]
  13. Lambert, S.; Wagner, M. Environmental performance of bio-based and biodegradable plastics: The road ahead. Chem. Soc. Rev. 2017, 46, 6855–6871. [Google Scholar] [CrossRef] [PubMed]
  14. Salit, M.S. Tropical Natural Fibre Composites—Chapter 3 Biopolymer. In Tropical Natural Fibre Composites; Engineering Materials; Springer: Singapore, 2014; p. 41. ISBN 978-981-287-154-1. [Google Scholar]
  15. Niranjana Prabhu, T.; Prashantha, K. A review on present status and future challenges of starch based polymer films and their composites in food packaging applications. Polym. Compos. 2018, 39, 2499–2522. [Google Scholar] [CrossRef]
  16. Tan, I.; Torley, P.J.; Halley, P.J. Combined rheological and optical investigation of maize, barley and wheat starch gelatinisation. Carbohydr. Polym. 2008, 72, 272–286. [Google Scholar] [CrossRef]
  17. Tang, X.Z.; Kumar, P.; Alavi, S.; Sandeep, K.P. Recent Advances in Biopolymers and Biopolymer-Based Nanocomposites for Food Packaging Materials. Crit. Rev. Food Sci. Nutr. 2012, 52, 426–442. [Google Scholar] [CrossRef]
  18. Bagheri, R. Effect of processing on the melt degradation of starch-filled polypropylene. Polym. Int. 1999, 48, 1257–1263. [Google Scholar] [CrossRef]
  19. Thunwall, M.; Kuthanova, V.; Boldizar, A.; Rigdahl, M. Film blowing of thermoplastic starch. Carbohydr. Polym. 2008, 71, 583–590. [Google Scholar] [CrossRef]
  20. Tabi, T.; Kovacs, J.G. Examination of injection moulded thermoplastic maize starch. Express Polym. Lett. 2007, 1, 804–809. [Google Scholar] [CrossRef]
  21. Rodríguez-Castellanos, W.; Martínez-Bustos, F.; Rodrigue, D.; Trujillo-Barragán, M. Extrusion blow molding of a starch–gelatin polymer matrix reinforced with cellulose. Eur. Polym. J. 2015, 73, 335–343. [Google Scholar] [CrossRef]
  22. Gumul, D.; Krystyjan, M.; Buksa, K.; Ziobro, R.; Zięba, T. The influence of oxidation, extrusion and oxidation/extrusion on physico-chemical properties of potato starch. Starch Stärke 2014, 66, 190–198. [Google Scholar] [CrossRef]
  23. Sapper, M.; Talens, P.; Chiralt, A. Improving Functional Properties of Cassava Starch-Based Films by Incorporating Xanthan, Gellan, or Pullulan Gums. Int. J. Polym. Sci. 2019, 2019, 5367164. [Google Scholar] [CrossRef] [Green Version]
  24. Ribba, L.; Garcia, N.L.; D’Accorso, N.; Goyanes, S. Disadvantages of Starch-Based Materials, Feasible Alternatives in Order to Overcome These Limitations. In Starch-Based Materials in Food Packaging; Elsevier: Amsterdam, The Netherlands, 2017; pp. 37–76. ISBN 9780128122570. [Google Scholar]
  25. Liu, H.; Xie, F.; Yu, L.; Chen, L.; Li, L. Thermal processing of starch-based polymers. Prog. Polym. Sci. 2009, 34, 1348–1368. [Google Scholar] [CrossRef]
  26. Vieira, M.G.A.; da Silva, M.A.; dos Santos, L.O.; Beppu, M.M. Natural-based plasticizers and biopolymer films: A review. Eur. Polym. J. 2011, 47, 254–263. [Google Scholar] [CrossRef] [Green Version]
  27. Grifffin, G.J.L. Biodegradable Fillers in Thermoplastics. In Advances in Chemistry; Deanin, R.D., Schott, N.R., Eds.; American Chemical Society: Washington, DC, USA, 1974; pp. 159–170. [Google Scholar]
  28. Tian, H.; Xu, G. Processing and Characterization of Glycerol-Plasticized Soy Protein Plastics Reinforced with Citric Acid-Modified Starch Nanoparticles. J. Polym. Environ. 2011, 19, 582–588. [Google Scholar] [CrossRef]
  29. Hochschule Hannover–University of Applied Sciences and Arts IfBB–Institute for Bioplastics and Biocomposites. Available online: https://www.ifbb-hannover.de/en/facts-and-statistics.html (accessed on 12 May 2020).
  30. Spierling, S.; Knüpffer, E.; Behnsen, H.; Mudersbach, M.; Krieg, H.; Springer, S.; Albrecht, S.; Herrmann, C.; Endres, H.J. Bio-based plastics-A review of environmental, social and economic impact assessments. J. Clean. Prod. 2018, 185, 476–491. [Google Scholar] [CrossRef]
  31. Reis, M.O.; Olivato, J.B.; Bilck, A.P.; Zanela, J.; Grossmann, M.V.E.; Yamashita, F. Biodegradable trays of thermoplastic starch/poly (lactic acid) coated with beeswax. Ind. Crops Prod. 2018, 112, 481–487. [Google Scholar] [CrossRef]
  32. Chollakup, R.; Pongburoos, S.; Boonsong, W.; Khanoonkon, N.; Kongsin, K.; Sothornvit, R.; Sukyai, P.; Sukatta, U.; Harnkarnsujarit, N. Antioxidant and antibacterial activities of cassava starch and whey protein blend films containing rambutan peel extract and cinnamon oil for active packaging. LWT 2020, 130, 109573. [Google Scholar] [CrossRef]
  33. George, J.; Sabapathi, S.N. Cellulose nanocrystals: Synthesis, functional properties, and applications. Nanotechnol. Sci. Appl. 2015, 8, 45. [Google Scholar] [CrossRef] [Green Version]
  34. Rajinipriya, M.; Nagalakshmaiah, M.; Robert, M.; Elkoun, S. Importance of Agricultural and Industrial Waste in the Field of Nanocellulose and Recent Industrial Developments of Wood Based Nanocellulose: A Review. ACS Sustain. Chem. Eng. 2018, 6, 2807–2828. [Google Scholar] [CrossRef]
  35. Lindman, B.; Medronho, B.; Alves, L.; Costa, C.; Edlund, H.; Norgren, M. The relevance of structural features of cellulose and its interactions to dissolution, regeneration, gelation and plasticization phenomena. Phys. Chem. Chem. Phys. 2017, 19, 23704–23718. [Google Scholar] [CrossRef] [Green Version]
  36. Mallick, P.K. Thermoplastics and thermoplastic–matrix composites for lightweight automotive structures. In Materials, Design and Manufacturing for Lightweight Vehicles; Elsevier: Amsterdam, The Netherlands, 2010; pp. 174–207. [Google Scholar]
  37. Peelman, N.; Ragaert, P.; De Meulenaer, B.; Adons, D.; Peeters, R.; Cardon, L.; Van Impe, F.; Devlieghere, F. Application of bioplastics for food packaging. Trends Food Sci. Technol. 2013, 32, 128–141. [Google Scholar] [CrossRef] [Green Version]
  38. Babu, R.P.; O’Connor, K.; Seeram, R. Current progress on bio-based polymers and their future trends. Prog. Biomater. 2013, 2, 8. [Google Scholar] [CrossRef] [Green Version]
  39. Fortunati, E.; Armentano, I.; Zhou, Q.; Iannoni, A.; Saino, E.; Visai, L.; Berglund, L.A.; Kenny, J.M. Multifunctional bionanocomposite films of poly(lactic acid), cellulose nanocrystals and silver nanoparticles. Carbohydr. Polym. 2012, 87, 1596–1605. [Google Scholar] [CrossRef]
  40. Jiang, S.; Zhang, T.; Song, Y.; Qian, F.; Tuo, Y.; Mu, G. Mechanical properties of whey protein concentrate based film improved by the coexistence of nanocrystalline cellulose and transglutaminase. Int. J. Biol. Macromol. 2019, 126, 1266–1272. [Google Scholar] [CrossRef] [PubMed]
  41. Mao, J.; Tang, Y.; Zhao, R.; Zhou, Y.; Wang, Z. Preparation of nanofibrillated cellulose and application in reinforced PLA/starch nanocomposite film. J. Polym. Environ. 2019, 27, 728–738. [Google Scholar] [CrossRef]
  42. Bergamonti, L.; Potenza, M.; Haghighi Poshtiri, A.; Lorenzi, A.; Sanangelantoni, A.M.; Lazzarini, L.; Lottici, P.P.; Graiff, C. Ag-functionalized nanocrystalline cellulose for paper preservation and strengthening. Carbohydr. Polym. 2020, 231, 115773. [Google Scholar] [CrossRef] [PubMed]
  43. Jiang, G.; Zhang, M.; Feng, J.; Zhang, S.; Wang, X. High Oxygen Barrier Property of Poly(propylene carbonate)/Polyethylene Glycol Nanocomposites with Low Loading of Cellulose Nanocrytals. ACS Sustain. Chem. Eng. 2017, 5, 11246–11254. [Google Scholar] [CrossRef]
  44. Dreyfuss-Deseigne, R. Nanocellulose Films in Art Conservation. J. Pap. Conserv. 2017, 18, 18–29. [Google Scholar] [CrossRef]
  45. Alavi, M. Modifications of microcrystalline cellulose (MCC), nanofibrillated cellulose (NFC), and nanocrystalline cellulose (NCC) for antimicrobial and wound healing applications. e-Polymers 2019, 19, 103–119. [Google Scholar] [CrossRef]
  46. Hedenqvist, M.S.; Plackett, D. V Properties of plasticized composite films prepared from nanofibrillated cellulose and birch wood xylan. Cellulose 2015, 19, 2015–2031. [Google Scholar] [CrossRef]
  47. Cherpinski, A.; Torres-Giner, S.; Vartiainen, J.; Peresin, M.S.; Lahtinen, P.; Lagaron, J.M. Improving the water resistance of nanocellulose-based films with polyhydroxyalkanoates processed by the electrospinning coating technique. Cellulose 2018, 25, 1291–1307. [Google Scholar]
  48. Dehnad, D.; Mirzaei, H.; Emam-Djomeh, Z.; Jafari, S.M.; Dadashi, S. Thermal and antimicrobial properties of chitosan-nanocellulose films for extending shelf life of ground meat. Carbohydr. Polym. 2014, 109, 148–154. [Google Scholar]
  49. Moreirinha, C.; Vilela, C.; Silva, N.H.C.S.; Pinto, R.R.J.; Almeida, A.; Rocha, M.A.M.; Coelho, E.; Coimbra, M.A.; Silvestre, A.A.J.; Freire, C.S.R. Antioxidant and antimicrobial films based on brewers spent grain arabinoxylans, nanocellulose and feruloylated compounds for active packaging. Food Hydrocoll. 2020, 108, 105836. [Google Scholar] [CrossRef]
  50. Cheng, S.; Zhang, Y.; Cha, R.; Yang, J.; Jiang, X. Water-soluble nanocrystalline cellulose films with highly transparent and oxygen barrier properties. Nanoscale 2016, 8, 973–978. [Google Scholar] [CrossRef]
  51. Missio, A.L.; Mattos, B.D.; Ferreira, D.d.F.; Magalhães, W.L.E.; Bertuol, D.A.; Gatto, D.A.; Petutschnigg, A.; Tondi, G. Nanocellulose-tannin films: From trees to sustainable active packaging. J. Clean. Prod. 2018, 184, 143–151. [Google Scholar] [CrossRef]
  52. Tyagi, P.; Lucia, L.A.; Hubbe, M.A.; Pal, L. Nanocellulose-based multilayer barrier coatings for gas, oil, and grease resistance. Carbohydr. Polym. 2019, 206, 281–288. [Google Scholar] [CrossRef]
  53. Pourmoazzen, Z.; Sadeghifar, H.; Chen, J.; Yang, G.; Zhang, K.; Lucia, L. The morphology, self-assembly, and host-guest properties of cellulose nanocrystals surface grafted with cholesterol. Carbohydr. Polym. 2020, 233, 115840. [Google Scholar] [CrossRef]
  54. Bian, H.; Gao, Y.; Wang, R.; Liu, Z.; Wu, W.; Dai, H. Contribution of lignin to the surface structure and physical performance of cellulose nanofibrils film. Cellulose 2018, 25, 1309–1318. [Google Scholar] [CrossRef]
  55. Aider, M. Chitosan application for active bio-based films production and potential in the food industry: Review. LWT Food Sci. Technol. 2010, 43, 837–842. [Google Scholar] [CrossRef]
  56. Ravi Kumar, M.N. A review of chitin and chitosan applications. React. Funct. Polym. 2000, 46, 1–27. [Google Scholar] [CrossRef]
  57. Francis Suh, J.-K.; Matthew, H.W. Application of chitosan-based polysaccharide biomaterials in cartilage tissue engineering: A review. Biomaterials 2000, 21, 2589–2598. [Google Scholar] [CrossRef]
  58. Mano, J.F. Viscoelastic Properties of Chitosan with Different Hydration Degrees as Studied by Dynamic Mechanical Analysis. Macromol. Biosci. 2008, 8, 69–76. [Google Scholar] [CrossRef]
  59. Liu, X.; Xu, Y.; Zhan, X.; Xie, W.; Yang, X.; Cui, S.W.; Xia, W. Development and properties of new kojic acid and chitosan composite biodegradable films for active packaging materials. Int. J. Biol. Macromol. 2020, 144, 483–490. [Google Scholar] [CrossRef] [PubMed]
  60. Liu, Y.; Yuan, Y.; Duan, S.; Li, C.; Hu, B.; Liu, A.; Wu, D.; Cui, H.; Lin, L.; He, J.; et al. Preparation and characterization of chitosan films with three kinds of molecular weight for food packaging. Int. J. Biol. Macromol. 2020, 155, 249–259. [Google Scholar] [CrossRef] [PubMed]
  61. Prata, A.S.; Grosso, C.R.F. Production of microparticles with gelatin and chitosan. Carbohydr. Polym. 2015, 116, 292–299. [Google Scholar] [CrossRef] [PubMed]
  62. Botelho da Silva, S.; Krolicka, M.; van den Broek, L.A.M.; Frissen, A.E.; Boeriu, C.G. Water-soluble chitosan derivatives and pH-responsive hydrogels by selective C-6 oxidation mediated by TEMPO-laccase redox system. Carbohydr. Polym. 2018, 186, 299–309. [Google Scholar] [CrossRef]
  63. Jayakumar, R.; Prabaharan, M.; Reis, R.L.; Mano, J.F. Graft copolymerized chitosan—Present status and applications. Carbohydr. Polym. 2005, 62, 142–158. [Google Scholar] [CrossRef] [Green Version]
  64. Braz, E.M.A.; Silva, S.C.C.C.; Sousa Brito, C.A.R.; Brito, L.M.; Barreto, H.M.; Carvalho, F.A.A.; Santos, L.S.; Lobo, A.O.; Osajima, J.A.; Sousa, K.S.; et al. Spectroscopic, thermal characterizations and bacteria inhibition of chemically modified chitosan with phthalic anhydride. Mater. Chem. Phys. 2020, 240, 122053. [Google Scholar] [CrossRef]
  65. Prabaharan, M.; Mano, J.F. Stimuli-Responsive Hydrogels Based on Polysaccharides Incorporated with Thermo-Responsive Polymers as Novel Biomaterials. Macromol. Biosci. 2006, 6, 991–1008. [Google Scholar] [CrossRef]
  66. Castillejo, M.; Rebollar, E.; Oujja, M.; Sanz, M.; Selimis, A.; Sigletou, M.; Psycharakis, S.; Ranella, A.; Fotakis, C. Fabrication of porous biopolymer substrates for cell growth by UV laser: The role of pulse duration. Appl. Surf. Sci. 2012, 258, 8919–8927. [Google Scholar] [CrossRef] [Green Version]
  67. Chan, S.W.; Mirhosseini, H.; Taip, F.S.; Ling, T.C.; Tan, C.P. Comparative study on the physicochemical properties of κ-carrageenan extracted from Kappaphycus alvarezii (doty) doty ex Silva in Tawau, Sabah, Malaysia and commercial κ-carrageenans. Food Hydrocoll. 2013, 30, 581–588. [Google Scholar] [CrossRef]
  68. Pu, S.; Li, J.; Sun, L.; Zhong, L.; Ma, Q. An in vitro comparison of the antioxidant activities of chitosan and green synthesized gold nanoparticles. Carbohydr. Polym. 2019, 211, 161–172. [Google Scholar] [CrossRef]
  69. Darmadji, P.; Izumimoto, M. Effect of chitosan in meat preservation. Meat Sci. 1994, 38, 243–254. [Google Scholar] [CrossRef]
  70. Ouattara, B.; Simard, R.E.; Piette, G.; Begin, A.; Holley, R.A. Diffusion of Acetic and Propionic Acids from Chitosan-based Antimicrobial Packaging Films. J. Food Sci. 2000, 65, 768–773. [Google Scholar] [CrossRef]
  71. Jost, V.; Kobsik, K.; Schmid, M.; Noller, K. Influence of plasticiser on the barrier, mechanical and grease resistance properties of alginate cast films. Carbohydr. Polym. 2014, 110, 309–319. [Google Scholar] [CrossRef]
  72. Lim, H.N.; Huang, N.M.; Loo, C.H. Facile preparation of graphene-based chitosan films: Enhanced thermal, mechanical and antibacterial properties. J. Non. Cryst. Solids 2012, 358, 525–530. [Google Scholar] [CrossRef]
  73. Carson, L.; Kelly-Brown, C.; Stewart, M.; Oki, A.; Regisford, G.; Luo, Z.; Bakhmutov, V.I. Synthesis and characterization of chitosan–carbon nanotube composites. Mater. Lett. 2009, 63, 617–620. [Google Scholar] [CrossRef] [Green Version]
  74. Draget, K.I.; Taylor, C. Chemical, physical and biological properties of alginates and their biomedical implications. Food Hydrocoll. 2011, 25, 251–256. [Google Scholar] [CrossRef]
  75. Lee, K.Y.; Mooney, D.J. Alginate: Properties and biomedical applications. Prog. Polym. Sci. 2012, 37, 106–126. [Google Scholar] [CrossRef] [Green Version]
  76. Senturk Parreidt, T.; Müller, K.; Schmid, M. Alginate-Based Edible Films and Coatings for Food Packaging Applications. Foods 2018, 7, 170. [Google Scholar] [CrossRef] [Green Version]
  77. Marcos, B.; Aymerich, T.; Monfort, J.M.; Garriga, M. Use of antimicrobial biodegradable packaging to control Listeria monocytogenes during storage of cooked ham. Int. J. Food Microbiol. 2007, 120, 152–158. [Google Scholar] [CrossRef] [Green Version]
  78. Jiang, T. Effect of alginate coating on physicochemical and sensory qualities of button mushrooms (Agaricus bisporus) under a high oxygen modified atmosphere. Postharvest Biol. Technol. 2013, 76, 91–97. [Google Scholar] [CrossRef]
  79. Rhim, J.W.; Wu, Y.; Weller, C.L.; Schnepf, M. Physical Characteristics of a Composite Film of Soy Protein Isolate and Propyleneglycol Alginate. J. Food Sci. 1999, 64, 149–152. [Google Scholar] [CrossRef]
  80. Gu, C.H.; Wang, J.J.; Yu, Y.; Sun, H.; Shuai, N.; Wei, B. Biodegradable multilayer barrier films based on alginate/polyethyleneimine and biaxially oriented poly(lactic acid). Carbohydr. Polym. 2013, 92, 1579–1585. [Google Scholar] [CrossRef] [PubMed]
  81. Mohammed Fayaz, A.; Balaji, K.; Girilal, M.; Kalaichelvan, P.T.; Venkatesan, R. Mycobased Synthesis of Silver Nanoparticles and Their Incorporation into Sodium Alginate Films for Vegetable and Fruit Preservation. J. Agric. Food Chem. 2009, 57, 6246–6252. [Google Scholar] [CrossRef] [PubMed]
  82. Cutter, C.N.; Siragusa, G.R. Reduction of Brochothrix thermosphacta on beef surfaces following immobilization of nisin in calcium alginate gels. Lett. Appl. Microbiol. 1996, 23, 9–12. [Google Scholar] [CrossRef] [Green Version]
  83. Priyadarshi, R.; Kim, H.J.; Rhim, J.W. Effect of sulfur nanoparticles on properties of alginate-based films for active food packaging applications. Food Hydrocoll. 2021, 110, 106155. [Google Scholar] [CrossRef]
  84. Pawar, S.G.; Pathade, G.R.; Rale, V.B. Pullulan production from cane molasses by Aureobasidium mausonii strain NCIM 1226. Anal. Chem. 2007, 6, 4–8. [Google Scholar]
  85. Wu, S.; Chen, H.; Jin, Z.; Tong, Q. Effect of two-stage temperature on pullulan production by Aureobasidium pullulans. World J. Microbiol. Biotechnol. 2010, 26, 737–741. [Google Scholar] [CrossRef]
  86. Ma, Z.-C.; Liu, N.-N.; Chi, Z.; Liu, G.-L.; Chi, Z.-M. Genetic Modification of the Marine-Isolated Yeast Aureobasidium melanogenum P16 for Efficient Pullulan Production from Inulin. Mar. Biotechnol. 2015, 17, 511–522. [Google Scholar] [CrossRef]
  87. Yurlova, N.A.; De Hoog, G.S. A new variety of Aureobasidium pullulans characterized by exopolysaccharide structure, nutritional physiology and molecular features. Antonie van Leeuwenhoek Int. J. Gen. Mol. Microbiol. 1997, 72, 141–147. [Google Scholar] [CrossRef]
  88. Liu, Y.; Liu, Y.; Han, K.; Cai, Y.; Ma, M.; Tong, Q.; Sheng, L. Effect of nano-TiO2 on the physical, mechanical and optical properties of pullulan film. Carbohydr. Polym. 2019, 218, 95–102. [Google Scholar] [CrossRef]
  89. Sheng, L.; Su, P.; Han, K.; Chen, J.; Cao, A.; Zhang, Z.; Jin, Y.; Ma, M. Synthesis and structural characterization of lysozyme–pullulan conjugates obtained by the Maillard reaction. Food Hydrocoll. 2017, 71, 1–7. [Google Scholar] [CrossRef]
  90. 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] [PubMed]
  91. Saeaeh, K.; Thummarungsan, N.; Paradee, N.; Choeichom, P.; Phasuksom, K.; Lerdwijitjarud, W.; Sirivat, A. Soft and highly responsive multi-walled carbon nanotube/pullulan hydrogel composites as electroactive materials. Eur. Polym. J. 2019, 120, 109231. [Google Scholar] [CrossRef]
  92. Brown, M.L. A Comparative Study of the Effects of Non-starch Polysaccharide Gums on Physical Properties of Single-screw Extruded Aquafeed. J. Food Process. Technol. 2015, 6, 457. [Google Scholar] [CrossRef]
  93. Sarraf, A.G.; Tissot, H.; Tissot, P.; Alfonso, D.; Gurny, R.; Doelker, E. Influence of hot-melt extrusion and compression molding on polymer structure organization, investigated by differential scanning calorimetry. J. Appl. Polym. Sci. 2001, 81, 3124–3132. [Google Scholar] [CrossRef]
  94. Ganduri, V.S.R. Evaluation of pullulan-based edible active coating methods on Rastali and Chakkarakeli bananas and their shelf-life extension parameters studies. J. Food Process. Preserv. 2020, 44, e14378. [Google Scholar] [CrossRef]
  95. Chu, Y.; Gao, C.; Liu, X.; Zhang, N.; Xu, T.; Feng, X.; Yang, Y.; Shen, X.; Tang, X. Improvement of storage quality of strawberries by pullulan coatings incorporated with cinnamon essential oil nanoemulsion. LWT 2020, 122, 109054. [Google Scholar] [CrossRef]
  96. Zhang, L.; Huang, C.; Zhao, H. Application of Pullulan and Chitosan Multilayer Coatings in Fresh Papayas. Coatings 2019, 9, 745. [Google Scholar] [CrossRef] [Green Version]
  97. Treviño-Garza, M.Z.; García, S.; Heredia, N.; Alanís-Guzmán, M.G.; Arévalo-Niño, K. Layer-by-layer edible coatings based on mucilages, pullulan and chitosan and its effect on quality and preservation of fresh-cut pineapple (Ananas comosus). Postharvest Biol. Technol. 2017, 128, 63–75. [Google Scholar] [CrossRef]
  98. Shao, P.; Niu, B.; Chen, H.; Sun, P. Fabrication and characterization of tea polyphenols loaded pullulan-CMC electrospun nanofiber for fruit preservation. Int. J. Biol. Macromol. 2018, 107, 1908–1914. [Google Scholar] [CrossRef]
  99. Qian, Y.F.; Zheng, L.J.; Song, R.Y.; Du, B. Electrospinning of Pullulan Nanofibers for Food Package Materials. Adv. Mater. Res. 2013, 821–822, 1321–1325. [Google Scholar] [CrossRef]
  100. Son, T.-W.; Lee, G.-M.; Lee, D.-W.; Lee, J.-H.; Lim, H.-S. Preparation and Characterization of Electrospun Pullulan Webs. Polym. Korea 2012, 36, 196–201. [Google Scholar] [CrossRef]
  101. Wu, J.; Zhong, F.; Li, Y.; Shoemaker, C.F.; Xia, W. Preparation and characterization of pullulan-chitosan and pullulan-carboxymethyl chitosan blended films. Food Hydrocoll. 2013, 30, 82–91. [Google Scholar] [CrossRef]
  102. Kanmani, P.; Lim, S.T. Development and characterization of novel probiotic-residing pullulan/starch edible films. Food Chem. 2013, 141, 1041–1049. [Google Scholar] [CrossRef] [PubMed]
  103. Pinto, R.J.B.; Almeida, A.; Fernandes, S.C.M.; Freire, C.S.R.; Silvestre, A.J.D.; Neto, C.P.; Trindade, T. Antifungal activity of transparent nanocomposite thin films of pullulan and silver against Aspergillus niger. Colloids Surf. B Biointerfaces 2013, 103, 143–148. [Google Scholar] [CrossRef]
  104. Kristo, E.; Biliaderis, C.G. Physical properties of starch nanocrystal-reinforced pullulan films. Carbohydr. Polym. 2007, 68, 146–158. [Google Scholar] [CrossRef]
  105. Lee, J.H.; Jeong, D.; Kanmani, P. Study on physical and mechanical properties of the biopolymer/silver based active nanocomposite films with antimicrobial activity. Carbohydr. Polym. 2019, 224, 115159. [Google Scholar] [CrossRef]
  106. Ramos, M.; Valdés, A.; Beltrán, A.; Garrigós, M. Gelatin-Based Films and Coatings for Food Packaging Applications. Coatings 2016, 6, 41. [Google Scholar] [CrossRef] [Green Version]
  107. Nur Hanani, Z.A.; Roos, Y.H.; Kerry, J.P. Use and application of gelatin as potential biodegradable packaging materials for food products. Int. J. Biol. Macromol. 2014, 71, 94–102. [Google Scholar] [CrossRef]
  108. Papon, P.; Leblond, J.; Meijer, P.H. (Eds.) Gelation and Transitions in Biopolymers. In The Physics of Phase Transitions; Springer: Berlin/Heidelberg, Germany, 2006; pp. 189–213. [Google Scholar]
  109. Shankar, S.; Jaiswal, L.; Rhim, J.-W. Gelatin-Based Nanocomposite Films. In Antimicrobial Food Packaging; Elsevier: Amsterdam, The Netherlands, 2016; pp. 339–348. [Google Scholar]
  110. Ortiz-Zarama, M.A.; Jiménez-Aparicio, A.R.; Solorza-Feria, J. Obtainment and partial characterization of biodegradable gelatin films with tannic acid, bentonite and glycerol. J. Sci. Food Agric. 2016, 96, 3424–3431. [Google Scholar] [CrossRef]
  111. Mellinas, C.; Valdés, A.; Ramos, M.; Burgos, N.; Garrigós, M.d.C.; Jiménez, A. Active edible films: Current state and future trends. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef] [Green Version]
  112. Tongdeesoontorn, W.; Mauer, L.J.; Wongruong, S.; Sriburi, P.; Rachtanapun, P. Mechanical and Physical Properties of Cassava Starch-Gelatin Composite Films. Int. J. Polym. Mater. 2012, 61, 778–792. [Google Scholar] [CrossRef]
  113. Malherbi, N.M.; Schmitz, A.C.; Grando, R.C.; Bilck, A.P.; Yamashita, F.; Tormen, L.; Fakhouri, F.M.; Velasco, J.I.; Bertan, L.C. Corn starch and gelatin-based films added with guabiroba pulp for application in food packaging. Food Packag. Shelf Life 2019, 19, 140–146. [Google Scholar] [CrossRef]
  114. 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]
  115. Zhao, J.; Wei, F.; Xu, W.; Han, X. Enhanced antibacterial performance of gelatin/chitosan film containing capsaicin loaded MOFs for food packaging. Appl. Surf. Sci. 2020, 510, 145418. [Google Scholar] [CrossRef]
  116. Halder, D.; Mitra, A.; Bag, S.; Raychaudhuri, U.; Chakraborty, R. Study on Gelatin-Silver Nanoparticle Composite Towards the Development of Bio-Based Antimicrobial Film. J. Nanosci. Nanotechnol. 2011, 11, 10374–10378. [Google Scholar] [CrossRef] [PubMed]
  117. Fakhouri, F.M.; Martelli, S.M.; Caon, T.; Velasco, J.I.; Mei, L.H.I. Edible films and coatings based on starch/gelatin: Film properties and effect of coatings on quality of refrigerated Red Crimson grapes. Postharvest Biol. Technol. 2015, 109, 57–64. [Google Scholar] [CrossRef]
  118. Jamróz, E.; Kopel, P.; Juszczak, L.; Kawecka, A.; Bytesnikova, Z.; Milosavljevic, V.; Makarewicz, M. Development of furcellaran-gelatin films with Se-AgNPs as an active packaging system for extension of mini kiwi shelf life. Food Packag. Shelf Life 2019, 21, 100339. [Google Scholar] [CrossRef]
  119. Samsi, M.S.; Kamari, A.; Din, S.M.; Lazar, G. Synthesis, characterization and application of gelatin–carboxymethyl cellulose blend films for preservation of cherry tomatoes and grapes. J. Food Sci. Technol. 2019, 56, 3099–3108. [Google Scholar] [CrossRef] [PubMed]
  120. Kingwascharapong, P.; Arisa, K.; Karnjanapratum, S.; Tanaka, F.; Tanaka, F. Effect of gelatin-based coating containing frog skin oil on the quality of persimmon and its characteristics. Sci. Hortic. Amst. 2020, 260, 108864. [Google Scholar] [CrossRef]
  121. Pellá, M.C.G.; Silva, O.A.; Pellá, M.G.; Beneton, A.G.; Caetano, J.; Simões, M.R.; Dragunski, D.C. Effect of gelatin and casein additions on starch edible biodegradable films for fruit surface coating. Food Chem. 2020, 309, 125764. [Google Scholar] [CrossRef]
  122. Sagis, L.M.C. Microencapsulation and Microspheres for Food Applications; Elsevier: Amsterdam, The Netherlands, 2015; ISBN 9780128003503. [Google Scholar]
  123. Fabra, M.J.; Lopez-Rubio, A.; Lagaron, J.M. High barrier polyhydroxyalcanoate food packaging film by means of nanostructured electrospun interlayers of zein. Food Hydrocoll. 2013, 32, 106–114. [Google Scholar] [CrossRef]
  124. Shukla, R.; Cheryan, M. Zein: The industrial protein from corn. Ind. Crops Prod. 2001, 13, 171–192. [Google Scholar] [CrossRef]
  125. Ahammed, S.; Liu, F.; Khin, M.N.; Yokoyama, W.H.; Zhong, F. Improvement of the water resistance and ductility of gelatin film by zein. Food Hydrocoll. 2020, 105, 105804. [Google Scholar] [CrossRef]
  126. Neo, Y.P.; Ray, S.; Jin, J.; Gizdavic-Nikolaidis, M.; Nieuwoudt, M.K.; Liu, D.; Quek, S.Y. Encapsulation of food grade antioxidant in natural biopolymer by electrospinning technique: A physicochemical study based on zein–gallic acid system. Food Chem. 2013, 136, 1013–1021. [Google Scholar] [CrossRef] [PubMed]
  127. Lawton, J.W. Plasticizers for Zein: Their Effect on Tensile Properties and Water Absorption of Zein Films. Cereal Chem. J. 2004, 81, 1–5. [Google Scholar] [CrossRef] [Green Version]
  128. Lai, H.-M.; Padua, G.W. Water Vapor Barrier Properties of Zein Films Plasticized with Oleic Acid. Cereal Chem. J. 1998, 75, 194–199. [Google Scholar] [CrossRef]
  129. Doğan Atik, İ.; Özen, B.; Tıhmınlıoğlu, F. Water vapour barrier performance of corn-zein coated polypropylene (PP) packaging films. J. Therm. Anal. Calorim. 2008, 94, 687–693. [Google Scholar] [CrossRef] [Green Version]
  130. Zhan, F.; Yan, X.; Sheng, F.; Li, B. Facile in situ synthesis of silver nanoparticles on tannic acid/zein electrospun membranes and their antibacterial, catalytic and antioxidant activities. Food Chem. 2020, 330, 127172. [Google Scholar] [CrossRef]
  131. Chen, G.; Chen, Y.; Jin, N.; Li, J.; Dong, S.; Li, S.; Zhang, Z.; Chen, Y. Zein films with porous polylactic acid coatings via cold plasma pre-treatment. Ind. Crops Prod. 2020, 150, 112382. [Google Scholar] [CrossRef]
  132. Park, H.-Y.; Kim, S.-J.; Kim, K.M.; You, Y.-S.; Kim, S.Y.; Han, J. Development of Antioxidant Packaging Material by Applying Corn-Zein to LLDPE Film in Combination with Phenolic Compounds. J. Food Sci. 2012, 77, E273–E279. [Google Scholar] [CrossRef] [PubMed]
  133. Aytac, Z.; Ipek, S.; Durgun, E.; Tekinay, T.; Uyar, T. Antibacterial electrospun zein nanofibrous web encapsulating thymol/cyclodextrin-inclusion complex for food packaging. Food Chem. 2017, 233, 117–124. [Google Scholar] [CrossRef] [PubMed]
  134. Alves de Oliveira, R.; Komesu, A.; Vaz Rossell, C.E.; Maciel Filho, R. Challenges and opportunities in lactic acid bioprocess design—From economic to production aspects. Biochem. Eng. J. 2018, 133, 219–239. [Google Scholar] [CrossRef]
  135. Suryanegara, L.; Nakagaito, A.N.; Yano, H. The effect of crystallization of PLA on the thermal and mechanical properties of microfibrillated cellulose-reinforced PLA composites. Compos. Sci. Technol. 2009, 69, 1187–1192. [Google Scholar] [CrossRef]
  136. Al-Tayyar, N.A.; Youssef, A.M.; Al-hindi, R. Antimicrobial food packaging based on sustainable Bio-based materials for reducing foodborne Pathogens: A review. Food Chem. 2020, 310, 125915. [Google Scholar] [CrossRef] [PubMed]
  137. Jamshidian, M.; Tehrany, E.A.; Imran, M.; Jacquot, M.; Desobry, S. Poly-Lactic Acid: Production, Applications, Nanocomposites, and Release Studies. Compr. Rev. Food Sci. Food Saf. 2010, 9, 552–571. [Google Scholar] [CrossRef]
  138. Mohsen, A.H.; Ali, N.A. Mechanical, Color and Barrier, Properties of Biodegradable Nanocomposites Polylactic Acid/Nanoclay. J. Bioremediation Biodegrad. 2018, 9, 455. [Google Scholar] [CrossRef]
  139. Gross, R.A. Biodegradable Polymers for the Environment. Science 2002, 297, 803–807. [Google Scholar] [CrossRef] [Green Version]
  140. Corneillie, S.; Smet, M. PLA architectures: The role of branching. Polym. Chem. 2015, 6, 850–867. [Google Scholar] [CrossRef] [Green Version]
  141. Esmaeili, M.; Pircheraghi, G.; Bagheri, R.; Altstädt, V. Poly(lactic acid)/coplasticized thermoplastic starch blend: Effect of plasticizer migration on rheological and mechanical properties. Polym. Adv. Technol. 2019, 30, 839–851. [Google Scholar] [CrossRef]
  142. Siakeng, R.; Jawaid, M.; Ariffin, H.; Sapuan, S.M. Mechanical, dynamic, and thermomechanical properties of coir/pineapple leaf fiber reinforced polylactic acid hybrid biocomposites. Polym. Compos. 2019, 40, 2000–2011. [Google Scholar] [CrossRef]
  143. Castro-Aguirre, E.; Iñiguez-Franco, F.; Samsudin, H.; Fang, X.; Auras, R. Poly(lactic acid)—Mass production, processing, industrial applications, and end of life. Adv. Drug Deliv. Rev. 2016, 107, 333–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Rivera-Briso, A.; Serrano-Aroca, Á. Poly(3-Hydroxybutyrate-co-3-Hydroxyvalerate): Enhancement Strategies for Advanced Applications. Polymers 2018, 10, 732. [Google Scholar] [CrossRef] [Green Version]
  145. Li, F.; Yu, H.; Wang, Y.; Zhou, Y.; Zhang, H.; Yao, J.; Abdalkarim, S.Y.H.; Tam, K.C. Natural Biodegradable Poly(3-hydroxybutyrate- co -3-hydroxyvalerate) Nanocomposites with Multifunctional Cellulose Nanocrystals/Graphene Oxide Hybrids for High-Performance Food Packaging. J. Agric. Food Chem. 2019, 67, 10954–10967. [Google Scholar] [CrossRef] [PubMed]
  146. Díez-Pascual, A.M.; Díez-Vicente, A.L. ZnO-Reinforced Poly(3-hydroxybutyrate- co -3-hydroxyvalerate) Bionanocomposites with Antimicrobial Function for Food Packaging. ACS Appl. Mater. Interfaces 2014, 6, 9822–9834. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Li, D.; Fu, J.; Ma, X. Improvement in thermal, mechanical, and barrier properties of biocomposite of poly (3-hydroxybutyrate-co-3-hydroxyhexanoate)/modified nano-SiO2. Polym. Compos. 2019, 41, 381–390. [Google Scholar] [CrossRef]
  148. Chen, G.X.; Hao, G.J.; Guo, T.Y.; Song, M.D.; Zhang, B.H. Structure and mechanical properties of poly (3-hydroxybutyrate- co-3-hydroxyvalerate) (PHBV)/clay nanocomposites. J. Mater. Sci. Lett. 2002, 21, 1587–1589. [Google Scholar] [CrossRef]
  149. Li, D.; Zhou, J.; Ma, X.; Li, J. Synthesis of a novel biocomposite of poly (3-hydroxybutyrate-co-3-hydroxyhexanoate) reinforced with acetylated cellulose nanocrystals. Cellulose 2019, 26, 8729–8743. [Google Scholar] [CrossRef]
  150. Vilela, C.; Kurek, M.; Hayouka, Z.; Röcker, B.; Yildirim, S.; Antunes, M.D.C.; Nilsen-Nygaard, J.; Pettersen, M.K.; Freire, C.S.R. A concise guide to active agents for active food packaging. Trends Food Sci. Technol. 2018, 80, 212–222. [Google Scholar] [CrossRef]
  151. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  152. Lyn, F.H.; Peng, T.C.; Ruzniza, M.Z.; Hanani, Z.A.N. Effect of oxidation degrees of graphene oxide (GO) on the structure and physical properties of chitosan/GO composite films. Food Packag. Shelf Life 2019, 21, 100373. [Google Scholar] [CrossRef]
  153. Purkait, T.; Singh, G.; Singh, M.; Kumar, D.; Dey, R.S. Large area few-layer graphene with scalable preparation from waste biomass for high-performance supercapacitor. Sci. Rep. 2017, 7, 1–14. [Google Scholar] [CrossRef] [PubMed]
  154. Sun, P.Z.; Yang, Q.; Kuang, W.J.; Stebunov, Y.V.; Xiong, W.Q.; Yu, J.; Nair, R.R.; Katsnelson, M.I.; Yuan, S.J.; Grigorieva, I.V.; et al. Limits on gas impermeability of graphene. Nature 2020, 579, 229–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wassei, J.K.; Kaner, R.B. Graphene, a promising transparent conductor. Mater. Today 2010, 13, 52–59. [Google Scholar] [CrossRef]
  156. Shokrieh, M.M.; Esmkhani, M.; Shahverdi, H.R.; Vahedi, F. Effect of graphene nanosheets (GNS) and graphite nanoplatelets (GNP) on the mechanical properties of epoxy nanocomposites. Sci. Adv. Mater. 2013, 5, 260–266. [Google Scholar] [CrossRef]
  157. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and graphene oxide: Synthesis, properties, and applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  158. Goncalves, G.; Marques, P.A.A.P.; Granadeiro, C.M.; Nogueira, H.I.S.; Singh, M.K.; Grácio, J. Surface Modification of Graphene Nanosheets with Gold Nanoparticles: The Role of Oxygen Moieties at Graphene Surface on Gold Nucleation and Growth. Chem. Mater. 2009, 21, 4796–4802. [Google Scholar] [CrossRef]
  159. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar] [CrossRef]
  160. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  161. Compton, O.C.; Jain, B.; Dikin, D.A.; Abouimrane, A.; Amine, K.; Nguyen, S.T. Chemically active reduced graphene oxide with tunable C/O ratios. ACS Nano 2011, 5, 4380–4391. [Google Scholar] [CrossRef]
  162. Zhou, X.; Zhang, J.; Wu, H.; Yang, H.; Zhang, J.; Guo, S. Reducing Graphene Oxide via Hydroxylamine: A Simple and Efficient Route to Graphene. J. Phys. Chem. C 2011, 115, 11957–11961. [Google Scholar] [CrossRef]
  163. Chua, C.K.; Pumera, M. Chemical reduction of graphene oxide: A synthetic chemistry viewpoint. Chem. Soc. Rev. 2014, 43, 291–312. [Google Scholar] [CrossRef] [PubMed]
  164. Konios, D.; Stylianakis, M.M.; Stratakis, E.; Kymakis, E. Dispersion behaviour of graphene oxide and reduced graphene oxide. J. Colloid Interface Sci. 2014, 430, 108–112. [Google Scholar] [CrossRef] [PubMed]
  165. Wick, P.; Louw-Gaume, A.E.; Kucki, M.; Krug, H.F.; Kostarelos, K.; Fadeel, B.; Dawson, K.A.; Salvati, A.; Vázquez, E.; Ballerini, L.; et al. Classification Framework for Graphene-Based Materials. Angew. Chemie Int. Ed. 2014, 53, 7714–7718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Sun, X.; Sun, H.; Li, H.; Peng, H. Developing polymer composite materials: Carbon nanotubes or graphene? Adv. Mater. 2013, 25, 5153–5176. [Google Scholar] [CrossRef]
  167. Wang, P.; Tang, S.; Sheng, F.; Cai, J.; Fei, P.; Nawaz, A.; Walayat, N.; Bakhsh, A.; Xiong, H. Crystallization, thermal stability, barrier property, and aging resistance application of multi-functionalized graphene oxide/poly (lactide)/starch nanocomposites. Int. J. Biol. Macromol. 2019, 132, 1208–1220. [Google Scholar] [CrossRef] [PubMed]
  168. Liu, Y.; Wang, S.; Lan, W.; Qin, W. Fabrication of polylactic acid/carbon nanotubes/chitosan composite fibers by electrospinning for strawberry preservation. Int. J. Biol. Macromol. 2019, 121, 1329–1336. [Google Scholar] [CrossRef]
  169. Gouvêa, R.F.; Del Aguils, E.M.; Paschoalin, V.M.F.; Andrade, C.T. Extruded hybrids based on poly (3-hydroxybutyrate- co -3-hydroxyvalerate) and reduced graphene oxide composite for active food packaging. Food Packag. Shelf Life 2018, 16, 77–85. [Google Scholar] [CrossRef]
  170. Barra, A.; Ferreira, N.M.; Martins, M.A.; Lazar, O.; Pantazi, A.; Jderu, A.A.; Neumayer, S.M.; Rodriguez, B.J.; Enăchescu, M.; Ferreira, P.; et al. Eco-friendly preparation of electrically conductive chitosan-reduced graphene oxide flexible bionanocomposites for food packaging and biological applications. Compos. Sci. Technol. 2019, 173, 53–60. [Google Scholar] [CrossRef]
  171. Afshar, S.; Banisadi, H. Investigation the effect of graphene oxide and gelatin/starch weight ratio on the properties of starch/gelatin/GO nanocomposite films: The RSM study. Int. J. Biol. Macromol. 2017, 109, 1019–1028. [Google Scholar] [CrossRef]
  172. Yu, B.; Zhao, Z.; Fu, S.; Meng, L.; Liu, Y.; Chen, F.; Wang, K.; Fu, Q. Fabrication of PLA/CNC/CNT conductive composites for high electromagnetic interference shielding based on Pickering emulsions method. Compos. Part A Appl. Sci. Manuf. 2019, 125, 105558. [Google Scholar] [CrossRef]
  173. Montes, S.; Etxeberria, A.; Mocholi, V.; Rekondo, A.; Grande, H.; Labidi, J. Effect of combining cellulose nanocrystals and graphene nanoplatelets on the properties of poly (lactic acid) based films. Polym. Lett. 2018, 12, 543–555. [Google Scholar] [CrossRef]
  174. Xu, L.; Zhang, X.; Cui, C.; Ren, P.; Yan, D.; Li, Z. Enhanced Mechanical Performance of Segregated Carbon Nanotube / Poly (lactic acid) Composite for E fficient Electromagnetic Interference Shielding. Ind. Eng. Chem. Res. 2019, 58, 4454–4461. [Google Scholar] [CrossRef]
  175. Ren, F.; Li, Z.; Xu, L.; Sun, Z.; Ren, P.; Yan, D.; Li, Z. Large-scale preparation of segregated PLA/carbon nanotube composite with high efficient electromagnetic interference shielding and favourable mechanical properties. Compos. Part B Eng. 2018, 155, 405–413. [Google Scholar] [CrossRef]
  176. Husna, N.; Che, M.; Mustapha, M.; Sallih, N.; Eng, K.; Ahmad, R. Progress in Organic Coatings Graphene-based polymer nanocomposites as barrier coatings for corrosion protection. Prog. Org. Coatings 2019, 135, 82–99. [Google Scholar] [CrossRef]
  177. Sharma, B.; Malik, P.; Jain, P. Biopolymer reinforced nanocomposites: A comprehensive review. Mater. Today Commun. 2018, 16, 353–363. [Google Scholar] [CrossRef]
  178. Terzioglu, P.; Altin, Y.; Kalemtas, A.; Celik Bedeloglu, A. Graphene oxide and zinc oxide decorated chitosan nanocomposite biofilms for packaging applications. J. Polym. Eng. 2020, 40, 152–157. [Google Scholar] [CrossRef] [Green Version]
  179. Weng, S.X.; Youse, N.; Tufenkji, N. Self-Assembly of Ultralarge Graphene Oxide Nanosheets and Alginate into Layered Nanocomposites for Robust Packaging Materials. ACS Appl. Nano Mater. 2019, 2, 1431–1444. [Google Scholar] [CrossRef]
  180. Xie, D.; Liu, Q.; Xu, D.; Wu, X. Graphene oxide–polyoctahedral silsesquioxane–chitosan composite films with improved mechanical and water-vapor-transport properties. J. Appl. Polym. Sci. 2019, 47748, 1–9. [Google Scholar] [CrossRef]
  181. Chen, Q.; Shi, Y.; Chen, G.; Cai, M. Enhanced mechanical and hydrophobic properties of composite cassava starch films with stearic acid modified MCC (microcrystalline cellulose)/NCC (nanocellulose) as strength agent. Int. J. Biol. Macromol. 2020, 142, 846–854. [Google Scholar] [CrossRef]
  182. Mallakpour, S.; Rashidimoghadam, S. Application of ultrasonic irradiation as a benign method for production of glycerol plasticized-starch/ascorbic acid functionalized MWCNTs nanocomposites: Investigation of methylene blue adsorption and electrical properties. Ultrason. Sonochem. 2018, 40, 419–432. [Google Scholar] [CrossRef]
  183. Mergen, O.B.; Arda, E.; Evingur, G.A. Electrical, optical and mechanical properties of chitosan biocomposites. J. Compos. Mater. 2019, 54, 846–854. [Google Scholar] [CrossRef]
  184. Jia, S.; Wang, Z.; Chen, L. A feasible strategy to constructing hybrid conductive networks in PLA—Based composites modified by CNT-d-RGO particles and PEG for mechanical and electrical properties. Polym. Adv. Technol. 2019, 31, 699–712. [Google Scholar] [CrossRef]
  185. Idumah, C.I.; Hassan, A.; Ihuoma, D.E. Recently emerging trends in polymer nanocomposites packaging materials. Polym. Technol. Mater. 2019, 58, 1054–1109. [Google Scholar] [CrossRef]
  186. Xie, X.; Li, L.; Ye, S.; Zhang, Q.; Chen, X.; Huang, X. Photocatalytic degradation of ethylene by TiO2 nanotubes/ reduced graphene oxide prepared by gamma irradiation. Radiat. Phys. Chem. 2019, 165, 108371. [Google Scholar] [CrossRef]
  187. 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]
  188. Shen, X.J.; Yang, S.; Shen, J.X.; Ma, J.L.; Wu, Y.Q.; Zeng, X.L.; Fu, S.Y. Improved mechanical and antibacterial properties of silver-graphene oxide hybrid/polylactid acid composites by in-situ polymerization. Ind. Crops Prod. 2019, 130, 571–579. [Google Scholar] [CrossRef]
  189. Bher, A.; Unalan, I.U.; Auras, R.; Rubino, M.; Schvezov, C.E. Graphene modifies the biodegradation of poly(lactic acid)-thermoplastic cassava starch reactive blend films. Polym. Degrad. Stab. 2019, 164, 187–197. [Google Scholar] [CrossRef]
  190. Ferreira, W.H.; Dahmouche, K.; Andrade, C.T. Dispersion of reduced graphene oxide within thermoplastic starch / poly (lactic acid) blends investigated by small-angle X-ray scattering. Carbohydr. Polym. 2019, 208, 124–132. [Google Scholar] [CrossRef]
  191. De Carvalho, A.P.A.; Conte Junior, C.A. Green strategies for active food packagings: A systematic review on active properties of graphene-based nanomaterials and biodegradable polymers. Trends Food Sci. Technol. 2020, 103, 130–143. [Google Scholar] [CrossRef]
  192. Cui, R.; Jiang, K.; Yuan, M.; Cao, J.; Li, L.; Tang, Z.; Qin, Y. Antimicrobial film based on polylactic acid and carbon nanotube for controlled cinnamaldehyde release. J. Mater. Res. Technol. 2020, 9, 10130–10138. [Google Scholar] [CrossRef]
  193. Xie, J.; Huang, L.; Wang, R.; Ye, S.; Song, X. Novel visible light-responsive graphene oxide/Bi2WO6/starch composite membrane for efficient degradation of ethylene. Carbohydr. Polym. 2020, 246, 116640. [Google Scholar] [CrossRef] [PubMed]
  194. Darder, M.; Aranda, P.; Ruiz-Hitzky, E. Bionanocomposites: A new concept of ecological, bioinspired, and functional hybrid materials. Adv. Mater. 2007, 19, 1309–1319. [Google Scholar] [CrossRef]
  195. Ruiz-Hitzky, E.; Ariga, K.; Lvov, Y.M. Bio-inorganic Hybrid Nanomaterials: Strategies, Synthesis, Characterization and Applications; Wiley-VCH: Weinheim, Germany, 2008; ISBN 978-3-527-31718-9. [Google Scholar]
  196. Fernandes, F.M.; Darder, M.; Ruiz, A.I.; Aranda, P.; Ruiz-Hitzky, E. Gelatine-based bio-nanocomposites. In Nanocomposites with Biodegradable Polymers: Synthesis, Properties, and Future Perspectives; Mittal, V., Ed.; Oxford University Press: New York, NY, USA, 2011; p. 209. [Google Scholar]
  197. Alcântara, A.C.S.; Darder, M.; Aranda, P.; Ruiz-Hitzky, E. Zein-fibrous clays biohybrid materials. Eur. J. Inorg. Chem. 2012, 5216–5224. [Google Scholar] [CrossRef]
  198. Aranda, P.; Darder, M.; Fernandes, F.M. Clay Mineral–Polymer Nanocomposites. In Handbook of Clay Science. Part A: Fundamentals; Bergaya, F., Lagaly, G., Eds.; Elsevier: Amsterdam, The Netherlands; Oxford, UK, 2013; p. 721. [Google Scholar]
  199. Ruiz-Hitzky, E.; Darder, M.; Fernandes, F.M.; Wicklein, B.; Alcântara, A.C.S.; Aranda, P. Fibrous clays based bionanocomposites. Prog. Polym. Sci. 2013, 38, 1392–1414. [Google Scholar] [CrossRef] [Green Version]
  200. Alcântara, A.C.S.; Darder, M.; Aranda, P.; Ayral, A.; Ruiz-Hitzky, E. Bionanocomposites based on polysaccharides and fibrous clays for packaging applications. J. Appl. Polym. Sci. 2016, 133, 42362. [Google Scholar] [CrossRef]
  201. González, M.M.; Darder, M.; Aranda, P.; Akkari, M.; Huttel, Y.; Mayoral, A.; Bettini, J.; Ruiz-hitzky, E. Functional Hybrid Nanopaper by Assembling Nanofibers of Cellulose and Sepiolite. Adv. Funct. Mater. 2018, 28, 1–13. [Google Scholar] [CrossRef]
  202. Darder, M.; Aranda, P.; Ruiz-García, C.; Fernandes, F.M.; Ruiz-Hitzky, E. The Meeting Point of Carbonaceous Materials and Clays: Toward a New Generation of Functional Composites. Adv. Funct. Mater. 2017, 28, 1704323. [Google Scholar] [CrossRef]
  203. Ruiz-García, C.; Darder, M.; Aranda, P.; Ruiz-Hitzky, E. Toward a green way for the chemical production of supported graphenes using porous solids. J. Mater. Chem. A 2017, 2, 2009–2017. [Google Scholar] [CrossRef]
  204. Bandyopadhyaya, R.; Nativ-Roth, E.; Regev, O.; Yerushalmi-Rozen, R. Stabilization of Individual Carbon Nanotubes in Aqueous Solutions. Nano Lett. 2002, 2, 25–28. [Google Scholar] [CrossRef]
  205. Jiang, L.; Gao, L.; Sun, J. Production of aqueous colloidal dispersions of carbon nanotubes. J. Colloid Interface Sci. 2003, 260, 89–94. [Google Scholar] [CrossRef]
  206. Huang, Y.Y.; Terentjev, E.M. Dispersion of carbon nanotubes: Mixing, sonication, stabilization, and composite properties. Polymers 2012, 4, 275–295. [Google Scholar] [CrossRef] [Green Version]
  207. Fernandes, R.M.F.; Abreu, B.; Claro, B.; Buzaglo, M.; Regev, O.; Furó, I.; Marques, E.F. Dispersing Carbon Nanotubes with Ionic Surfactants under Controlled Conditions: Comparisons and Insight. Langmuir 2015, 31, 10955–10965. [Google Scholar] [CrossRef] [PubMed]
  208. Kharisov, B.I.; Kharissova, O.V.; Dimas, A.V. The dispersion, solubilization and stabilization in “solution” of single-walled carbon nanotubes. RSC Adv. 2016, 6, 68760–68787. [Google Scholar] [CrossRef]
  209. Fernandes, F.M.; Ruiz-Hitzky, E. Assembling nanotubes and nanofibres: Cooperativeness in sepiolite-carbon nanotube materials. Carbon N. Y. 2014, 72, 296–303. [Google Scholar] [CrossRef]
  210. Ruiz-Hitzky, E.; Sobral, M.M.C.; Gómez-Avilés, A.; Nunes, C.; Ruiz-García, C.; Ferreira, P.; Aranda, P. Clay-Graphene Nanoplatelets Functional Conducting Composites. Adv. Funct. Mater. 2016, 26, 7394–7405. [Google Scholar] [CrossRef]
  211. Pillet, F.; Formosa-Dague, C.; Baaziz, H.; Dague, E.; Rols, M.P. Cell wall as a target for bacteria inactivation by pulsed electric fields. Sci. Rep. 2016, 6, 19778. [Google Scholar] [CrossRef]
  212. Sanchez-Vega, R.; Elez-Martinez, P.; Martin-Belloso, O. Influence of high-intensity pulsed electric field processing parameters on antioxidant compounds of broccoli juice. Innov. Food Sci. Emerg. Technol. 2015, 29, 70–77. [Google Scholar] [CrossRef]
  213. Quintão-Teixeira, L.J.; Soliva-Fortuny, R.; Mota Ramos, A.; Martín-Belloso, O. Kinetics of peroxidase inactivation in carrot juice treated with pulsed electric fields. J. Food Sci. 2013, 78, E222–E228. [Google Scholar] [CrossRef] [PubMed]
  214. Huang, K.; Wang, J. Designs of pulsed electric fields treatment chambers for liquid foods pasteurization process: A review. J. Food Eng. 2009, 95, 227–239. [Google Scholar] [CrossRef]
  215. Roodenburg, B.; Morren, J.; Berg, H.E.; de Haan, S.W.H. Metal release in a stainless steel pulsed electric field (PEF) system Part II. The treatment of orange juice; related to legislation and treatment chamber lifetime. Innov. Food Sci. Emerg. Technol. 2005, 6, 337–345. [Google Scholar] [CrossRef]
  216. Roodenburg, B.; de Haan, S.W.H.; van Boxtel, L.B.J.; Hatt, V.; Wouters, P.C.; Coronel, P.; Ferreira, J.A. Conductive plastic film electrodes for Pulsed Electric Field (PEF) treatment-A proof of principle. Innov. Food Sci. Emerg. Technol. 2010, 11, 274–282. [Google Scholar] [CrossRef]
  217. Roodenburg, B.; De Haan, S.W.H.; Ferreira, J.A.; Coronel, P.; Wouters, P.C.; Hatt, V. Toward 6 log10 pulsed electric field inactivation with conductive plastic packaging material. J. Food Process Eng. 2013, 36, 77–86. [Google Scholar] [CrossRef]
  218. Wang, Z.; Zhao, S.; Kang, H.; Zhang, W.; Zhang, S.; Li, J. Mussel byssus-inspired engineering of synergistic nanointerfacial interactions as sacrificial bonds into carbon nanotube-reinforced soy protein/nanofibrillated cellulose nanocomposites: Versatile mechanical enhancement. Appl. Surf. Sci. 2018, 434, 1086–1100. [Google Scholar] [CrossRef]
  219. Guo, Y.; Zuo, X.; Xue, Y.; Tang, J.; Gouzman, M.; Fang, Y.; Zhou, Y.; Wang, L.; Yu, Y.; Rafailovich, M.H. Engineering thermally and electrically conductive biodegradable polymer nanocomposites. Compos. Part B Eng. 2020, 189, 107905. [Google Scholar] [CrossRef]
  220. Chen, K.; Zhang, H. Alginate/pectin aerogel microspheres for controlled release of proanthocyanidins. Int. J. Biol. Macromol. 2019, 136, 936–943. [Google Scholar] [CrossRef] [PubMed]
  221. Lo Dico, G.; Wicklein, B.; Lisuzzo, L.; Lazzara, G.; Aranda, P.; Ruiz-Hitzky, E. Multicomponent bionanocomposites based on clay nanoarchitectures for electrochemical devices. Beilstein J. Nanotechnol. 2019, 10, 1303–1315. [Google Scholar] [CrossRef]
  222. Wang, X.-Y.; Narita, A.; Müllen, K. Precision synthesis versus bulk-scale fabrication of graphenes. Nat. Rev. Chem. 2018, 2, 1–10. [Google Scholar] [CrossRef]
  223. Phiri, J.; Johansson, L.; Gane, P.; Maloney, T.C. Co-exfoliation and fabrication of graphene based microfibrillated cellulose composites–mechanical and thermal stability and functional conductive properties. Nano Res. 2018, 10, 9569–9582. [Google Scholar] [CrossRef] [Green Version]
  224. Cobos, M.; González, B.; Fernández, M.J.; Fernández, M.D. International Journal of Biological Macromolecules Study on the effect of graphene and glycerol plasticizer on the properties of chitosan-graphene nanocomposites via in situ green chemical reduction of graphene oxide. Int. J. Biol. Macromol. 2018, 114, 599–613. [Google Scholar] [CrossRef]
  225. Chen, J.; Li, H.; Zhang, L.; Du, C.; Fang, T.; Hu, J. Direct Reduction of Graphene Oxide/Nanofibrillated Cellulose Composite Film and its Electrical Conductivity Research. Sci. Rep. 2020, 10, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Kinloch, I.A.; Suhr, J.; Lou, J.; Young, R.J.; Ajayan, P.M. Composites with carbon nanotubes and graphene: An outlook. Science 2018, 362, 547–553. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Kashi, S.; Gupta, R.K.; Baum, T.; Kao, N.; Bhattacharya, S.N. Phase transition and anomalous rheological behaviour of polylactide/graphene nanocomposites. Compos. Part B Eng. 2018, 135, 25–34. [Google Scholar] [CrossRef]
  228. Fryń, P.; Bogdanowicz, K.A.; Górska, N.; Rysz, J.; Krysiak, P.; Marzec, M.; Marzec, M.; Iwan, A.; Januszko, A. Hybrid materials based on L,D-poly(lactic acid) and Single-Walled Carbon Nanotubes as flexible substrate for organic devices. Polymers 2018, 10, 1271. [Google Scholar] [CrossRef] [Green Version]
  229. Kashi, S.; Gupta, R.K.; Kao, N.; Hadigheh, S.A.; Bhattacharya, S.N. Influence of graphene nanoplatelet incorporation and dispersion state on thermal, mechanical and electrical properties of biodegradable matrices. J. Mater. Sci. Technol. 2018, 34, 1026–1034. [Google Scholar] [CrossRef]
  230. Dhar, P.; Pratto, B.; Jose, A.; Cruz, G.; Bankar, S. Valorization of sugarcane straw to produce highly conductive bacterial cellulose/graphene nanocomposite fi lms through in situ fermentation: Kinetic analysis and property evaluation. J. Clean. Prod. 2019, 238, 117859. [Google Scholar] [CrossRef]
  231. Zhang, H.; Sun, X.; Hubbe, M.A.; Pal, L. Highly conductive carbon nanotubes and flexible cellulose nano fibers composite membranes with semi-interpenetrating networks structure. Carbohydr. Polym. 2019, 222, 115013. [Google Scholar] [CrossRef]
  232. Siripongpreda, T.; Siralertmukul, K.; Rodthongkum, N. Colorimetric sensor and LDI-MS detection of biogenic amines in food spoilage based on porous PLA and graphene oxide. Food Chem. 2020, 329, 127165. [Google Scholar] [CrossRef]
  233. Ruiz-Hitzky, E.; Aranda, P.; Akkari, M.; Khaorapapong, N.; Ogawa, M. Photoactive nanoarchitectures based on clays incorporating TiO2 and ZnO nanoparticles. Beilstein J. Nanotechnol. 2019, 10, 1140–1156. [Google Scholar] [CrossRef] [Green Version]
  234. Lizundia, E.; Ruiz-Rubio, L.; Vilas, J.L.; León, L.M. Poly(L-lactide)/ZnO nanocomposites as efficient UV-shielding coatings for packaging applications. J. Appl. Polym. Sci. 2016, 133, 42426. [Google Scholar] [CrossRef]
  235. Ruiz-Hitzky, E. Molecular access to intracrystalline tunnels of sepiolite. J. Mater. Chem. 2001, 11, 86–91. [Google Scholar] [CrossRef]
  236. Ruiz-Hitzky, E.; Aranda, P. Novel architectures in porous materials based on clays. J. Sol Gel Sci. Technol. 2014, 70, 307–316. [Google Scholar] [CrossRef]
  237. Aranda, P.; Ruiz-Hitzky, E. Immobilization of Nanoparticles on Fibrous Clay Surfaces: Towards Promising Nanoplatforms for Advanced Functional Applications. Chem. Rec. 2018, 18, 1125–1137. [Google Scholar] [CrossRef] [PubMed]
  238. Fadeel, B.; Bussy, C.; Merino, S.; Vázquez, E.; Flahaut, E.; Mouchet, F.; Evariste, L.; Gauthier, L.; Koivisto, A.J.; Vogel, U.; et al. Safety Assessment of Graphene-Based Materials: Focus on Human Health and the Environment. ACS Nano 2018, 12, 10582–10620. [Google Scholar] [CrossRef] [PubMed]
  239. Newman, L.; Rodrigues, A.F.; Jasim, D.A.; Vacchi, I.A.; Ménard-Moyon, C.; Bianco, A.; Bussy, C.; Kostarelos, K. Nose-to-Brain Translocation and Cerebral Biodegradation of Thin Graphene Oxide Nanosheets. Cell Rep. Phys. Sci. 2020, 1, 100176. [Google Scholar] [CrossRef]
  240. Rodrigues, A.F.; Newman, L.; Jasim, D.; Mukherjee, S.P.; Wang, J.; Vacchi, I.A.; Ménard-Moyon, C.; Bianco, A.; Fadeel, B.; Kostarelos, K.; et al. Size-Dependent Pulmonary Impact of Thin Graphene Oxide Sheets in Mice: Toward Safe-by-Design. Adv. Sci. 2020, 7, 1–17. [Google Scholar] [CrossRef]
Figure 1. Upper scheme: (A) Evolution of commercial production capacities of bio-based and biodegradable polymers within the packaging plastic market between 2017 and 2019; (B) the representativity of different biopolymers on the 2019 market. Data for the graphical construction acquired from the European Bioplastics Report [4]. Lower scheme: (C) Number of published research articles focused on the various bio-based and biodegradable polymers combined with the term “food packaging” in 2017 and 2020; (D) the distribution of the diverse biopolymers in research in 2020 (accessed on 7 October 2020 through “The Web of Science”).
Figure 1. Upper scheme: (A) Evolution of commercial production capacities of bio-based and biodegradable polymers within the packaging plastic market between 2017 and 2019; (B) the representativity of different biopolymers on the 2019 market. Data for the graphical construction acquired from the European Bioplastics Report [4]. Lower scheme: (C) Number of published research articles focused on the various bio-based and biodegradable polymers combined with the term “food packaging” in 2017 and 2020; (D) the distribution of the diverse biopolymers in research in 2020 (accessed on 7 October 2020 through “The Web of Science”).
Nanomaterials 10 02077 g001
Figure 2. Structures of biopolymers.
Figure 2. Structures of biopolymers.
Nanomaterials 10 02077 g002
Figure 3. Classification grid for the categorization of different graphene types according to three fundamental graphene derivatives properties. The different materials drawn at the six corners of the box represent the ideal cases according to the lateral dimensions and the number of layers reported in the literature. The values of the three axes are related to the graphene derivatives at the nanoscale, but it is feasible to expand the values to the microscale. Reproduced with permission from [165]. Copyright Wiley-VCH Verlag GmbH & Co, KGaA, Weinheim, Germany, 2014.
Figure 3. Classification grid for the categorization of different graphene types according to three fundamental graphene derivatives properties. The different materials drawn at the six corners of the box represent the ideal cases according to the lateral dimensions and the number of layers reported in the literature. The values of the three axes are related to the graphene derivatives at the nanoscale, but it is feasible to expand the values to the microscale. Reproduced with permission from [165]. Copyright Wiley-VCH Verlag GmbH & Co, KGaA, Weinheim, Germany, 2014.
Nanomaterials 10 02077 g003
Figure 4. (Upper scheme) Dispersion of platelet-like fillers into polymer composites. Adapted from reference [177]. Copyright Elsevier, 2018. (Lower scheme) Interactions between typical carbon nanostructures and different polymers in composites.
Figure 4. (Upper scheme) Dispersion of platelet-like fillers into polymer composites. Adapted from reference [177]. Copyright Elsevier, 2018. (Lower scheme) Interactions between typical carbon nanostructures and different polymers in composites.
Nanomaterials 10 02077 g004
Figure 5. PEF processing in-pack. Food is packaged into an electrically conductive food packaging prior to PEF Table 2. Adapted from reference [217] Copyright Wiley-VCH Verlag GmbH & Co, KGaA, Weinheim, 2011.
Figure 5. PEF processing in-pack. Food is packaged into an electrically conductive food packaging prior to PEF Table 2. Adapted from reference [217] Copyright Wiley-VCH Verlag GmbH & Co, KGaA, Weinheim, 2011.
Nanomaterials 10 02077 g005
Table 1. Preparation methods and mechanical properties of biopolymer/carbon nanostructures composites.
Table 1. Preparation methods and mechanical properties of biopolymer/carbon nanostructures composites.
PolymerNanomaterialPreparation MethodNanomaterial Dispersion StrategiesMain Mechanical EffectsRef.
PHBV0.5 to 0.7 wt% GO,
1:0.5 wt% non-grafted GO/CNC,
1 wt% grafted GO-CNC
Solvent castingPhysical blending (stirring); chemical graftingCovalently grafted GO-CNC achieved the highest YM, TS, and EB values, which were up to 138%, 170%, and 52% higher than neat polymer.[149]
CS0.5 wt% GO with different degrees of oxidationSolvent castingUltrasonic dispersionBy increasing of oxidation degree of GO, the TS and YM increase and the EB decreases.[152]
0.25 wt% GO and 3 wt% ball-milled maleamic acid–isobutyl polyoctahedral silsesquioxanes (MAIPS)Solvent castingPhysical blendingSynergistic reinforcements were found on the composite with GO and MAIPS: highest YM and TS (e.g., 50% and 38% higher, respectively, than neat polymer).[186]
5 wt% GNP and 5 wt% ZnOSolvent castingUltrasonic dispersionThe simultaneous incorporation of GNP and ZnO lead to highest values of YM and TS, and to a slight decrease of EB.[178]
GO (0, 25, 40, 45, 48, or 50 wt%, in relation to
CS weight)
Solvent castingUltrasonic dispersionCS/GO showed higher TS (improvements of 70% to 110%), YM (improvements of 500%), and lower EB (decay of 90%) when compared with chitosan films. No significative differences were found in CS-based composites with 40 to 50 wt% of GO.[170]
0 to 30 wt% GNP or MWCNTSolvent castingUltrasonic dispersionAt the same ratios, CS/GNP and CS/MWCNT exhibited similar TS and YM values. The highest values of TS were achieved by incorporation of 15 wt% of GNP or MWCNT, which represented improvements of 49% and 64% when compared to those of neat polymer. In turn, the highest values of YM were achieved by incorporation of 30 wt% of GNP or MWCNT, which represented improvements of 109% and 115% when compared to those of neat polymer.[183]
Starch3, 6, and 9 wt% MWCNT grafted with ascorbic acid (AA-MWCNT)Solvent castingUltrasonic dispersionThe YM and TS were reduced and the EB was increased by enhancing the AA-MWCNT loading in the composite.[182]
PLA1 wt% GNP and CNC (ratio 50/50)Hot pressingMelt blending with the Triton X-100 surfactantImprovements on YM, TS, and EB were achieved by simultaneous incorporation of both nanofillers.[173]
0.5 wt% GO and 1 wt% CNCSolvent castingPhysical blendingIncrease of PLA/CNC/rGO nanocomposite TS up to 23%.[187]
0.05 to 2 wt% GO-Ag hybridsSolvent casting or direct mechanical melt blendingPhysical blending or melt blendingHigher flexural strength was achieved when higher amounts of GO-Ag hybrids were added and when physical blending and solvent casting subsequent methods were applied.[188]
2 wt% CNTSolvent castingUltrasonic dispersionThe TS and EB have an enhancement of 52% and 36%, respectively, in comparison with PLA films.[174]
0.5, 1.0, 2.0, and 3.0 wt% MWCNTInjection moldingMechanical blendingIncrements of 32.70% and 67.17% were obtained for the TS and EB with the inclusion of 3 wt% of MWCNT.[175]
PLA/CNCs0.9 to 8.3 wt% CNTCompression
molding
Pickering emulsionsThe mechanical performance of the sample was maintained a high level (tensile strength: 45.52 MPa, Young’s modulus: 3152 MPa) after the incorporation of 4.3 wt% CNT.[172]
Alginate0 to 25 wt% GOSolvent castingPhysical blendingThe inclusion of >2 wt% GO content into alginate-based composites demonstrated remarkable improvements in YM. The maximum upgrade achieved was of 230% in comparison with pure alginate (15 wt% GO). The evolution of the TS suggested the inclusion of defects in the microstructure as GO increased.[179]
PHBV: poly(3-hydroxybutyrate- co -3-hydroxyvalerate). GO: graphene oxide. CNC: cellulose nanocrystals. YM: Young’s modulus. TS: tensile strength. EB: elongation at break. CS: chitosan. MAIPS: ball-milled maleamic acid–isobutyl polyoctahedral silsesquioxanes. CNT: carbon nanotubes. PLA: polylactic acid. MWCNT: multi-walled carbon nanotubes. GNP: graphene nanoplatelets. ZnO: zinc oxide. Ag: silver particles.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Barra, A.; Santos, J.D.C.; Silva, M.R.F.; Nunes, C.; Ruiz-Hitzky, E.; Gonçalves, I.; Yildirim, S.; Ferreira, P.; Marques, P.A.A.P. Graphene Derivatives in Biopolymer-Based Composites for Food Packaging Applications. Nanomaterials 2020, 10, 2077. https://doi.org/10.3390/nano10102077

AMA Style

Barra A, Santos JDC, Silva MRF, Nunes C, Ruiz-Hitzky E, Gonçalves I, Yildirim S, Ferreira P, Marques PAAP. Graphene Derivatives in Biopolymer-Based Composites for Food Packaging Applications. Nanomaterials. 2020; 10(10):2077. https://doi.org/10.3390/nano10102077

Chicago/Turabian Style

Barra, Ana, Jéssica D. C. Santos, Mariana R. F. Silva, Cláudia Nunes, Eduardo Ruiz-Hitzky, Idalina Gonçalves, Selçuk Yildirim, Paula Ferreira, and Paula A. A. P. Marques. 2020. "Graphene Derivatives in Biopolymer-Based Composites for Food Packaging Applications" Nanomaterials 10, no. 10: 2077. https://doi.org/10.3390/nano10102077

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