Next Article in Journal
Optimization of the Electrospray Process to Produce Lignin Nanoparticles for PLA-Based Food Packaging
Next Article in Special Issue
High-Barrier, Biodegradable Films with Polyvinyl Alcohol/Polylactic Acid + Wax Double Coatings: Influence of Relative Humidity on Transport Properties and Suitability for Modified Atmosphere Packaging Applications
Previous Article in Journal
Thermal Performance of Multifunctional Facade Solution Containing Phase Change Materials: Experimental and Numerical Analysis
Previous Article in Special Issue
Bioactive Absorbent Chitosan Aerogels Reinforced with Bay Tree Pruning Waste Nanocellulose with Antioxidant Properties for Burger Meat Preservation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Reinforcements and Additives in Starch-Based Composites for Food Packaging

by
Pedro Francisco Muñoz-Gimena
,
Víctor Oliver-Cuenca
,
Laura Peponi
* and
Daniel López
*
Instituto de Ciencia y Tecnología de Polímeros (ICTP-CSIC), C/Juan de la Cierva 3, 28006 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(13), 2972; https://doi.org/10.3390/polym15132972
Submission received: 7 June 2023 / Revised: 30 June 2023 / Accepted: 4 July 2023 / Published: 7 July 2023

Abstract

:
The research of starch as a matrix material for manufacturing biodegradable films has been gaining popularity in recent years, indicating its potential and possible limitations. To compete with conventional petroleum-based plastics, an enhancement of their low resistance to water and limited mechanical properties is essential. This review aims to discuss the various types of nanofillers and additives that have been used in plasticized starch films including nanoclays (montmorillonite, halloysite, kaolinite, etc.), poly-saccharide nanofillers (cellulose, starch, chitin, and chitosan nanomaterials), metal oxides (titanium dioxide, zinc oxide, zirconium oxide, etc.), and essential oils (carvacrol, eugenol, cinnamic acid). These reinforcements are frequently used to enhance several physical characteristics including mechanical properties, thermal stability, moisture resistance, oxygen barrier capabilities, and biodegradation rate, providing antimicrobial and antioxidant properties. This paper will provide an overview of the development of starch-based nanocomposite films and coatings applied in food packaging systems through the application of reinforcements and additives.

Graphical Abstract

1. Introduction

The use of plastics is essential to many sectors of our economy [1] due to their low production costs, durability, good mechanical properties, and resistance to oil and most chemicals. Consequently, the food packaging sector is the sector with the highest percentage of plastic use [2]. Although the use of non-biodegradable plastics is a matter of concern in society and multiple industries [1], the accumulation of the residues produced by these plastics is a critical problem for the environment, causing several problems in our daily lives and affecting nearly every ecosystem. Moreover, plastics pose a threat to oceans, wildlife, public health, and economic sectors that directly depend on marine wildlife health [3,4,5].
Therefore, to minimize the impact of plastics, it is vital to develop new polymeric matrices that are biodegradable and eco-friendly. Polymers derived from renewable sources, such as starch, cellulose, or chitin, are an excellent approach to fulfill these conditions. However, the current transition rate from petroleum-based polymers to biodegradable polymers is rather low and still inconsistent. Companies prefer to recycle already-used plastics instead of investing in new polymer matrices due to the poorer mechanical and barrier properties that these newly developed matrices show [6,7].
To improve these limiting properties, a push toward the research and development of new and suitable matrices has to be made. In this context, starch-based polymers rise as one of the most promising alternatives to traditional petroleum-based polymers in the food packaging sector. The growing interest in sustainable plastics can be seen in Figure 1 by the overwhelming increase in research papers and reviews published dealing with the topic: since 2003, almost 600 scientific papers have been published, of which 559 were published after 2013, and 132 during the last year, 2022 (source: Scopus; keywords: starch, active and packaging). The significant increase in research on this topic highlights the demand for sustainable packaging materials.
Starch is a natural biopolymer extracted from conventional botanical sources such as potatoes, rice, or corn, but can also be extracted from non-conventional resources including roots, skins, and pits from vegetables and fruits, most of which are discarded and thrown away. Therefore, starch is abundant and can give value to what is commonly considered “waste”.
Starch usage has been described for thousands of years, from the ancient Egyptians (~4000 BC) and later the Romans using it as an adhesive or the ancient Greeks incorporating it in medical preparations [8]. Nowadays, starch applications range from ingredients in multiple food products (e.g., canned goods, bakery products, baby food, soups and creams, sauces, and dressings) to adhesives, aerogel, films, or edible coatings in the textile, pharmaceutical, paper, packaging, and cosmetic industries [9]. Moreover, starch is considered one of the most promising sustainable materials due to its abundance and excellent biodegradability, with new applications being developed across multiple sectors: green chemistry, water treatments, medical applications (tissue engineering, controlled drug release, and medical dressings) and flexible electronics (supercapacitors, sensors, and conductive substrate films) [10]. However, since native starch granules do not dissolve in cold water, transformation into de-structured thermoplastic starch (TPS) is needed for most starch-based applications [11]. As an example, in the food sector, gelatinization through cooking or processing greatly increases starch digestibility, as the disruption of the ordered structure of gelatinized starch allows easy access of digestive enzymes to starch molecules [12].
In the packaging sector, thermoplastic starch (TPS)-based films not only match the properties shown by conventional petroleum-based polymer films, but in some cases, they show a superior capacity to form colorless, tasteless, and translucent layers [13] as well as having excellent gas barrier properties against oxygen and carbon dioxide [14]. TPS can be obtained through two methods: the casting solution method and the extrusion process. Nevertheless, starch films have multiple drawbacks: high hydrophilic behavior, which makes them difficult to use in contact with food that has a high water content, and lower mechanical and water vapor properties compared to petroleum-based polymer films [15,16]. Moreover, starch-based films generally present no antimicrobial or antioxidant properties by themselves. To enhance performance and processability, frequently used strategies include using polymer blends, coating articles with high-barrier materials, using multilayered films containing other films, chemical modification of natural biopolymers, and producing composite films using different fillers [17].
Composite films consist of two phases: a continuous polymeric matrix and a discontinuous filler (organic or inorganic materials with specific shapes) [18]. Moreover, it has been well established that the filler properties and size greatly influence the properties of the nanocomposite, with smaller/nanosized fillers generally having a greater impact [19].
Thus, the objective of this review is to report an interesting scientific discussion on starch-based films for food packaging applications, focusing attention on the reinforcements and additives used and their main effects on starch-based film properties. In particular, after analyzing starch botanical sources and their extraction techniques, the properties of thermoplastic starch are discussed, including common preparation techniques such as solvent casting, extrusion, and film blowing. Then, the most typical organic and inorganic nanofillers used as reinforcements are described, detailing their effects on the mechanical response. Furthermore, additives such as essential oils and functional organic molecules such as carvacrol, eugenol, and cinnamic acid are reported to focus on their antioxidant and/or antimicrobial properties and their application in smart packaging. In the end, a short discussion of their perspective is provided from a circular economy point of view.

2. Starch

2.1. Starch Sources and Extraction

The structure of starch is formed by the succession of α-D-glucose linked to each other through α (1 → 4) glycosidic bonds [20]. Two different structures are found inside starch matrices: amylose, which has a non-branched, lineal, helical structure with only α (1 → 4) glycosidic bonds, and amylopectin, which has a highly branched structure with 24 to 30 residues of glucose units, as seen in Figure 2. This structure comes as a result of the presence of α (1 → 6) glycosidic bonds, which are responsible for the branched structure, as well as α (1 → 4) glycosidic bonds, like amylose [21]. The ratio between both components fluctuates due to multiple factors, the most important being the source of extraction of starch or even the conditions of the vegetable source. In general, most native starches contain 20–30% amylose and 70–80% amylopectin [22].
Starch can be extracted from multiple sources and, as a result, the ratio between amylose and amylopectin as well as the characteristics and properties of the obtained native starch can vary. Several methods have been used for starch extraction, of which two are the most commonly used. In the first method, starch extraction is carried out after treatment with a dissolution of hydrogen sulfite 0.1 M at 50 °C for 24 h, after which the substrate is ground in a laboratory grinder with distilled water to extract all the starch. After several screenings through different mesh sizes, the starch slurry is centrifuged several times and then dried in an oven for 12 h at 40 °C [24]. In the second method, starch is obtained through treatment with an aqueous solution of sodium metabisulfite 0.1 M for 30 min, after which the subtract is ground with distilled cold water in a laboratory grinder. The homogenate is then screened through different mesh sizes and then the filtrate is kept undisturbed until the complete precipitation of starch which is recovered and treated with an aqueous solution of NaOH 0.2 M for 1–2 h and successive washes with distilled water, and finally dried at 40 °C for 12 h [25]. Among the many possible botanical sources for starch extraction, Table 1 shows the main characteristics of three of the most common starch sources: corn, potato, and rice.
The shape and size of the starch granules are characteristic of the source from which the starch was extracted. In corn, starch granules are angular shaped with a diameter of 11.5 ± 4.3 µm; potato starch granules exhibit large irregular and oval or cuboidal forms, with typical granule size for small and large potato starch granules between 1 and 20 µm. Meanwhile, rice starch granules present pentagonal and angular forms with a size of 3 to 5 µm [37]. A wide variety of starch botanical sources and their granular forms is reported in Table 2 and Figure 3.
Table 2. Size and shape of starch granules from different botanical sources [38].
Table 2. Size and shape of starch granules from different botanical sources [38].
Plant SourceSize (µm)Shape
Maize2–30Round and polyhedral
Waxy maize2–30Round and polyhedral
Wheat10 and 10–30Discs
Waxy wheat>10Spherical and lenticular
Rice<20 and 2–8Polygonal and angular
Tapioca5–45Spherical/lenticular
Cassava30Kettle-drum truncated
Potato<110Oval and irregular
Sweet Potato1–100Oval, spherical, round
Pea2–40Oval, spherical, round
Lotus14–31Oblong
Figure 3. SEM images of various starches: (A) potato (1000×), (B) corn (1000×), (C) wheat (500×), and (D) rice (2000×) [39].
Figure 3. SEM images of various starches: (A) potato (1000×), (B) corn (1000×), (C) wheat (500×), and (D) rice (2000×) [39].
Polymers 15 02972 g003

2.2. Thermoplastic Starch, TPS

Starch has received significant interest in the form of a thermoplastic polymer, namely thermoplastic starch, TPS. Even though starch is a poor choice to replace plastic as it does not exhibit true thermoplastic behavior, it can be transformed into a continuous polymeric-entangled phase through the addition of sufficient water and/or plasticizers [40]. During the processing of TPS, various physical and chemical reactions are involved such as water diffusion, expansion of granules, gelatinization, melting, and crystallization [41]. The disruption process of starch to form thermoplastic starch requires generating an amorphous matrix by the complete or partial disintegration of the granular architecture. In particular, starch conversion into TPS in the presence of water follows a three-step process: granular swelling, amylopectin cluster breakage, and amylose/amylopectin solubilization, as seen in Figure 4 [42]. During the first stage, the absorption of water in its amorphous domains swells the starch granule. This process is reversible when the temperature remains low; however, if the temperature exceeds the gelatinization temperature (Tgel), the granular structure starts to collapse. The specific Tgel is a characteristic property of each starch, linked to the branch chain length of amylopectin, with values for common starches ranging between 50 and 85 °C [43]. However, gelatinization is a complex process, as it depends on multiple variables including botanical source, water content, and composition (amylose/amylopectin ratio, protein, lipids, ash content, etc.), as well as morphology and size of the starch granules [44]. For instance, under differential scanning calorimetry (DSC) study, a single peak is observed when the gelatinization water content is in an excessive amount, while multiple peaks and bumps can be found from gelatinization thermograms when the water content is below, e.g., 60% [45]. On a similar note, amylose content has been reported to both positively and negatively influence the gelatinization temperature [44,46].
The glass transition temperature (Tg) of starch is substantially higher than its decomposition temperature [47], thereby impeding its potential usage as a bioplastic. To attain the capacity to process native starch, it is essential to utilize plasticizers, small molecules with low volatility, to decrease the melting point. Additionally, the inherent brittleness, inadequate mechanical properties, and barrier properties of starch-only films pose a significant challenge in their handling. During the thermoplastic process, water contained in starch and the addition of plasticizers (such as glycerol or sorbitol) decrease the internal starch hydrogen bonding [48], increasing the flexibility of the resultant films and improving the barrier and sorption properties of films, causing better packaging qualities [49].
The process of generating a starch film initiates with the application of heat to starch granules in an excess of water, thus facilitating the creation of a viscous solution. The viscoelastic behavior of TPS can be classified according to the quantity of plasticizer present: a glassy profile (hard and stiff but brittle) at low plasticizer concentrations, followed by a rubbery behavior (flexible) characterized by a glass-to-rubber transition as plasticizer content increases, concluding in a gel-like state at very high plasticizer content [50]. The microstructure, particle size, and roughness of the starch source influence the viscosity of the film-forming suspension, promoting a modification of the filmogenic matrix after the drying process, which affects the physicochemical and mechanical properties of the resulting film [51,52].
TPS exhibits a few limitations in its utilization, including: a significant hydrophilic nature, making it susceptible to water; relatively poor mechanical attributes when compared to conventional polymers; and a significant variation in post-processing properties [53]. In addition, retrogradation and crystallization of the mobile starch chains lead to an undesired change in thermomechanical properties [54]. Susceptibility to humidity can be observed in a typical plasticized TPS model system. Under a low humidity environment, stability is adversely affected due to the diffusion of plasticizer out of the TPS matrix, while during high humidity conditions, water diffuses into the product. Consequently, the loss of plasticizer in a low-humidity environment prompts brittleness, whilst the excess absorption of moisture in a high-humidity environment causes shape- and texture-retention concerns [55]. The relative air humidity and the plasticizer type and concentration affect the film water vapor permeability (WVP) [56]. Another undesired effect is starch retrogradation, which is a phenomenon used to describe the stiffening and decreased extensibility of TPS materials over time, linked to the association of dissociated amylose and amylopectin chains in a gelatinized starch dispersion forming more ordered structures [51].

2.3. Preparation of Starch-Based Films

There are three main techniques for the preparation of starch-based films; the casting procedure, the extrusion/thermocompression process, and the blow molding approach, which are briefly discussed in this review. Additionally, other alternative processing techniques based on the plastic industry have also been used for the elaboration of starch films, including injection molding, compression molding, foaming, and spinning (melt-, wet- and electro-spinning) [57].

2.3.1. Solvent Casting

The solvent casting technique is by far the most popular in starch film production and has been widely documented in the academic literature. The technique was developed at the end of the 19th century to manufacture thin plastic films for the emerging industry of movies and photography [58].
The casting of starch-based films typically includes four steps: solution preparation, gelatinization, casting, and drying. As a general method [59,60], the starch and plasticizers are mixed in water and heated to 80–95 °C for 10–15 min while under constant shaking. The gelatinized suspensions are poured on Teflon or acrylic plate and dried in an oven (40–75 °C) for 24 h to constant weight. The film thickness, usually 0.02–0.10 mm, can be adjusted by the amount and concentration of the starch suspension poured into the mold.
Despite its simplicity and popularity, the long drying phase makes this method slow and discontinuous, producing small quantities of material after a long process. Moreover, the casting methodology presents certain drawbacks: difficulty when combining materials of different natures; separating the film from the underlying support can be challenging; and limited capacity of scaling up the technique from an experimental setup to industrial production [61].

2.3.2. Tape Casting

Tape casting was first introduced by Glenn Howatt in the 1940s, to produce capacitor dielectrics, and, to this day, continues to be one of the main applications for this technology [62]. The technique can be seen as an upgraded version of the solvent casting method, as it has the potential to facilitate researchers utilizing the classical casting method to scale up production and create biopolymeric films with significantly larger dimensions. However, there is very limited literature about the utilization of tape casting on starch-based films, with the majority elaborated by the same research group, and the method is probably limited by the difficult technological transfer from other industries [63].
The tape casting method allows the casting of the film-forming solution on large supports (batch process) and on continuous belts (continuous process), where the film thickness can be modulated by an adaptable blade positioned at the base of the spreading device [64], as seen in Figure 5. The support where the film suspension is placed permits a regulated drying via heat conduction, hot air circulation (heat convection), and/or infrared radiation; with a consequent reduction in the final thickness, usually ranging between 20 μm and 1 mm [65]. Scheibe et al. [66] successfully tape-casted starch-glycerol-fiber films and starch-glycerol-cellulose fiber bags with different properties by changing the film thickness.

2.3.3. Extrusion

From an industrial standpoint, the manufacturing of TPS demands a highly effective mass-production technique; a simple solution is using processes with the standard equipment used in the manufacture of synthetic polymers. Considering this, hot extrusion presents itself as an answer to scale up biopolymeric film production.
Throughout the extrusion process, starch is subject to elevated pressure and temperature, as well as the application of mechanical shear forces, leading to a range of chemical and physical reactions, disintegrating the starch granules [68]. The plasticization and melting of TPS can be facilitated through the heating and shearing effects resulting from the rotation of internal screws within the barrels utilizing electric heating, where temperature manipulation can be employed across various zones within cylindrical barrels to attain desired outcomes relating to mixing, flow, pressure, and flow rates [69].
González-Seligra et al. [70] studied the effect of screw speed on starch-glycerol-water mixes extruding under the same temperature profile and process conditions. A slow screw speed (40 rpm) resulted in high non-gelatinized broken grains and uneven glycerol distribution, leading to high crystallinity and low deformation resistance. Extrusion at 80 rpm was ideal for homogeneous TPS in threads and films but presented accelerated retrogradation. Finally, high speed extrusion (120 rpm) achieved stronger amorphous TPS films with slower retrogradation. On a similar note, Li et al. [71] studied the effect of modifying amylose content on extruded TPS films. The study found that processability decreased with increased amylose content, but the effects could be countered by modifying the extrusion parameters. Higher amylose films presented better mechanical properties and high impact strength. Moreover, the glass transition temperature (Tg) of the extruded films was studied using DMA, and showed a rising trend with increasing amylose content, attributed to the different molecular sizes of amylose and amylopectin and their phase transitions.

2.3.4. Film Blowing

The process of film blowing is a technique frequently used to manufacture self-supporting plastic films, which involves extruding a hollow tube and then expanding it into a bubble by raising the pressure within the tube [57]. The popularity of film blowing in plastic film production is due to its simple production equipment, low investment, continuous production, and adjustable film size [72].
Rejak, A. and Mosciki, L. [73] processed different starches through film blowing and found maize and wheat starch to be unfit for packaging due to their brittle, opaque and thick films obtained. However, potato starch films led to semi-transparent films amenable to blow-molding. Furthermore, the addition of emulsifiers further improved film flexibility and strength. On a similar note, Dang & Yoksan [74] examined the effect of replacing 50% of glycerol with larger-molecular plasticizers, xylitol or sorbitol, in TPS blown films. While the glycerol-only films presented higher melt flowability and elongation at break, and lower Tg, they also displayed stickiness and shrinkage. Meanwhile, the mixed plasticizers achieved significant tensile strength, stiffness, water vapor barrier and oxygen barrier properties, and dimension stability improvements.
Nonetheless, TPS production by film blowing may result to be complex due to the low melting tenacity of starch [57]. Therefore, a frequent approach to ease the processability and mechanical properties of the resulting films is the preparation of starch blends combined with other biodegradable polymers, such as PLA [75], PCL [76], PBAT [77,78] or chitosan [79].

3. Nanoparticles

The hydrophilic characteristic of starch results in elevated water vapor permeability (WVP) and low tensile strength and Young’s modulus in contrast to synthetic materials of manufactured products. A feasible approach to enhance these attributes involves the integration of reinforcement fillers to create starch-based nanocomposites.
The incorporation of nanoparticles as fillers in polymers is a widely recognized technique to enhance the properties of polymeric materials while using less product. For an equal amount of filler, if the dispersion exhibits homogeneity, the influence on the macroscopic characteristics will be more pronounced for nanoparticles of a smaller size, attributable to the augmented polymer/filler interface [80]. However, the aggregation of nanoparticles may result in an adverse impact on the modulus and strength of the material as a consequence of the formation of rigid regions within the polymer matrix and a decline in the interfacial surface area [81]. Several techniques are often used to obtain a complete and homogenous dispersion of NPs in the polymeric matrix: mechanical mixing (i.e., high-speed shear milling, three-roll milling, and high-pressure homogenizing) or ultrasound sonication [82,83].
Nanoparticles can be grouped by their nature into inorganic nanofillers (e.g., metal and metal oxide nanoparticles, carbon nanotubes, and nanolayered clays) and organic nanofillers (e.g., cellulosic whiskers, lignocellulosic materials, chitin whisker, and starch nanocrystal) [84]. Alternatively, nanoparticles can also be classified into three main categories according to how many of their dimensions are on the nanometric scale and according to their geometry [84,85,86], as seen in Figure 6.
Recently, Wolf et al. [84] compared hundreds of nanocomposite studies in an attempt to compare the tortuosity effect of different nanoparticle shapes on barrier properties. While the review did confirm that layered nanoparticles were more efficient than isodimensional and elongated nanoparticles, no clear trend was established on the selective transport mechanisms.
As nanotechnology innovations make their way into different industrial sectors, many countries are establishing controls over the use of nano-based products. However, regulation frameworks applied to control and ensure the safe use of nanotechnologies differ across countries and regions [87]. For example, TiO2 has approval for use in human food, medicines, cosmetics, and food contact products by the US FDA [88]. However, due to the potential genotoxicity of TiO2 NPs, the European Commission in 2022 banned the use of TiO2 as a food additive and encourage its replacement in medicines over health safety concerns [89].
In the EU, nanomaterials are covered by the same regulatory framework that ensures the safe use of all chemicals and mixtures, i.e., the REACH and CLP regulations. The EU has added regulatory measures for nanomaterials in the food and feed chain, together with risk assessment guidance. Current EU legislation includes Regulation (EC) No. 1169/2011 on the provision of food information to consumers, which defines ‘engineered nanomaterial’ as any intentionally produced material with at least one or more dimensions of the order of 100 nm or less, including structures, agglomerates and aggregates which may exceed this size but retain properties of the nanoscale [90]. EU Regulation on Plastic Food Contact Materials and Articles (10/2011) sets an overall migration limit (OML) and specific migration limits (SML). OML is set at 60 mg (of substance) per kg (of food or food stimulant) to guarantee that packaging materials do not transmit excessive amounts of compounds which, even if they are not hazardous, might cause an undesirable change in the composition of the food [91]. As an example of a safety assessment, the study on silver nanoparticles (Ag NPs) for use in food contact materials [92] determined that Ag NPs remained embedded in the polymeric matrix and did not pose a toxicological threat. In ionic form, silver did present low levels of migration which did not exceed the proposed acceptable daily intake (ADI) of 0.9 µg/kg per day. Nevertheless, the study highlights that estimated exposure combined with other dietary sources already exceeded the set ADI limit.

3.1. Inorganic Nanofillers

Incorporating inorganic nanoparticles is a highly effective form of enhancing the mechanical and physical properties of starch-based films. The incorporation of these additives in polymeric nanocomposite causes an improvement in mechanical and barrier qualities, including enhanced gas barrier qualities, and stability across a range of temperature and moisture levels [93]. Furthermore, some of them can also add new functional properties to these matrices including antimicrobial properties by reducing the growth of microorganisms [94]. There are many different types of fillers in this category, including metal nanoparticles, metal oxides, and nanoclays.

3.1.1. Nanoclays

An interesting class of nanofillers that solves the issues of starch matrices previously mentioned is nanoscale lamellar silicates, commonly referred to as nanoclays. The economic cost, biocompatible nature, and widespread availability of nanoclays have seen them being frequently utilized for the production of biodegradable films and coatings, as well as synthetic films to improve their barrier and thermal stability properties [95]. According to reports [96], the incorporation of nanoclays, even at low concentrations, can produce a reduction in gas permeability through polymer films of 50–500 times. The integration of nanoclays into polymers emerged as a pioneering polymer nanocomposite technology, yielding a novel material for the purpose of food packaging [97], and representing nearly 70% of the market volume of commercial nanoscale applications in the early 2010s [98].
Typically, within a polymer matrix, nanoclays can display three distinct morphologies of nanoclay, depending on the dispersion state of the nanoclay, and the degree of separation between the silicate layers, as well as the level of polymer penetration between them. These morphologies are: aggregated, intercalated, and exfoliated [99], as seen in Figure 7. In nanocomposites, the clay layers must be uniformly dispersed in the polymer matrix (intercalated or exfoliated), as opposed to being aggregated as tactoids [100].
The impermeable properties of nanoclays dispersed within the film oblige water molecules to travel through a convoluted path to cross the matrix, improving barrier properties to water vapor [100]. Increasing the aspect ratio and the volume fraction of the nanoplatelets improves the barrier properties of nanocomposites [96]. Several studies have documented nanoclay-based packaging materials employed in conjunction with different active agents to achieve film formation that manifests antimicrobial activity [101]. Moreover, the enhanced tortuosity impact of dispersed nanoclays effectively prolongs the retention or regulates the transfer of antibacterial agents through polymer matrices.
Osman et al. [102] demonstrated the importance of nanofiller dispersion by contrasting dolomite/starch films versus ultrasonicated dolomite/starch; sonication reduced particle size, increased crystallinity and matrix homogeneity, and significantly improved mechanical properties. On a similar note, Gao et al. [103] tested the effect of five clays with different hydrophilicity levels on the properties of starch-clay nanocomposites by film blowing. All nanoclay-reinforced films presented improved mechanical and barrier properties compared to the control film. While the higher hydrophilic clay presented the lowest WVP, the medium hydrophilic clay exhibited better compatibility and dispersion within the starch matrix, which resulted in the highest tensile strength and lowest elongation.
Interestingly, Müller et al. [61] compared two preparation methods: dispersing nanoclays in glycerol for 24 h and subsequently incorporating this into the starch, versus adding all the ingredients in a single step. While both thermopressed bentonite-starch films presented improved mechanical and barrier properties compared to the control films, the previously dispersed films showed significantly higher reinforcing effects and lower WVP (60% lower than control samples). The procedure led to an enhancement in the interactions between the nanoclays and glycerol, limiting the availability of glycerol and water interactions with starch, reducing hygroscopicity and plasticization. The study demonstrates that the properties of nanocomposite materials are a direct consequence of the interactions that take place between the elements comprising the composites.

Montmorillonite (MMT)

Montmorillonite (MMT) is a relatively cheap and widely available natural clay derived from volcanic ash/rocks [98], popular as a reinforcement in polymer composites due to its high surface area, significant aspect ratio and high swelling properties in polar spaces [99,104]. MMT is comprised of two silica tetrahedral sheets that are interconnected to a central octahedral sheet (2:1 layered structure) composed of aluminum or magnesium, where each layer is about 1 nm thick and with a lateral dimension of 100–1000 nm [105]. The hydrophilic nature of natural MMT, which originates from the presence of sodium cations within its interlayer spaces, makes it compatible with hydrophilic polymers like starch [106], as the polar hydroxyl groups of water in the TPS chain strongly interact with the silicate layer of pristine MMT, allowing the polymer chains to intercalate [107].
Integration of MMT NPs as a filler in starch-based films is frequently used to overcome low mechanical resistance and poor barrier properties [108]. Huang et al. [109] presented corn starch/MMT nanocomposites with remarkable mechanical augmentation through the incorporation of 5% nanoclays, specifically in the tensile strength and strain with an increase of 450% and 20%, respectively. Tang et al. [100] compared the influence of incorporating MMT and modified-MMT nanoclays. The starch/MMT composite films exhibited superior mechanical properties and lower WVP compared to both the starch/modified-MMT and the blank reference. Conversely, the modified MMT presented a hydrophobic surface and was, therefore, less miscible in the hydrophilic starch matrix resulting in little or no intercalation or exfoliation.

Other Nanoclays

Kaolinite is an abundant 1:1 clay with little research into starch composites; moreover, kaolinite-starch interactions are considered very weak, as kaolinite clay sheets present strong hydrogen bonds complicating dispersion in aqueous solutions and polymeric matrices [110]. However, the use of aprotic polar molecules like DMSO is a usual approach to weaken the clay sheets’ bonds and intercalate kaolinite. Mbey et al. [110,111] studied the DMSO intercalation effect on cassava starch/kaolinite films. The DMSO enhanced nanoclay dispersion and, as a result, films presented augmented transparency, water uptake, and UV barrier effect, while lowering Tg. An interesting next step could be the addition of Ag NPs in the interlamellar space of kaolinite with DMSO by chemical-induced reduction [112], which Girdthep et al. [113] used in PLA films to further increase nanoclay dispersion, enhance gas barrier properties, and provide long-lasting antimicrobial properties.
Ren et al. [114] prepared potato TPS/halloysite nanobiocomposites to study the influence of two plasticizers and the effect of nanoclay content. Halloysite nanoclay increased the thermal stability and reduced the moisture content of films independent of the plasticizer used. Glycerol improved halloysite nanotube dispersion in the matrix more than sorbitol, due to stronger hydrogen bonds, allowing a more obvious enhancement in tensile properties.

3.1.2. Metal Nanoparticles

Silver Nanoparticles (Ag NPs)

Looking at the use of metallic nanoreinforcements, a variety of metal nanoparticles have been incorporated in multiple polymeric nanocomposites for a wide range of applications, for example, gold, copper, silver, palladium, and cobalt [115]; however, silver is by far the most popular. In starch matrices silver nanoparticles are an important additive due to their ability to be synthesized via green chemistry [116,117]. This synthesis also presents the added value of being inexpensive, having a straightforward one-step process, being easily scaled up for large-scale synthesis, and not requiring hazardous chemicals, high pressure, energy, or temperature [118]. Moreover, silver is a well-known antibacterial component that can kill over 650 distinct types of microorganisms, including gram-positive and gram-negative bacteria, fungi, and viruses [118]. The concentration of free Ag2+ ions affects the numerous processes that Ag uses. These are the suggested mechanisms:
(1)
Ag2+ interacts with sulfhydryl groups in the cell wall to generate insoluble compounds that are parts of several enzymes involved in the production of transmembrane energy and the transport of electrolytes.
(2)
Ag2+ inhibits the cytochrome oxidase and NADH-succinate-dehydrogenase regions of the bacterial respiratory chain.
(3)
Ag2+ enters the cell and interacts with the bacterial DNA. This reversible binding to the nucleotide bases results in denaturation by breaking the hydrogen bonds between the neighboring purines and pyrimidines [119], as seen in Figure 8.
Ag NPs biosynthesized from plant extracts have demonstrated the capacity to scavenge free radicals in addition to their antibacterial capabilities. This is primarily because phenolic chemicals are present in the structure of Ag NPs [121].
In thermoplastic starch matrices for food packaging Ag nanoparticles have been used as an important additive highly increasing the antimicrobial properties of the matrix, more than essential oils and nano-emulsions. Multiple edible films containing starch as a matrix and essential oil or silver nanoparticles were obtained via solvent casting. The antimicrobial activity of these films was measured proving that edible films containing silver nanoparticles showed almost a three-times-bigger diameter of the inhibition zone [122]. The antioxidant capacity in starch matrices has not been proved yet; nevertheless, Ag Np greatly increases mechanical properties, as explained in a recent study [123] where Ag NP/starch edible films showed a great increase in the physical resistance of films. Khurshid et al. [124] proposed a green approach for the synthesis of Ag NPs from S. asoca leaf extract; the obtained Ag NP/starch films exhibited lower water solubility and WVP with impending antibacterial activity. The addition of Ag NPs to starch and chitosan coatings on paper surfaces upgraded the barrier and mechanical properties and repressed the growth of S. enteritidis and E. coli [125]. Similarly, Ag NPs were dispersed in mushroom starch/agar films to increase antibacterial activities, water vapor properties and hydrophobicity [126].

Zinc Oxide (ZnO) NP

Compared to other metal oxide NPs, ZnO has received the highest attention from researchers. Among its many attributes are biocompatibility, nontoxic, chemically stable under exposure to both high temperatures and UV light, and the recognition by the US Food and Drug Administration (FDA) as a generally recognized as safe substance (GRAS) [127,128].
ZnO NPs’ inclusion in buckwheat starch films enhanced physical and water barrier qualities as well as their antimicrobial activity. The 3% ZnO NPs film in particular showed antibacterial action against L. monocytogenes and preserved the quality of fresh-cut mushrooms throughout storage [129]. Anugrahwidya et al. [130] successfully employed ZnO NPs to increase the tensile strength of starch films and the shelf life of bread. Degradation tests in soil and seawater showed NPs slightly delayed degradation. However, the addition of chitosan to the film matrix resulted in even higher mechanical properties and faster degradation.
In another study, ZnO nanoparticles and fennel essential oil (FEO) were combined on potato starch films produced by the casting method. The study reveals an improvement in barrier properties, a decrease in water vapor and oxygen permeability, and tensile strength. The antimicrobial activity of starch films was tested against S. aureus, E. coli, and A. flavus with results showing a significant improvement as the concentration of ZnO and FEO increased [131].
Ni et al. [132] developed a cornstarch-based coating with 5% ZnO NPs for food packaging paper with high hydrophobicity and antimicrobial properties against E. coli., and with ZnO migration below the overall migration legislative limits; however, ZnO NPs over 7% led to sharply decreased transparency. In a similar manner, mangos were directly coated with starch and ZnO NPs to prevent post-harvest degradation caused by C. gleosporides, prolong shelf life, and preserve the quality of the fruit while being stored [133].

Titanium Dioxide (TiO2) NP

TiO2 is commonly used as a white pigment in the paint and coating industries due to its high refractive index and low absorption, resulting in a high light scattering that renders the materials opaque with outstanding thermal stability [134]. In the packaging sector, TiO2 NPs have been used to modify the properties of different biopolymer films including starch-based nanocomposites [135]. In addition, TiO2 has antimicrobial properties that destroy microbial cells through the production of active oxygen species (ROS) under adequate light energy [136], which can eliminate pathogenic microorganisms, including bacteria, fungi, and viruses. The antimicrobial activity of TiO2 has been widely tested on a variety of microorganisms, including E. coli, S. aureus, P. putida, and L. innocua on multiple surfaces [132,137].
Tunma et al. [138] incorporated 0.01% w/w TiO2 into starch films and managed to extend the shelf life of fresh fruits over petrol-based films. Moreover, the composites decomposed under 14 days leaving no trace of NP residue in the soil. However, higher concentrations of NPs led to brittleness and fracture of the packaging. Ostafinska et al. [139] compared the addition of TiO2 NPs and titanium nanotubes in TPS films using a combination of solvent casting and melt mixing procedures to increase dispersion. Although individual nanotubes were smaller than the TiO2 NP, they tended to form big agglomerates, whereas TiO2 NPs exhibited better dispersion and higher interface surface than agglomerated titanium nanotubes, and as a result, the highest increase in shear moduli.
Arezoo et al. [140] combined cinnamon essential oil (CEO) and TiO2 NPs in sago starch-based film which presented excellent antimicrobial activity against E coli, S. typhimurium, and S. aureus. On the other hand, the additives caused colliding effects in mechanical and barrier properties, a higher TiO2 NP concentration increased tensile strength and Young’s modulus, and lowered elongation at break, WVP, and oxygen permeability, while CEO had the opposite change. The developed films allowed an increase in shelf life and better preservation of fresh pistachios due to the antioxidant and antimicrobial properties of CEO and TiO2 combined.

Copper Oxide (CuO)

Copper oxide nanoparticles do not only act as reinforcements for the polymeric matrix. Bumbudsanpharoke et al. [141] studied blends of PBAT and TPS reinforced with CuO NP, obtained via extrusion using 30% glycerol w/w starch as a plasticizer and different amounts of CuO concentrations (0.05–2 wt%). These nanoreinforced blends present higher mechanical properties (TS increased 29.1% and EB increased 67% over the PBAT/TPS film), increased vapor barrier properties, and an antimicrobial rate of over 99% against E. coli, as well as presenting great antimicrobial activity against bacteria (gram-positive and gram-negative bacteria) [142], and also some antimicrobial activity against fungi [143], viruses [144] and algae [145].
Due to their small size and large specific surface area of CuO, nanoparticles have remarkable antibacterial potential because they can interact intimately with cell membranes [146]. Copper ions function by rapidly receiving and giving electrons, having a high redox potential, and showing the ability to damage the structural elements of microbial cells to the point of death [147].
Numerous studies have reported that edible films containing CuO nanoparticles presented superior properties to other films in terms of their antioxidant qualities. For example, Rehana et al. [148] reported a case used in medicine against cancerous activity. In another case, CuO nanoparticles were used as an additive in a chitosan edible film. Reinforced films presented higher thermal stability than neat chitosan films and had excellent antibacterial activity against gram-positive and gram-negative bacteria, as well as antioxidant properties [149]. In another example, nanoparticles were synthesized by a green method, and their antimicrobial and antioxidant properties were measured showing excellent results in antimicrobial and antioxidant properties [150]. In DPPH tests, CuO nanoparticles can transmit their electron density to the free radical situated at the nitrogen atom [151].
When used in thermoplastic starch matrices, they not only produce excellent antimicrobial [141] response but also antioxidant and mechanical properties are increased when working as additives as well as reinforcements, respectively. In this case, an active edible film containing Ag, ZnO, and CuO nanoparticles, or a combination of these nanoparticles, in a starch matrix has been obtained via the solvent casting method. The addition of this additive has proven to increase mechanical properties as well as antioxidant properties [93].

Other Metallic NPs

The incorporation of magnesium nanoparticles (MgO NPs) in starch films resulted in increased thickness, enhanced antioxidant activities and antimicrobial activity against E. coli and S. aureus, and decreased solubility, moisture content, and WVP [152]. In the same way, MgO NPS and tea polyphenols reinforced potato starch films [153] and exhibited strong antioxidant activity and strong antibacterial activity against E. coli and S. aureus with an inhibition zone diameter of 25.60 mm and 27.50 mm, respectively. Moreover, the lower WVP and oxygen permeability of the reinforced films proved to prolong the shelf life of various fruits, as they reduced weight loss of various fruits, while the reduced enzyme activity and inhibition of the microbial growth on surfaces delayed the decline in firmness.
Zirconium oxide nanoparticles (ZrO2 NPs) combined with encapsulated essential oil increased the antioxidant, antibacterial, and mechanical properties of potato starch and apple peel composite films [154]. The study also highlighted that the augmented films expanded the shelf life of quail meat by reducing WVP and moisture content, whereas incorporating a combination of Ag, CuO, and ZnO NPs in lower concentrations enhanced mechanical and antibacterial properties while decreasing toxicity against eukaryotic cells [93].

3.2. Organic Nanofillers

The exploitation of organic fillers as additives in packaging materials has primarily been motivated by the imperative to mitigate the toxicity concerns associated with synthetic nanoparticles [149], alongside their inherent renewability, sustainability, and abundance. The production of nanoparticles (starch nanoparticles, cellulose nanofibers, chitosan nanoparticles, etc.) in the food industry is gaining attention due to their controlled release, stability, bioavailability, and ability to deliver active ingredients [155]. The combination of biodegradable polymers as a matrix and bio-based filler provides a wide range of opportunities to develop improved, safe, and environmentally friendly packaging materials.

3.2.1. Starch Nanomaterials

These organic nanomaterials possess a multitude of advantageous properties: renewability, biodegradability, unique morphology, high crystalline orientation, surface chemistry, and rheological tendencies, which have attracted significant attention over the past decades [156]. The addition of starch nanofillers has resulted in significant enhancements in the properties of nanocomposites, including mechanical strength, thermal stability, resistance to moisture, oxygen barrier ability, and biodegradation rate [155].
Currently, several terms have been used to describe starch nanomaterials that coexist and have created confusion among researchers. However, there are two distinct classifications of starch nanomaterials persistently reported in the literature: starch nanocrystals (SNC) and starch nanoparticles (SNP) [157]. SNC are crystalline platelets that are derived from starch granules by eliminating the amorphous elements, typically through acid hydrolysis [158,159]. On the other side, SNPs are characterized by their nanoscale size and varying dimensions, which can come from different sources of starch, exhibit variability in crystallinity, and in some cases, be completely amorphous [159,160].
Starch nanocrystals (SNC) constitute crystalline fragments produced through the disintegration of amorphous sectors in starch granules and exhibit nanometer scale dimensions (<100 nm) in at least one dimension. SNC can be obtained from different sources and multiple approaches, with acid hydrolysis with sulfuric acid being the most frequent procedure. In this method [161,162], starch powder is mixed with a 3.16 M H2SO4 solution (15% w/v), at 40 °C under mechanical stirring for 5 days. The remaining suspension is repeatedly washed with distilled water and separated using centrifugation at 10,000 rpm and 4 °C until a neutral pH is reached. Finally, the suspension is submitted to mechanical treatment using an Ultra Turrax to avoid aggregates before being freeze-dried.
During the initial stage of acid hydrolysis (days 0–2), the amorphous regions, which are more susceptible to acid, are quickly degraded resulting in an increase in crystallinity. The second phase (days 2–5) consists of slow hydrolysis of the semi-crystalline rings, breaking the starch granules into nanocrystals [158,163]. Acid hydrolysis procedures were originally carried out using 2.2 M HCl for longer periods of time [164,165], which resulted in non-charge SNC surfaces compared to the negatively charged SNC prepared utilizing H2SO4 [166]. When studying the properties of different SNC, Le Corre et al. [167] discovered that botanical origin had a small influence on nanocrystal size, while amylopectin content and molecular structure had a significant effect on the shape and crystallinity; square-shaped SNC were obtained from A-type starches, while round-shaped SNC were generated from B-type starches, as seen in Figure 9.
Nevertheless, the application of SNC generated through traditional acid hydrolysis is restricted by their susceptibility to aggregation, combined with long elaboration periods and low extraction yields [168]. Moreover, the acid hydrolysis approach poses challenges to its implementation in industrial settings due to environmental concerns [169]. In recent years, alternative methods to obtain starch nanomaterials have been proposed, including physical treatments such as ultrasonication [157,170], stirred media milling [171], gamma radiation [172] and high-pressure homogenization [173,174]; enzymatic treatments [169]; nanoprecipitation in ethanol [175,176]; self-assembly methods; and combinations of methods such as acid hydrolysis with ultrasound [168,177], acid hydrolysis with precipitation [178], ball milling with acid hydrolysis [179], or enzymatic treatments with acid hydrolysis [180].
Starch nanomaterials have successfully been used as reinforcement fillers in many natural and synthetic matrices. The incorporation of SNC into starch-based films increased mechanical and water vapor barrier properties [170] and thermal stability [181]. Similarly, mango seed starch-based films, containing mango SNC at a concentration of 5–7.5%, exhibited increased tensile strength and modulus, as well as decreased water vapor permeability. However, the film’s transparency and elasticity were hindered [182]. Some researchers have proposed modification of SNC to increase the effects of the nanofillers. Surface acetylation of SNC improved their affinity to the matrix, increasing the oxygen barrier and mechanical properties of the films [183]. The addition of modified SNP with citric acid by dry precipitation augmented tensile yield strength (from 3.94 to 8.12 MPa) and Young’s modulus (from 49.8 to 125.1 MPa) while lowering elongation (from 42.2 to 35.0%) and WVP (from 4.76 × 10−10 to 2.72 × 10−10 g m−1 s−1 Pa−1) of starch nanocomposites [184].

3.2.2. Cellulose Nanomaterials

Cellulose is the most abundant natural polysaccharide, composed of linear chains of β-1,4 linked D-glucose units, which form strong intramolecular hydrogen bonds which shape its crystalline characteristics [185]. Cellulose serves as the principal structural component of plants, algae, and bacterial organisms, and is considered a valuable resource that can be obtained from various natural sources such as trees, plant stalks, and agricultural residue [186]. In nature, cellulose exists as an aggregation of distinct cellulose chain-forming fibers known as protofibrils, which subsequently organize into larger units or microfibrils, eventually resulting in the formation of recognizable cellulose fibers [187], as seen in Figure 10. The crystalline structure is upheld through the cohesive interactions of hydrogen bonds and Van der Waals forces, whereas the amorphous structure is characterized by variabilities in the orientation and arrangement of molecular entities, resulting in twisted or contorted structures.
This polysaccharide has received substantial attention in various forms of cellulose nanomaterials (CN), including cellulose nanocrystals (CNC), cellulose nanofibrils (CNF), and bacterial nanocellulose (BNC) [188], as seen in Figure 10 and Figure 11. Unlike starch nanomaterials, complete standard terms and definitions for CN are proposed in ISO Standard TC 20477 [189]. Applications for CN include reinforcing agents in different nanocomposites; medical applications such as bone tissue engineering [190] or drug delivery [191]; wastewater treatment [192]; and cosmetics [193]. CN obtained from different plants and trees exhibit resemblances between them, although minor differences exist. On the contrary, vast differences in particle shape, size, cross-sectional morphology, and crystal structure can be observed when comparing CN extracted from plants, tunicates, algae, or bacteria against each other [194]. Jonoobi et al. [195] reviewed the different CN preparation methods and reported that even with significant variations in the cellulose content of source materials (20–80%), the resulting purified or bleached cellulose exhibited less diversity in terms of cellulose content, with all measurements exceeding 80%.
Figure 10. Nanocellulose production through enzymatic hydrolysis. (A) Pretreatments for cellulose extraction; (B) Controlled enzymatic hydrolysis for production of cellulose nanofibers and nanocrystals; (C) Application of mechanical treatment to obtain more uniform particles [196].
Figure 10. Nanocellulose production through enzymatic hydrolysis. (A) Pretreatments for cellulose extraction; (B) Controlled enzymatic hydrolysis for production of cellulose nanofibers and nanocrystals; (C) Application of mechanical treatment to obtain more uniform particles [196].
Polymers 15 02972 g010
Figure 11. Comparison between the different types of nanocellulose [197].
Figure 11. Comparison between the different types of nanocellulose [197].
Polymers 15 02972 g011

Cellulose Nanocrystals (CNC)

Cellulose Nanocrystals (CNC), also called nanowhiskers [198], are a versatile class of natural polymer materials with high strength, low density, large specific surface area, high aspect ratio, and hydrophilicity. Moreover, CNC preserve desirable natural cellulose properties like biocompatibility and renewable degradation, with the surface and size effects of nanomaterials [199]. The ISO Standard defines CNC as ‘composed predominantly of cellulose-containing crystalline and paracrystalline regions, with at least one elementary fibril not exhibiting longitudinal splits’, with an ‘aspect ratio (longest to smallest dimension) usually smaller than 50 but usually bigger than 5′ and includes as similarly used terms ‘nanocrystalline cellulose, nanowhiskers cellulose and cellulose crystallites and cellulose microcrystals’ [189].
Similar to starch, cellulose is commonly treated by acid hydrolysis to form CNC, but through a two-step procedure. First, a pre-treatment of the source material is needed to eliminate the hemicellulose, lignin, and other components of the matrix, normally through chemical treatment (pulping) or steam explosion [200]. In the second phase, a hydrolysis reaction is used to destroy the cellulose amorphous fibers and liberate the crystalline region. Usual hydrolysis conditions include a 64–65% (w/w) sulfuric acid aqueous solution with the ratio of 1:8 to 1:20 (w/v) of raw material containing cellulose vs. acid at 45 °C, followed by centrifugation washing to a neutral pH, homogenization, and freeze-drying [201], as seen Figure 12. The diversity of dimensions, aspect ratio, morphologies, and degree of crystallinity of CNC are defined by the type of cellulosic source, the conditions during the preparation process, and the experimental methodology employed [187].
The addition of small amounts of wood CNC (1.5–2.5%) to extruded native potato starch formed strong hydrogen-hydroxyl bonds, which improved the mechanical properties while reducing swelling and enzymatic degradation [202]. In contrast, when CNC content exceeded 5 wt%, nanocrystals aggregated, ultimately resulting in a negative impact on physiological conditions [202,203]. The issue of CNC aggregation often arises due to the intermolecular hydrogen bonding interactions that occur between the hydroxyl (OH) groups present on the surface of the nanocrystals [204]. Similarly, beeswax CNC/cornstarch films used to preserve the freshness of fruits displayed oxygen and water barrier capacity and improved tensile strength and modulus [205].
The use of cellulose nanocrystals (CNC) can be impeded by their intrinsic tendency to self-aggregate and hydrophilic behavior. Consequently, several chemical modifications and copolymerization have been documented as strategies to address these deficiencies [206], and to grant additional functions. Some examples include surface modification of CNC through vinyl silane grafting [207]; modified CNC by silane coupling agent and graphene oxide to mechanically reinforce corn starch films [208]; crosslinking CNC by citric acid and amidation of chitosan, which resulted in improved nanocomposites with increased tensile strength, lower water absorption and WVP and good antimicrobial activities against E. coli and S. aureus [204].

Cellulose Nanofibers (CNF)

Cellulose nanofibers (CNF) consist of cellulose chains that are laterally stabilized by strong hydrogen bonds that form between the oxygen and hydroxyl groups of neighboring molecules [195]. The ISO Standard defines CNF as ‘composed of at least one elementary fibril that can contain branches a significant fraction of which are in the nanoscale’, with ‘dimensions of 3–100 nm in cross-section and typically up to 110 µm’, and includes as similarly used terms ‘nanofibrillated cellulose, nanofibrillar cellulose and cellulose nanofibers’ [189]. The diameter of CNF, which can range from approximately 2–20 nm in diameter and extend several microns in length, is primarily influenced by their origin [195]. In contrast to CNC, CNF possess both crystalline and amorphous regions and are generally obtained through a sequence of high-shear mechanical processes of the cellulosic source, mainly high-pressure homogenization, microfluidization, or grinding [156].
Production of CNF using mechanical processes is very energy-intensive; however, pretreatment methods offer an efficient approach to reduce energy input [209,210]. Examples of pretreatment methods include enzymatic treatments [211,212], carboxymethylation [213] and TEMPO oxidation [214]. Pretreatment methods also have a great influence on CNF dimensions; current technology produces CNF with 4–10 nm diameter, while untreated CNF yield coarser fibrils, normally 20–100 nm in diameter, known as cellulose microfibers (CMF) [194]. The downside of these procedures is their relatively high cost and possible unwanted side effects, such as the shortening of CNF length [215]. Moreover, CNF have traditionally been extracted from wood products, but efforts to reduce production costs and increase biomass sustainability have seen several studies detailing the isolation of CNF from alternative sources including agricultural residues and industrial wastes [216,217,218,219].
Nanocomposite films based on potato starch and pineapple leaf CNF (3% w/w) proved to be UV resistant and have increased water barrier properties [219]. Introducing only 0.4 wt% of henequen CNF in TPS films was enough to increase tensile strength (+80%) and elastic modulus (+170%), and lower WVP (−86%) and oxygen permeability (−94%) [220]. Similarly, extruded starch/wood-CNF composites presented increased mechanical properties and better water sensitivity compared to neat TPS samples, while preserving their translucency [221]. Additionally, all three studies reported that higher CNF concentrations resulted in the aggregation of fibers and lower reinforcement effects.
Chen et al. [222] extracted three kinds of non-wood CNF with the same production method and compared the reinforcement effects on starch-based films. The botanical source had almost no impact on morphology and chemical composition but influenced the size and crystallinity of CNF. The addition of CNF with the highest aspect ratio provided superior mechanical and barrier properties. On a similar note, Zhan et al. [223] compared the effect of adding CNC and CNF to pumpkin starch-based films, revealing CNC provided better reinforcing capacities while CNF showed higher thermal stability. The combination of CNC with clove bud oil (CBO) proved to be an effective additive in starch-based films to preserve red grapes, with CBO further increasing mechanical properties, lowering WVP, and providing antimicrobial properties to the CNC-enhanced films [224].
Biodegradability tests of several nanoparticles in an aqueous environment [225] showed that cellulose and starch NPs exhibited a faster rate of degradation compared to their macroscopic counterparts. Conversely, other nanoparticles with significant and widespread use, such as fullerenes and carbon nanotubes, displayed no detectable biodegradability.

Bacterial Cellulose (BC)

In addition to plant cellulose, the production of cellulose can be achieved by means of diverse enzymatic mechanisms, and chemical synthesis from glucose-derived compounds, as well as biosynthesis facilitated by algae, fungi, and bacteria [226]. The cellulose produced by bacterial strains possesses an identical molecular formula and polymeric structure to plant cellulose, with higher levels of crystallinity, and free of lignin, hemicellulose, and pectin [227]. As a result, bacterial cellulose purification and production are easier as there is no need for harsh chemical or high-energy treatments [228]. Moreover, microorganisms can be genetically adjusted to deliver BC with the desired properties and agro-industrial wastes can be utilized as a growth medium to lower production costs [229]. Nevertheless, carbon sources and culture media can have effects on the production and structural properties of bacterial cellulose [230], similar to how the botanical source affects the properties of plant cellulose.
BC is composed of a network of ultrafine nanoscale fibers with extensive polymerization and possesses remarkable physicochemical and mechanical characteristics, comprising high levels of purity, crystallinity, porosity, surface area, translucency, chemical stability, tensile robustness, proton conductivity, and water retention capability [228,231]. Moreover, toxicology studies strongly suggest the use of BC does not pose any adverse side effects when used in human foods [227]. Consequently, BC has been utilized in multiple areas including wound dressing and blood vessel regeneration in the medical field, as biomaterial in electrical instruments, paper restoration, and food ingredients, and as a reinforcement in food packaging [232].
Melt-mixed corn TPS films containing a 15% bacterial CNC loading prepared by Fabra et al. [233] enhanced oxygen (−95%) and WVP (−46%) barrier properties, preserved transparency, and stiffened the material, with elongation at break quickly decreasing with high loadings. Grande et al. [234] developed bacterial cellulose NF/starch nanocomposites by self-assembly. Unlike previously mentioned methods where cellulose is disintegrated to obtain BCNF, self-assembly starch films capitalize on the inherent ability of certain bacteria to expel cellulose nanofibers and the transport mechanisms of starch gelatinization to form the nanocomposite film, as seen in Figure 13.

3.2.3. Chitin/Chitosan Nanoparticles

Chitin is an abundant biopolymer that originates naturally in exoskeletons of arthropods, cuticles of insects, and in certain fungi and seaweed algae, primarily assuming a supporting function [235]. Chitin nanocrystals can be produced through acid hydrolysis, oxidation (TEMPO-mediated or ammonium persulfate (APS)) and mechanical treatments, while nanofibers can be obtained by a simple grinding process [236].
Chitosan is a partially or completed deacetylated derivative of chitin, resulting in a linear polysaccharide consisting of (1,4)-linked 2-amino-deoxy-b-D-glucan, as seen in Figure 14. Chitosan offers multiple valuable properties for nanobiocomposite development due to its non-toxic, biodegradable, biofunctional, and biocompatible features, coupled with antimicrobial properties [236,237]. Chitosan nanoparticles can be obtained through various methods including ionic gelation, reverse emulsion, precipitation, and polyelectrolyte complexation [238]. The most promising method for food packaging and pharmaceutical application is ionotropic gelatinization, a physical crosslinking between tripolyphosphate (TPP) and protonated chitosan, that avoids the potential toxicity and undesired effects of chemical crosslink [239,240].
The chemical similarity between the starch matrix and chitin/chitosan NPs creates a strong interfacial interaction, resulting in a reinforcement effect, visible through an increase in tensile strength and decreased elongation at break; an improvement in the intermolecular interaction increasing thermal stability as Tg shifts to higher temperatures; and the creation of a more tortuous path that lowers WVP [241,242]. Additionally, Shapi’i et al. [238] reported that chitosan NP/starch films were more efficient at inhibiting the growth of gram-positive (B. cereus and S. aureus) bacteria compared to gram-negative (E. coli and S. typhi) bacteria; a similar conclusion was reached on chitosan NP/tara gum edible films [243].

4. Additives

As mentioned before, the properties of materials based on a thermoplastic starch matrix, despite their many benefits, have some well-known drawbacks. Therefore, it is necessary to incorporate additives to improve the material’s intrinsic characteristics or even introduce some new functional properties, such as antimicrobial or antioxidant activity or even some mechanism that can detect the state of degradation of the material [244]. The classification of these additives can be organized into three main groups: plasticizers, essential oils (EOs), and organic molecules.

4.1. Plasticizers

A plasticizer is a polymer additive, with a low molecular weight and non-volatile nature, that is added to the polymeric matrix to provide elasticity to the base polymer [245]. Plasticizers increase film flexibility due to their ability to reduce internal hydrogen bonds between polymer chains while increasing molecular space, lowering the glass transition temperature, and increasing the flexibility of the material [246,247]. Moreover, the properties of films can be influenced by the specific plasticizer utilized, as they present varying origins, molecular weights, compositions, and the quantity used, which is normally around 10–40% of dry weight depending on the stiffness of the polymer used [248]. Water is an excellent starch plasticizer but, when removed, films become brittle; therefore, other plasticizers are needed [42]. Common plasticizers in polymeric films include sugars (glucose and sucrose), polyols (glycerol, PEGs, sorbitol, and mannitol), and lipids (phospholipids and fatty acids) [13].
The introduction of plasticizers together with elevated temperatures and shear increases the flexibility of native starches similar to conventional thermoplastic polymers [249]. The most effective plasticizers often resemble the structure of the polymer they plasticize, so the most common plasticizers used in starch-based films are polyols, such as sorbitol and glycerol [247]. During plasticization, polar groups in polyols interact with the hydroxyl groups of starch molecules and reduce internal starch hydrogen bonding. Meanwhile, non-polar groups weaken the van der Waals forces between starch molecules and improve the mobility of starch chains, consequently increasing the tensile properties in films [250]. Consequently, a higher polyol concentration translates into thicker films with lower density, as they increase the disruption and restructuration of the intermolecular polymer chain network creating more free volume [249].
In packaging applications, glycerol is the most utilized plasticizer in starch-based films, due to its stability and compatibility with the hydrophilic bio-polymeric packaging chain [251]. The efficacy of glycerol to generate TPS biodegradable films is largely attributed to its diminutive scale, which grants it enhanced intercalation capabilities between the polymeric chains, and thus confers greater influence over the resulting mechanical properties [252]. Increasing glycerol content has been related to an increase in moisture content, water solubility, water absorption, water vapor permeability, density, elongation at break, lightness, and thickness, while tensile strength and color parameters decreased [253]. As well as a plasticizer for bio-based polymers, this polyol plays a crucial role in another polymer field, particularly as a feedstock to produce synthetic polymers (i.e., polyurethane and polyester) [254]. Interestingly, glycerol is the major by-product of the biodiesel production line, with each cycle of processing yielding around 10% (w/w) of waste glycerol [254]. Therefore, from a circular economy point of view, consuming glycerol for film production purposes would further promote the sustainability of biodiesel and packaging production.
When comparing glycerol and sorbitol, starch-glycerol films presented lower tensile strength and elastic modulus values and higher elongation compared to starch-sorbitol films, indicating the pronounced plasticizing influence of glycerol [255,256]. Additionally, sorbitol films presented higher thermal stability, as the higher molecular weight of sorbitol decreased the molecular mobility of the polymer chains [256].
Similarly, Maniglia et al. [257] studied the effect of plasticizers on starch-based films from starch obtained through different extraction techniques. Plasticizers with lower molecular weight such as glycerol and urea yielded starch films with higher solubility in water, greater hydrophilicity, higher WVP, and less crystalline structure. On the other hand, plasticizers with a greater number of -OH groups such as sorbitol and glucose were more effective for starches with higher amylose, fiber, and lipid contents and less agglomerated structures.
Ma et al. [250] compared sweet potato starch films with mannitol and sorbitol in different proportions. With the increase in the percentage of sorbitol, more hydrogen bonds formed between the plasticizer and starch molecules, decreasing the tensile strength of starch films. On the other side, higher mannitol content translated to lower freedom of water molecules, limiting the mobility of starch molecules, and thereby resulting in a reduction in water loss and starch regeneration. Finally, the starch film with mannitol:sorbitol (60:40) presented the highest tensile strength, the lowest WVP and was the most effective in delaying retrogradation.

4.2. Essential Oils (EOs)

Due to the many characteristics of these additives, essential oils have recently captured the interest of many researchers in the field. In general, essential oils come from a variety of aromatic plant parts, including seeds, bark, flowers, peel, fruit, roots, leaves, and wood. Their biological actions are caused by complex combinations of more than a hundred molecules, such as terpenes and isoprenoids, which are examples of volatile compounds and lipophilic compounds [258]. The EOs are extracted using different techniques such as mechanical extraction, dry distillation, or water hydro distillation that do not include temperature changes [259].
Additionally, EOs have the ability to neutralize a variety of bacteria [260], as well as having antioxidant properties [261]. These oils are particularly interesting additives for their application in thermoplastic starch matrices since they are non-toxic. However, they show some significant drawbacks when used in combination with TPS films. In fact, they tend to decay due to their volatile, unstable composition, limited water solubility, and exposure to light, oxidation, and heat from the outside, among other factors [262]. In order to solve these kinds of issues, encapsulation in polymeric particles [263], liposomes [264] and solid lipid nanoparticles [265] are interesting solutions recently reported in the scientific literature.
  • Terpenes: With more than 40,000 distinct chemical structures, terpenes are one of the principal constituents of essential oils and one of the most abundant and diverse categories of plant natural products [266]. There are at least two possible ways to biosynthesize terpenes: from mevalonic acid or from metileritritol phosphate [267]. However, all terpenes are derivates of the isoprene structure (Figure 15).
Figure 15. Terpene structure.
Figure 15. Terpene structure.
Polymers 15 02972 g015
Hemiterpenes (C5), monoterpenes (C10), sesquiterpenes (C15), diterpenes (C20), sesterpenes (C25), triterpenes (C30), and tetraterpenes/carotenoids (C40) are the different categories of terpenes based on the number of times this structure is repeated [268].
  • Polyphenols: These molecules can be divided into different classes depending on their chemical structure. Polyphenols are found in nature as a product of the secondary metabolism of most plants [269]. Two main classes of phenolic acids can be distinguished depending on their structure. Both of them have benzene as a basic bond to a carboxylic group (benzoic acids) or to a propionic group (cinnamic acids), and both structures can be found with different hydroxylation levels [270]. In addition, some examples of molecules inside this group that are not phenolic acids are: coumarins, lignans, chalcones, flavonoids, lignins, and stilbenes [271].
EOs’ constitutions, and therefore their method of action, determine their antibacterial activity [272]. The bacteriostatic or bactericide impact of essential oils is related to their potency [273]. In this regard, the EOs’ efficacy against gram-positive and gram-negative bacteria is very diverse, except for cinnamon, clove, and oregano. The phospholipid bilayer, the membrane proteins, and the degradation of the cell wall by essential oils can enhance the permeability of the cell membrane and cause the loss of cellular components [274].

4.2.1. Rosemary Essential Oil (REO)

Rosemary essential oil (REO) has been widely studied in various matrices, such as PLA. Fiore et al. [275] successfully developed five different biopolymers with the addition of chitosan-based active coatings in which the antioxidant effects of REO delayed the lipid oxidation of raw chicken meat, in comparison to the control sample. In another study [276], active nanocomposite films based on a sodium caseinate matrix were synthesized via the casting method. The films reinforced with ZnO NPs and REO showed better mechanical, physical and barrier properties than the blank films, as well as antimicrobial properties.
Bio-chitosan films were produced via the solvent casting method, using montmorillonite as reinforcement and REO as well as ginger essential oil to add some antimicrobial and antioxidant behavior to the matrix [277]. The measurement of the film properties proved the antioxidant capacity of the essential oils, since the active compounds contained in the chitosan film continued to have the capacity to scavenge the DPPH radical even after casting and migration experiments. Additionally, the films also showed antimicrobial activity against five foodborne pathogens [231].
REO has also been used as an additive in starch-based films. In particular, Restrepo et al. [278] formed edible films using starch as the polymer matrix, enhancing its properties with rosemary and lemongrass essential oils by forming nano-emulsions due to the essential oil’s low water solubility. Edible films were obtained using the solvent casting method, glycerol as a plasticizer (30% w/w starch) and the EOs’ nanoemulsion concentrations (0.3%, 0.5%, and 1% w/v). As a result, it was observed that these essential oils caused a plasticizing effect on the matrix that allowed higher transparency and elongation at break.
Moreover, REO has been proven to have high antimicrobial activity against gram-positive bacteria (S. epidermidis, S. aureus, and B. subtilis), gram-negative bacteria (P. vulgaris, P. aeruginosa, and E. coli), and fungi (C. albicans and A. niger) [279] and to increase antioxidant activity on films coated with REO [275]. The primary components of REO, which constitute 97.41% of the entire oil, are 1,8-Cineole (26.54%), α-Pinene (20.14%), Camphor (12.88%), Camphene (11.38%), and α-Pinene (6.95%), and, according to one hypothesis, the antibacterial properties are a result of their combination [280].
Due to all these antimicrobial characteristics, REO has been one of the most used in starch matrices. In particular, Tafa et al. [281] obtained a nanocomposite of cassava starch, silica nanoparticles, and REO extracted from rosemary leaves, that was prepared to improve the antimicrobial activity, physicochemical properties and water-absorbing and holding properties of starch matrix.
As mentioned before, REO can act as a plasticizer, allowing for greater transparency and elongation at break, both of which are desired in the food sector and raise water vapor permeation values. However, this impact appears to be offset by the EOs’ hydrophobic properties [278].

4.2.2. Cinnamon Essential Oil (CEO)

Cinnamon essential oil (CEO) is one of the most significant and promising EOs being used in active food packaging due to its antibacterial and antioxidant capabilities. In general, cinnamaldehyde has been primarily involved in CEO’s excellent antibacterial and antioxidant activities. CEO is considered one of the most common EOs in starch matrices for packaging applications because of its great antioxidant properties. In particular, the addition of CEO reduces the moisture content of the films and the water solubility due to its hydrophobic behavior and provides good thermal stability and antioxidant capacity as shown in a recent article where films were obtained via solvent casting, using glycerol as a plasticizer (25% w/w starch) and varying the bioactive compound [282]. In another study, the addition of CEO in a matrix of starch and sugar palm-based cellulose was studied [283], where an increase in tensile strength and tensile modulus was observed due to the group interactions in the films. Also, the addition of CEO had a plasticization effect on the matrix with a significant increase in the elongation at break and improved ductility of the films [284].
Phenylpropanoid cinnamaldehyde is the primary component of CEO, and its values range from 49.8% to 91% [285,286], followed by Cerpene cinnamyl-acetate [287,288]. Propanol, (+) 3-Carene, Copaene, Eugenol, and α-Pinene are further minor chemicals. Regarding the antimicrobial mechanism, the active ingredients of CEO can go through bacterial membranes and exert their bactericidal effect by acting on a variety of cellular targets. Additionally, they also act on cell physiology causing clumps and self-aggregation on the bacterial membrane [289].

4.2.3. Thyme Essential Oil (TEO)

Thyme essential oil (TEO) is also one of the most promising and widely used essential oils, due to the number of residues that can be used to obtain it and its antibacterial properties against all types of bacteria. TEO has been used in direct contact with foods such as lettuce, meat and milk, showing good results against pathogens of food origin, but less efficacy and less antioxidant activity in food matrices than by itself [290].
Nieto et al. [291], studied the antioxidant capacity of TEO using liposomes as biological membrane model system. To achieve this, after extracting and characterizing the extracted essential oil, they mixed it with a previously prepared liposome solution where peroxidation was induced by adding AAPH in a sodium phosphate buffer (pH 7.4). The reaction was monitored by observing the absorbance at 234 nm, which is the maximum absorption of the conjugated dienes. They obtained promising results on the ability to inhibit the formation of conjugated dienes in a liposome system.
TEO is widely studied as an additive in biodegradable polymers for food packaging, including starch matrices, due to its non-toxic, antimicrobial and antioxidant nature. Cai et al. [292] studied the physicochemical characteristics of a starch film containing TEO microcapsules. Edible TEO films were prepared by coprecipitation technology. In particular, in this process, 30% by weight of glycerol was added as a plasticizer. Samples with different amounts of essential oils, 10%, 20% and 30% by weight of starch and 10% by weight of citric acid, were analyzed. As a result of the study, it was found that the addition of TEO had the ability to inhibit Botryodiplodia theobromae Pat and Colletotrichum gloeosporioides Penz on the surface of mango fruit. In another study, a starch matrix nanocomposite activated with TEO was observed to improve the shelf life of vegetables, in this case baby spinach leaves. For this, starch/clay nanocomposite films (montmorillonite) and TEO were prepared by the solvent casting method and used as packaging for spinach leaves. As a result, the film could improve the acceptability of leaves by around 8 days by controlling the growth of bacteria such as E. coli and S. Typhimurium and improving the microbiological quality of baby spinach leaves during refrigerated storage. [293]. However, as the addition of TEO does not affect the mechanical properties, it is usually used with additional reinforcements to enhance mechanical properties [294].
TEO is formed by 52 distinct constituents that have been identified. Moreover, it was discovered that thymol and carvacrol were the 2 most prevalent chemicals along with terpineol, linalool, p-cymene, and other substances [295]. Due to thymol being the primary component of TEO, its antibacterial activity is closely tied to its mode of action. The structural and functional damage to the cell membrane as well as the increase in its permeability are connected to the mechanism of action of thymol. This activity is connected to the loss of inorganic ions, the production of nucleic acids, the decline in membrane potential, protein dysfunction, the collapse of the proton pump, and the depletion of ATP. Before entering the phospholipid bilayer of the cell membrane, it binds to the membrane. Moreover, thymol makes the plasma membrane more permeable to protons and ions, degrades lipid and protein layers, causes the cytoplasm to coagulate and intracellular component leakage, interferes with enzyme systems and cellular metabolism, and finally leads to cell death [296].
Aldehydes, ketones, and carboxylic acids are produced at the end of the polymer chain by photooxidation. Antioxidants stop these chain reactions and prevent subsequent oxidation processes by eliminating the free radical intermediates [151]. To perform its antioxidant function, thymol donates electrons to the first hydrogen atom that initiates the oxidation chain [297], as seen summarized in Figure 16.

4.3. Organic Molecules

The majority of the components included in this category, such as carvacrol, ferulic acid, and cinnamic acid, can also be found in essential oils, although their usage as single molecules is different [298] or even combined between them [299]. Without the other components that are part of the essential oils, the characteristics of the films that can be obtained directly using the simple molecules can be controlled in a more focused and meaningful way than with the use of the essential oils.
The number of procedures required for the purification of such components is its principal drawback. These molecules are often derived from the essential oils collected from plant resources; as a result, the purifying process entails an increase in overall cost and difficulty.

4.3.1. Carvacrol

Carvacrol (Figure 17), a phenol, is the primary component of oregano and thyme essential oils, with a range of 5 to 50% for thyme [295] and between 18–71% in the case of oregano [300,301]. These variances can be due to the different extraction techniques and the parts of the plant that were utilized to extract the molecules.
The antimicrobial activity of carvacrol is well known and studied. Carvacrol has free phenolic hydroxyl groups that can form hydrogen bridges and release protons that can damage the membrane structure and lead to bacterial cell death [302].
Eliuz et al. [303] exposed the carvacrol present in essential oil from Thymbra spicata leaves to radiation (at a wavelength of λ > 200 nm), which caused the OH groups to break apart and the H atom released in the orto- or para-position to recombine. However, as different types of cell walls have distinct reactions to UV light, the activity varies between cell walls.
In starch matrices, enhancement of antimicrobial and antioxidant properties via the addition of carvacrol has been widely reported. Edible starch films were obtained using carvacrol and thymol as a method of reducing the incidence of anthracnose symptoms in mango and papaya, enhancing fruit quality as long as possible by an additive effect of both molecules [304]. Mao et al. [305] reported that carvacrol nano-emulsion caused the formation of a porous structure and an increase in the thickness of the film. Moreover, the incorporation of carvacrol decreased the mechanical properties of the nanocomposites, caused by the interactions between carvacrol and the starch matrix via hydrogen bonding.

4.3.2. Eugenol

The primary ingredient in some essential oils, such as clove essential oil, which has a phenol content ranging from 76% to 82%, is eugenol (Figure 18) [306,307].
The free hydroxyl groups of this molecule are responsible for most of its antibacterial action, as it can damage cell membranes, increasing nonspecific permeability, which disrupts intercellular ATP and ion transport [308]. Moreover, it has a double bond in the side chain, in the α,β positions, and a methyl group in the position located in the γ position [309]. Due to its hydrophobic properties, it can penetrate the lipopolysaccharide of gram-negative bacteria’s plasma membrane and change the cell structure, which causes intracellular components to infiltrate [310].
The addition of eugenol in starch matrices has been reported in several articles to increase antimicrobial properties. In a study [311], eugenol was encapsulated into 3D-porous starch micro-networks to improve the antimicrobial properties of a corn starch matrix. This system displayed effective antibacterial activities against E. coli, S. aureus and B. subtilis and improved mechanical properties. In a second example [312], eugenol gelatin microspheres showed significantly increased antioxidant capacity and good antibacterial activity.
In relation to the mechanical properties response, these systems show a big dependence on the plasticizer due to the possible hydrogen bond interactions between eugenol and the plasticizer. In particular, Cheng et al. [313] reported that the hydrogen bonds formed between starch and sorbitol showed increased mechanical properties as sorbitol was added into the matrix as well as increased barrier capacity against moisture and oxygen.

4.3.3. Cinnamic Acids and Derivates

Cinnamic acids (CA) and their derivatives (Figure 19) are a wide family of molecules that are usually a minority compound in various types of plants [314,315,316].
The exact molecular mechanisms by which CA acts are unknown. However, two possible mechanisms have been proposed: CA blocks the post-translational modification of proteins that control cell growth, such as some p21ras proteins, thus preventing protein prenylation by blocking the synthesis of residues derived from mevalonate, or CA inhibits the expression of genes that promote cell proliferation [317]. CA exhibits antibacterial efficacy against all gram-negative bacteria, but not against gram-positive bacteria [318].
Moreover, the antimicrobial activity of CA in starch matrices has been reported. In particular, Ordoñez et al. [319] incorporated cinnamic and ferulic acid into starch monolayer films. These films were obtained via the solvent casting method using glycerol as the plasticizer in a ratio of 0.30 g/g starch and 1 and 2% w/w of ferulic and cinnamic acid, respectively. It was only observed that antimicrobial inhibition in films containing 2% acid was more effective than the ones with just cinnamic acid. There are still no studies that imply any antioxidant effect or modification of mechanical properties by the addition of CA in food packaging applications.

4.3.4. Organic Molecules in Smart Packaging

A new application in food packaging that has been gaining interest from researchers is food safety and quality monitoring, known as smart or real-time packaging. There are three main types of smart packaging, including indicators, data carriers, and sensors. In this regard, food packaging containing an indicator film can detect changes in the quality or freshness of the product, and respond by providing a visual warning to consumers. Briefly summarized, food spoilage can be tracked by alterations in pH as a reaction to the growth and metabolism of microorganisms or the development of organic acids and/or volatile amines (e.g., lactic acid formation in milk or ammonia/amino-sugar formation in meat) [320]. Commonly used synthetic indicators, such as bromothymol blue, methyl red and xylenol dyes, are known to be toxic and cancerogenic to human health [321]. However, natural pH-responsive color-changing indicators offer a simple, non-invasive, non-destructive, inexpensive, and fast-reacting quality-tracking method [321]. Many studies described the preparation of smart packaging films for monitoring food quality by combining biological macromolecules (e.g., starch, cellulose, chitin) with natural colorimetric indicators (chlorophyll, carotenoids, curcumin, betalains, and anthocyanins) [322,323,324].
In terms of natural pH indicators in starch matrices, anthocyanins are by far the most studied, as seen in Table 3. These phenolic compounds can be extracted from a wide variety of fruits and vegetables including red grape, purple corn, black rice, or different kinds of berries [325]. Moreover, anthocyanins are pH-sensitive substances that exhibit a gradient color change for different pH values. Anthocyanins undergo a continuous structural change from flavonoid ions under acidic conditions to quinone structures under alkaline conditions, as seen in Figure 20 [326]. Additionally, anthocyanins also display potent antioxidant activity, which can be useful to prevent color change or lipid oxidation of packaged food products [327].
Figure 20. Color changes in response to pH variation of (A) anthocyanin, (B) curcumin, (C) alizarin, (D) shikonin, and (E) betalains [321].
Figure 20. Color changes in response to pH variation of (A) anthocyanin, (B) curcumin, (C) alizarin, (D) shikonin, and (E) betalains [321].
Polymers 15 02972 g020
Fotouhi et al. [328] developed smart starch/gelatin biocomposite films that determined milk freshness and the presence of some preservatives, providing an economical, rapid, and biocompatible diagnostic. The films combined the use of color-changeable films embedded with small anthocyanin, ferric ion, iodide, and diphenyl carbazide analytes with QR codes. Interestingly, Qin et al. [329] studied the effect of encapsulating anthocyanins in starch/PVA blends. When compared to films with unrestricted anthocyanins, the encapsulated films presented denser micro-structures, stronger intermolecular interactions, higher water vapor blocking and mechanical properties, and higher stability when exposed to heat (35 °C) and light radiation.
Table 3. Natural pH indicators in starch matrices.
Table 3. Natural pH indicators in starch matrices.
MatrixIndicatorBotanical SourcePreparation MethodFood TestedColor ChangeRef.
cassava starchanthocyaningrape skinextrusionmeatblack to purple[330]
starch + PVAanthocyaninroselle calycessolvent castingfishpink to blue
to yellow
[331]
agar/potato starchanthocyaninpurple sweet
potato
-meatred to green[332]
starch + gelatinanthocyaninpurple sweet
potato
solvent castingmushroomsgreen to gray
to yellow
[333]
native/hydrolyzed
cassava starch
anthocyaninyerba mateextrusion-white to
yellow/brown
[334]
sago starchanthocyaninred cabbagesolvent casting-pink to green
to yellow
[320]
starch + gelatinanthocyanin
+ [KI/DPC/Fe(III)]
red cabbagesolvent castingmilkpink to blue
to yellow/green
[328]
cassava starch
+ PVA
encapsulated
anthocyanin
black fruit
wolfberry
solvent castingfishpink to blue
to yellow/green
[329]
cassava starch
+ PVA
betaninbeetroot redsolvent casting-red brown to
brownish yellow
[335]
potato starchbetacyaninpaperflowersolvent castingfishpink to yellow[336]
corn starch
+ PVA
curcumin-microwave-ultrasound
reaction
fishyellow to red[337]
TPS + PEcurcumin-extrusionmeatlight yellow to
light brown
[338]
corn starchshikoninzicao rootssolvent castingshrimpsred to blue[339]
Besides anthocyanin, curcumin has also been proposed as a pH indicator for smart packaging applications. While curcumin does not possess the wide color range of anthocyanin, it presents a clearer reaction to alkaline environments [337]. Curcumin is a hydrophobic polyphenolic compound, derived from the root of Curcuma longa, and is yellow-orange in its natural state [340]. Liu et al. [337] presented starch/PVA based films incorporated with curcumin-loaded Pickering emulsions that presented antioxidant and antimicrobial properties against S. aureus, B. subtilis, and E. coli.
In regards to film elaboration, the use of solvent casting and extrusion techniques has been frequently reported, as seen in Table 3. However, the addition of natural indicators does not solve the mechanical, barrier and processability weaknesses of starch-based films. Consequently, many researchers have often resorted to blends and nanofillers to further enhance film properties for food packaging applications.

5. Conclusions and Perspectives

In recent years, biopolymeric materials have been gaining popularity as an alternative to traditional plastic packaging. Moreover, stricter regulations on plastic packaging and the recent interest in the revalorization of agro-alimentary and industrial waste may increase the demand for biopackaging in the near future. Among biopolymers, starch is an abundant, sustainable, and worldwide-available source with interesting properties, that possesses plastic-like behavior in its thermoplastic form. However, the poor mechanical properties and hydrophilic nature make the processed material vulnerable and of poorer quality than its synthetic counterpart.
The integration of nanofillers into starch-based films has been studied and found to enhance several of their physical characteristics, including mechanical properties, thermal stability, moisture resistance, oxygen barrier capabilities, and biodegradation rate. Nevertheless, excessive nanofiller content can lead to negative effects due to the agglomeration and formation of rigid regions within the polymer matrix, and the consequent decline in the interfacial surface area. On the other hand, incorporating additives, plasticizers, essential oils, and organic molecules into the starch films offers interesting functional properties such as antimicrobial, antioxidant or pH-indicative properties. Nevertheless, the addition of nanofillers in nanocomposites is not limited to one element. Multiple research papers have proposed a combination of reinforcement and additives to obtain the benefits of each element while minimizing the downside effects. In summary, from a circular economy point of view, the key point is that thermoplastic starch biocomposites can challenge synthetic packaging when modified with adequate additives and nanofillers.
Therefore, the objective of this review is to discuss starch-based films for food packaging applications. In particular, after analyzing different starch botanical sources and their extraction techniques, the properties of thermoplastic starch are considered, including the main preparation techniques such as solvent casting, extrusion, and film blowing. Then, the most typical organic and inorganic nanofillers used as reinforcements are described, detailing their effects on the mechanical response. Furthermore, additives, plasticizers, essential oils, and functional organic molecules are reported to focus on their antioxidant and/or antimicrobial properties and their application in smart packaging.

Author Contributions

Conceptualization, D.L. and L.P.; resources, D.L. and L.P.; writing—original draft preparation, P.F.M.-G. and V.O.-C.; writing—review and editing, D.L. and L.P.; funding acquisition, D.L. and L.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Agencia Estatal de Investigación (AEI, MICINN, Spain) and Fondo Europeo de Desarrollo Regional (FEDER, EU), PID2021-123753NB-C31 and NextGenerationEU MRR TED2021-129335B-C21.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Filho, W.L.; Salvia, A.L.; Bonoli, A.; Saari, U.A.; Voronova, V.; Klõga, M.; Kumbhar, S.S.; Olszewski, K.; De Quevedo, D.M.; Barbir, J. An Assessment of Attitudes towards Plastics and Bioplastics in Europe. Sci. Total Environ. 2021, 755, 142732. [Google Scholar] [CrossRef] [PubMed]
  2. Plastics Europe. Plastics-the Facts 2022 OCTOBER. 2022. Available online: https://plasticseurope.org/wp-content/uploads/2022/10/PE-PLASTICS-THE-FACTS_V7-Tue_19-10-1.pdf (accessed on 1 May 2023).
  3. Beaumont, N.J.; Aanesen, M.; Austen, M.C.; Börger, T.; Clark, J.R.; Cole, M.; Hooper, T.; Lindeque, P.K.; Pascoe, C.; Wyles, K.J. Global Ecological, Social and Economic Impacts of Marine Plastic. Mar. Pollut. Bull. 2019, 142, 189–195. [Google Scholar] [CrossRef] [PubMed]
  4. Jambeck, J.R.; Geyer, R.; Wilcox, C.; Siegler, T.R.; Perryman, M.; Andrady, A.; Narayan, R.; Law, K.L. Plastic Waste Inputs from Land into the Ocean. Science 1979 2015, 347, 768–771. [Google Scholar] [CrossRef] [PubMed]
  5. Borrelle, S.B.; Ringma, J.; Law, K.L.; Monnahan, C.C.; Lebreton, L.; McGivern, A.; Murphy, E.; Jambeck, J.; Leonard, G.H.; Hilleary, M.A.; et al. Predicted Growth in Plastic Waste Exceeds Efforts to Mitigate Plastic Pollution. Science 2020, 369, 1515–1518. [Google Scholar] [CrossRef]
  6. Phelan, A.A.; Meissner, K.; Humphrey, J.; Ross, H. Plastic Pollution and Packaging: Corporate Commitments and Actions from the Food and Beverage Sector. J. Clean. Prod. 2022, 331, 129827. [Google Scholar] [CrossRef]
  7. Kan, M.; Miller, S.A. Environmental Impacts of Plastic Packaging of Food Products. Resour. Conserv. Recycl. 2022, 180, 106156. [Google Scholar] [CrossRef]
  8. Tester, R.F.; Karkalas, J. The Effects of Environmental Conditions on the Structural Features and Physico-Chemical Properties of Starches. Starch-Staerke 2001, 53, 513–519. [Google Scholar] [CrossRef]
  9. Magallanes-Cruz, P.A.; Duque-Buitrago, L.F.; del Rocío Martínez-Ruiz, N. Native and Modified Starches from Underutilized Seeds: Characteristics, Functional Properties and Potential Applications. Food Res. Int. 2023, 169, 112875. [Google Scholar] [CrossRef]
  10. Wang, S.; Zhang, P.; Li, Y.; Li, J.; Li, X.; Yang, J.; Ji, M.; Li, F.; Zhang, C. Recent Advances and Future Challenges of the Starch-Based Bio-Composites for Engineering Applications. Carbohydr. Polym. 2023, 307, 120627. [Google Scholar] [CrossRef]
  11. Carlstedt, J.; Wojtasz, J.; Fyhr, P.; Kocherbitov, V. Understanding Starch Gelatinization: The Phase Diagram Approach. Carbohydr. Polym. 2015, 129, 62–69. [Google Scholar] [CrossRef]
  12. Wang, S.; Sun, Y.; Wang, J.; Wang, S.; Copeland, L. Molecular Disassembly of Rice and Lotus Starches during Thermal Processing and Its Effect on Starch Digestibility. Food Funct. 2016, 7, 1188–1195. [Google Scholar] [CrossRef]
  13. Majeed, T.; Dar, A.H.; Pandey, V.K.; Dash, K.K.; Srivastava, S.; Shams, R.; Jeevarathinam, G.; Singh, P.; Echegaray, N.; Pandiselvam, R. Role of Additives in Starch-Based Edible Films and Coating: A Review with Current Knowledge. Prog. Org. Coat. 2023, 181, 107597. [Google Scholar] [CrossRef]
  14. Karimi Sani, I.; Masoudpour-Behabadi, M.; Alizadeh Sani, M.; Motalebinejad, H.; Juma, A.S.M.; Asdagh, A.; Eghbaljoo, H.; Khodaei, S.M.; Rhim, J.W.; Mohammadi, F. Value-Added Utilization of Fruit and Vegetable Processing by-Products for the Manufacture of Biodegradable Food Packaging Films. Food Chem. 2023, 405, 134964. [Google Scholar] [CrossRef]
  15. Farajpour, R.; Emam Djomeh, Z.; Moeini, S.; Tavahkolipour, H.; Safayan, S. Structural and Physico-Mechanical Properties of Potato Starch-Olive Oil Edible Films Reinforced with Zein Nanoparticles. Int. J. Biol. Macromol. 2020, 149, 941–950. [Google Scholar] [CrossRef]
  16. Thakur, R.; Pristijono, P.; Scarlett, C.J.; Bowyer, M.; Singh, S.P.; Vuong, Q.V. Starch-Based Films: Major Factors Affecting Their Properties. Int. J. Biol. Macromol. 2019, 132, 1079–1089. [Google Scholar] [CrossRef]
  17. Arora, A.; Padua, G.W. Review: Nanocomposites in Food Packaging. J. Food Sci. 2010, 75, 43–49. [Google Scholar] [CrossRef]
  18. Ghanbarzadeh, B.; Oleyaei, S.A.; Almasi, H. Nanostructured Materials Utilized in Biopolymer-Based Plastics for Food Packaging Applications. Crit. Rev. Food Sci. Nutr. 2015, 55, 1699–1723. [Google Scholar] [CrossRef]
  19. García, N.L.; Famá, L.; D’Accorso, N.B.; Goyanes, S. Biodegradable Starch Nanocomposites. Adv. Struct. Mater. 2015, 75, 17–77. [Google Scholar] [CrossRef]
  20. 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]
  21. Liu, D.; Zhao, P.; Chen, J.; Yan, Y.; Wu, Z. Recent Advances and Applications in Starch for Intelligent Active Food Packaging: A Review. Foods 2022, 11, 2879. [Google Scholar] [CrossRef]
  22. Jane, J.; Chen, Y.Y.; Lee, L.F.; McPherson, A.E.; Wong, K.S.; Radosavljevic, M.; Kasemsuwan, T. Effects of Amylopectin Branch Chain Length and Amylose Content on the Gelatinization and Pasting Properties of Starch. Cereal Chem. 1999, 76, 629–637. [Google Scholar] [CrossRef]
  23. Willfahrt, A.; Steiner, E.; Hötzel, J.; Crispin, X. Printable Acid-Modified Corn Starch as Non-Toxic, Disposable Hydrogel-Polymer Electrolyte in Supercapacitors. Appl. Phys. A Mater. Sci. Process 2019, 125, 474. [Google Scholar] [CrossRef]
  24. Singh Sandhu, K.; Singh, N. Relationships between Selected Properties of Starches from Different Corn Lines. Int. J. Food Prop. 2005, 8, 481–491. [Google Scholar] [CrossRef]
  25. Singh, G.D.; Bawa, A.S.; Singh, S.; Saxena, D.C. Physicochemical, Pasting, Thermal and Morphological Characteristics of Indian Water Chestnut (Trapa natans) Starch. Starch-Staerke 2009, 61, 35–42. [Google Scholar] [CrossRef]
  26. Mohammed, A.A.B.A.; Hasan, Z.; Omran, A.A.B.; Kumar, V.V.; Elfaghi, A.M.; Ilyas, R.A.; Sapuan, S.M. Corn: Its Structure, Polymer, Fiber, Composite, Properties, and Applications. Polymers 2022, 14, 4396. [Google Scholar] [CrossRef]
  27. Ibrahim, M.I.J.; Sapuan, S.M.; Zainudin, E.S.; Zuhri, M.Y.M. Extraction, chemical composition, and characterization of potential lignocellulosic biomasses and polymers from corn plant parts. BioResources 2019, 14, 6485–6500. [Google Scholar] [CrossRef]
  28. Marichelvam, M.K.; Jawaid, M.; Asim, M. Corn and Rice Starch-Based Bio-Plastics as Alternative Packaging Materials. Fibers 2019, 7, 32. [Google Scholar] [CrossRef] [Green Version]
  29. Liu, H.; Yu, L.; Xie, F.; Chen, L. Gelatinization of Cornstarch with Different Amylose/Amylopectin Content. Carbohydr. Polym. 2006, 65, 357–363. [Google Scholar] [CrossRef]
  30. Fredrikssona, H.; Silveriob, J.; Andemon’, R.; Eliassonb, A.-C.; Amanta, P. The Influence of Amylose and Amylopectin Characteristics on Gelatinization and Retrogradation Properties of Different Starches. Carbohydr. Polym. 1998, 35, 119–134. [Google Scholar] [CrossRef]
  31. Dhital, S.; Shrestha, A.K.; Hasjim, J.; Gidley, M.J. Physicochemical and Structural Properties of Maize and Potato Starches as a Function of Granule Size. J. Agric. Food Chem. 2011, 59, 10151–10161. [Google Scholar] [CrossRef]
  32. Simsek, S.; Ovando-Martinez, M.; Marefati, A.; Sj, M.; Rayner, M. Chemical Composition, Digestibility and Emulsification Properties of Octenyl Succinic Esters of Various Starches. Food Res. Int. 2015, 75, 41–49. [Google Scholar] [CrossRef]
  33. Waterschoot, J.; Gomand, S.V.; Fierens, E.; Delcour, J.A. Production, Structure, Physicochemical and Functional Properties of Maize, Cassava, Wheat, Potato and Rice Starches. Starch-Staerke 2015, 67, 14–29. [Google Scholar] [CrossRef]
  34. Vandeputte, G.E.; Vermeylen, R.; Geeroms, J.; Delcour, J.A. Rice Starches. I. Structural Aspects Provide Insight into Crystallinity Characteristics and Gelatinisation Behaviour of Granular Starch. J. Cereal Sci. 2003, 38, 43–52. [Google Scholar] [CrossRef]
  35. Juliano, B.O.; Villareal, R.M.; Perez, C.M.; Villareal, C.P.; Baños, L.; Takeda, V.; Hizukuri, S. Varietal Differences in Properties Among High Amylose Rice Starches. Starch-Stärke 1987, 39, 390–393. [Google Scholar] [CrossRef]
  36. Singh, N.; Kaur, L.; Sandhu, K.S.; Kaur, J.; Nishinari, K. Relationships between Physicochemical, Morphological, Thermal, Rheological Properties of Rice Starches. Food Hydrocoll. 2006, 20, 532–542. [Google Scholar] [CrossRef]
  37. Singh, N.; Singh, J.; Kaur, L.; Sodhi, N.S.; Gill, S. Morphological, Thermal and Rheological Properties of Starches from Different Botanical Sources. Food Chem. 2003, 81, 219–231. [Google Scholar] [CrossRef]
  38. Kumari, S.; Yadav, B.S.; Yadav, R.B. Synthesis and Modification Approaches for Starch Nanoparticles for Their Emerging Food Industrial Applications: A Review. Food Res. Int. 2020, 128, 108765. [Google Scholar] [CrossRef]
  39. Domene-López, D.; García-Quesada, J.C.; Martin-Gullon, I.; Montalbán, M.G. Influence of Starch Composition and Molecular Weight on Physicochemical Properties of Biodegradable Films. Polymers 2019, 11, 1084. [Google Scholar] [CrossRef] [Green Version]
  40. García, N.L.; Ribba, L.; Dufresne, A.; Aranguren, M.; Goyanes, S. Effect of Glycerol on the Morphology of Nanocomposites Made from Thermoplastic Starch and Starch Nanocrystals. Carbohydr. Polym. 2011, 84, 203–210. [Google Scholar] [CrossRef]
  41. Khan, B.; Bilal Khan Niazi, M.; Samin, G.; Jahan, Z. Thermoplastic Starch: A Possible Biodegradable Food Packaging Material—A Review. J. Food Process Eng. 2017, 40, e12447. [Google Scholar] [CrossRef]
  42. Pelissari, F.M.; Ferreira, D.C.; Louzada, L.B.; Dos Santos, F.; Corrêa, A.C.; Moreira, F.K.V.; Mattoso, L.H. Starch-Based Edible Films and Coatings: An Eco-Friendly Alternative for Food Packaging; Academic Press: Cambridge, MA, USA, 2018; ISBN 9780128094402. [Google Scholar]
  43. Agarwal, S. Major Factors Affecting the Characteristics of Starch Based Biopolymer Films. Eur. Polym. J. 2021, 160, 110788. [Google Scholar] [CrossRef]
  44. Li, C. Recent Progress in Understanding Starch Gelatinization—An Important Property Determining Food Quality. Carbohydr. Polym. 2022, 293, 119735. [Google Scholar] [CrossRef] [PubMed]
  45. Li, C.; Hu, Y. A Kinetics-Based Decomposition Approach to Reveal the Nature of Starch Asymmetric Gelatinization Thermograms at Non-Isothermal Conditions. Food Chem. 2021, 344, 128697. [Google Scholar] [CrossRef] [PubMed]
  46. Varavinit, S.; Shobsngob, S.; Varanyanond, W.; Chinachoti, P.; Naivikul, O. Effect of Amylose Content on Gelatinization, Retrogradation and Pasting Properties of Flours from Different Cultivars of Thai Rice. Starch-Staerke 2003, 55, 410–415. [Google Scholar] [CrossRef]
  47. Mathew, A.P.; Dufresne, A. Morphological Investigation of Nanocomposites from Sorbitol Plasticized Starch and Tunicin Whiskers. Biomacromolecules 2002, 3, 609–617. [Google Scholar] [CrossRef]
  48. Ma, X.; Yu, J. Formamide as the Plasticizer for Thermoplastic Starch. J. Appl. Polym. Sci. 2004, 93, 1769–1773. [Google Scholar] [CrossRef]
  49. Nawab, A.; Alam, F.; Haq, M.A.; Lutfi, Z.; Hasnain, A. Mango Kernel Starch-Gum Composite Films: Physical, Mechanical and Barrier Properties. Int. J. Biol. Macromol. 2017, 98, 869–876. [Google Scholar] [CrossRef]
  50. Van Soest, J.J.G.; Vliegenthart, J.F.G. Crystallinity in Starch Plastics: Consequences for Material Properties. Trends Biotechnol. 1997, 15, 208–213. [Google Scholar] [CrossRef]
  51. Chen, Y.L.; Shull, K.R. Controlling the Properties of Thermoplastic Starch Films with Hydrogen Bonding Plasticizers. Carbohydr. Polym. Technol. Appl. 2023, 5, 100291. [Google Scholar] [CrossRef]
  52. Luchese, C.L.; Spada, J.C.; Tessaro, I.C. Starch Content Affects Physicochemical Properties of Corn and Cassava Starch-Based Films. Ind. Crops Prod. 2017, 109, 619–626. [Google Scholar] [CrossRef]
  53. Averous, L.; Boquillon, N. Biocomposites Based on Plasticized Starch: Thermal and Mechanical Behaviours. Carbohydr. Polym. 2004, 56, 111–122. [Google Scholar] [CrossRef]
  54. Angellier, H.; Molina-Boisseau, S.; Dole, P.; Dufresne, A. Thermoplastic Starch—Waxy Maize Starch Nanocrystals Nanocomposites. Biomacromolecules 2006, 7, 531–539. [Google Scholar] [CrossRef]
  55. Ceseracciu, L.; Heredia-Guerrero, J.A.; Dante, S.; Athanassiou, A.; Bayer, I.S. Robust and Biodegradable Elastomers Based on Corn Starch and Polydimethylsiloxane (PDMS). ACS Appl. Mater. Interfaces 2015, 7, 3742–3753. [Google Scholar] [CrossRef]
  56. Müller, C.M.O.; Yamashita, F.; Laurindo, J.B. Evaluation of the Effects of Glycerol and Sorbitol Concentration and Water Activity on the Water Barrier Properties of Cassava Starch Films through a Solubility Approach. Carbohydr. Polym. 2008, 72, 82–87. [Google Scholar] [CrossRef]
  57. Li, J.; Luo, X. Strach-Based Blends. In Starch-Based Blends, Composites and Nanocomposites; Royal Society of Chemistry: London, UK, 2016; pp. 263–325. [Google Scholar]
  58. Siemann, U. Solvent Cast Technology—A Versatile Tool for Thin Film Production. Prog. Colloid. Polym. Sci. 2005, 130, 1–14. [Google Scholar] [CrossRef]
  59. 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]
  60. Sessini, V.; Arrieta, M.P.; Kenny, J.M.; Peponi, L. Processing of Edible Films Based on Nanoreinforced Gelatinized Starch. Polym. Degrad. Stab. 2016, 132, 157–168. [Google Scholar] [CrossRef]
  61. Müller, C.M.O.; Laurindo, J.B.; Yamashita, F. Effect of Nanoclay Incorporation Method on Mechanical and Water Vapor Barrier Properties of Starch-Based Films. Ind. Crops Prod. 2011, 33, 605–610. [Google Scholar] [CrossRef]
  62. Richard, E.; Mistler, E.R.T. Tape Casting: Theory and Practice; Wiley: Hoboken, NJ, USA, 2000; ISBN 978-1-574-98029-5. [Google Scholar]
  63. do Val Siqueira, L.; Arias, C.I.L.F.; Maniglia, B.C.; Tadini, C.C. Starch-Based Biodegradable Plastics: Methods of Production, Challenges and Future Perspectives. Curr. Opin. Food Sci. 2021, 38, 122–130. [Google Scholar] [CrossRef]
  64. Oliveira de Moraes, J.; Scheibe, A.S.; Augusto, B.; Carciofi, M.; Laurindo, J.B. Conductive Drying of Starch-Fiber Films Prepared by Tape Casting: Drying Rates and Film Properties. LWT—Food Sci. Technol. 2015, 64, 356–366. [Google Scholar] [CrossRef]
  65. De Moraes, J.O.; Scheibe, A.S.; Sereno, A.; Laurindo, J.B. Scale-up of the Production of Cassava Starch Based Films Using Tape-Casting. J. Food Eng. 2013, 119, 800–808. [Google Scholar] [CrossRef] [Green Version]
  66. Scheibe, A.S.; De Moraes, J.O.; Laurindo, J.B. Production and Characterization of Bags from Biocomposite Films of Starch-Vegetal Fibers Prepared by Tape Casting. J. Food Process Eng. 2014, 37, 482–492. [Google Scholar] [CrossRef]
  67. Nishihora, R.K.; Rachadel, P.L.; Quadri, M.G.N.; Hotza, D. Manufacturing Porous Ceramic Materials by Tape Casting—A Review. J. Eur. Ceram. Soc. 2018, 38, 988–1001. [Google Scholar] [CrossRef]
  68. Von Borries-Medrano, E.; Jaime-Fonseca, M.R.; Aguilar-Méndez, M.A. Starch-Guar Gum Extrudates: Microstructure, Physicochemical Properties and in-Vitro Digestion. Food Chem. 2016, 194, 891–899. [Google Scholar] [CrossRef] [PubMed]
  69. Harnkarnsujarit, N.; Wongphan, P.; Chatkitanan, T.; Laorenza, Y.; Srisa, A. Bioplastic for Sustainable Food Packaging; Elsevier Inc.: Amsterdam, The Netherlands, 2021; ISBN 9780128227145. [Google Scholar]
  70. González-Seligra, P.; Guz, L.; Ochoa-Yepes, O.; Goyanes, S.; Famá, L. Influence of Extrusion Process Conditions on Starch Film Morphology. LWT 2017, 84, 520–528. [Google Scholar] [CrossRef]
  71. Li, M.; Liu, P.; Zou, W.; Yu, L.; Xie, F.; Pu, H.; Liu, H.; Chen, L. Extrusion Processing and Characterization of Edible Starch Films with Different Amylose Contents. J. Food Eng. 2011, 106, 95–101. [Google Scholar] [CrossRef]
  72. Liu, W.; Wang, Z.; Liu, J.; Dai, B.; Hu, S.; Hong, R.; Xie, H.; Li, Z.; Chen, Y.; Zeng, G. Preparation, Reinforcement and Properties of Thermoplastic Starch Film by Film Blowing. Food Hydrocoll. 2020, 108, 106006. [Google Scholar] [CrossRef]
  73. Rejak, A.; Moscicki, L. TPS Film—Blowing. In Thermoplastic Starch. A Green Material for Various Industries; Janssen, L.P.B.M., Moscicki, L., Eds.; Wiley-VCH Verlag: Weinheim, Germany, 2009; pp. 173–184. ISBN 978-3-527-32528-3. [Google Scholar]
  74. Dang, K.M.; Yoksan, R. Thermoplastic Starch Blown Films with Improved Mechanical and Barrier Properties. Int. J. Biol. Macromol. 2021, 188, 290–299. [Google Scholar] [CrossRef]
  75. Palai, B.; Biswal, M.; Mohanty, S.; Nayak, S.K. In Situ Reactive Compatibilization of Polylactic Acid (PLA) and Thermoplastic Starch (TPS) Blends; Synthesis and Evaluation of Extrusion Blown Films Thereof. Ind. Crops Prod. 2019, 141, 111748. [Google Scholar] [CrossRef]
  76. Matzinos, P.; Tserki, V.; Kontoyiannis, A.; Panayiotou, C. Processing and Characterization of Starch/Polycaprolactone Products. Polym. Degrad. Stab. 2002, 77, 17–24. [Google Scholar] [CrossRef]
  77. de Campos, S.S.; de Oliveira, A.; Moreira, T.F.M.; da Silva, T.B.V.; da Silva, M.V.; Pinto, J.A.; Bilck, A.P.; Gonçalves, O.H.; Fernandes, I.P.; Barreiro, M.F.; et al. TPCS/PBAT Blown Extruded Films Added with Curcumin as a Technological Approach for Active Packaging Materials. Food Packag. Shelf Life 2019, 22, 100424. [Google Scholar] [CrossRef] [Green Version]
  78. Phothisarattana, D.; Wongphan, P.; Promhuad, K.; Promsorn, J.; Harnkarnsujarit, N. Blown Film Extrusion of PBAT/TPS/ZnO Nanocomposites for Shelf-Life Extension of Meat Packaging. Colloids Surf. B Biointerfaces 2022, 214, 112472. [Google Scholar] [CrossRef]
  79. Pelissari, F.M.; Grossmann, M.V.E.; Yamashita, F.; Pined, E.A.G. Antimicrobial, Mechanical, and Barrier Properties of Cassava Starch-Chitosan Films Incorporated with Oregano Essential Oil. J. Agric. Food Chem. 2009, 57, 7499–7504. [Google Scholar] [CrossRef]
  80. Cho, J.; Joshi, M.S.; Sun, C.T. Effect of Inclusion Size on Mechanical Properties of Polymeric Composites with Micro and Nano Particles. Compos. Sci. Technol. 2006, 66, 1941–1952. [Google Scholar] [CrossRef]
  81. Zare, Y. The Roles of Nanoparticles Accumulation and Interphase Properties in Properties of Polymer Particulate Nanocomposites by a Multi-Step Methodology. Compos. Part A Appl. Sci. Manuf. 2016, 91, 127–132. [Google Scholar] [CrossRef]
  82. Nguyen, V.S.; Rouxel, D.; Vincent, B. Dispersion of Nanoparticles: From Organic Solvents to Polymer Solutions. Ultrason. Sonochem. 2014, 21, 149–153. [Google Scholar] [CrossRef]
  83. Zhou, L.; Zhang, H.; Zhang, Z. Homogeneous Nanoparticle Dispersion Prepared with Impurity-Free Dispersant by the Ball Mill Technique. Particuology 2013, 11, 441–447. [Google Scholar] [CrossRef]
  84. Wolf, C.; Angellier-Coussy, H.; Gontard, N.; Doghieri, F.; Guillard, V. How the Shape of Fillers Affects the Barrier Properties of Polymer/Non-Porous Particles Nanocomposites: A Review. J. Memb. Sci. 2018, 556, 393–418. [Google Scholar] [CrossRef]
  85. Lengowski, E.C.; Bonfatti Júnior, E.A.; de Muñiz, G.I.B.; Satyanarayana, K.G. Barrier Properties of Nanoparticle-Based Polymer Composites. In Nanoparticle-Based Polymer Composites; Woodhead Publishing: Sawston, UK, 2022; pp. 219–241. [Google Scholar] [CrossRef]
  86. Ahmed, S.; Annu. Bionanocomposites: An Overview; Elsevier Ltd.: Amsterdam, The Netherlands, 2021; ISBN 9780128212806. [Google Scholar]
  87. Amenta, V.; Aschberger, K.; Arena, M.; Bouwmeester, H.; Botelho, F.; Brandhoff, P.; Gottardo, S.; Marvin, H.J.P.; Mech, A.; Quiros, L.; et al. Regulatory Aspects of Nanotechnology in the Agri/Feed/Food Sector in EU and Non-EU Countries. Regul. Toxicol. Pharmacol. 2015, 73, 463–476. [Google Scholar] [CrossRef]
  88. Food and Drug Administration (US). 21 CFR Part 11 Title 21—Food and Drugs Chapter I—Food and Drug Administration, Department of Health and Human Services Subchapter A—General; Food and Drug Administration: Silver Spring, MD, USA, 2023; Volume 575, pp. 1–2. [Google Scholar]
  89. European Commission Commission Regulation (EU). 2022/63 of 14 January 2022 Amending Annexes II and III to Regulation (EC) No 1333/2008 of European Parliament and of the Council as Regards the Food Additive Titanium Dioxide (E171). Off. J. Eur. Union 2022, 11, 1–5. [Google Scholar]
  90. European Parliament and Council. Regulation (EU) No 1169/2011 of the European Parliament and of the Council of 25 October 2011 on the Provision of Food Information to Consumers; European Parliament and Council: Brussels, Belgium, 2011; Volume L 304. [Google Scholar]
  91. Simoneau, C. Guidelines on Testing Conditions for Articles in Contact with Foodstuffs; European Communities: Brussels, Belgium, 2009. [Google Scholar]
  92. Lambré, C.; Barat Baviera, J.M.; Bolognesi, C.; Chesson, A.; Cocconcelli, P.S.; Crebelli, R.; Gott, D.M.; Grob, K.; Lampi, E.; Mengelers, M.; et al. Safety Assessment of the Substance Silver Nanoparticles for Use in Food Contact Materials. EFSA J. 2021, 19, e06790. [Google Scholar] [CrossRef] [PubMed]
  93. Peighambardoust, S.J.; Peighambardoust, S.H.; Mohammadzadeh Pournasir, N.; Pakdel, P. Properties of Active Starch-Based Films Incorporating a Combination of Ag, ZnO and CuO Nanoparticles for Potential Use in Food Packaging Applications. Food Packag. Shelf Life 2019, 22, 100420. [Google Scholar] [CrossRef]
  94. das Neves, M.D.S.; Scandorieiro, S.; Pereira, G.N.; Ribeiro, J.M.; Seabra, A.B.; Dias, A.P.; Yamashita, F.; Martinez, C.B.D.R.; Kobayashi, R.K.T.; Nakazato, G. Antibacterial Activity of Biodegradable Films Incorporated with Biologically-Synthesized Silver Nanoparticles and the Evaluation of Their Migration to Chicken Meat. Antibiotics 2023, 12, 178. [Google Scholar] [CrossRef] [PubMed]
  95. Nath, D.; Santhosh, R.; Pal, K.; Sarkar, P. Nanoclay-Based Active Food Packaging Systems: A Review. Food Packag. Shelf Life 2022, 31, 100803. [Google Scholar] [CrossRef]
  96. Choudalakis, G.; Gotsis, A.D. Permeability of Polymer/Clay Nanocomposites: A Review. Eur. Polym. J. 2009, 45, 967–984. [Google Scholar] [CrossRef]
  97. Huang, J.; Li, X.; Zhou, W. Safety Assessment of Nanocomposite for Food Packaging Application. Trends Food Sci. Technol. 2015, 45, 187–199. [Google Scholar] [CrossRef]
  98. Silvestre, C.; Duraccio, D.; Cimmino, S. Food Packaging Based on Polymer Nanomaterials. Prog. Polym. Sci. 2011, 36, 1766–1782. [Google Scholar] [CrossRef]
  99. Rafiee, R.; Shahzadi, R. Mechanical Properties of Nanoclay and Nanoclay Reinforced Polymers: A Review. Polym. Compos. 2019, 40, 431–445. [Google Scholar] [CrossRef]
  100. Tang, X.; Alavi, S.; Herald, T.J. Barrier and Mechanical Properties of Starch-Clay Nanocomposite Films. Cereal Chem. 2008, 85, 433–439. [Google Scholar] [CrossRef] [Green Version]
  101. Girdthep, S.; Worajittiphon, P.; Molloy, R.; Leejarkpai, T.; Punyodom, W. Formulation and Characterization of Compatibilized Poly(Lactic Acid)-Based Blends and Their Nanocomposites with Silver-Loaded Kaolinite. Polym. Int. 2015, 64, 203–211. [Google Scholar] [CrossRef]
  102. Osman, A.F.; Siah, L.; Alrashdi, A.A.; Ul-Hamid, A.; Ibrahim, I. Improving the Tensile and Tear Properties of Thermoplastic Starch/Dolomite Biocomposite Film through Sonication Process. Polymers 2021, 13, 274. [Google Scholar] [CrossRef]
  103. Gao, W.; Dong, H.; Hou, H.; Zhang, H. Effects of Clays with Various Hydrophilicities on Properties of Starch–Clay Nanocomposites by Film Blowing. Carbohydr. Polym. 2012, 88, 321–328. [Google Scholar] [CrossRef]
  104. Ruamcharoen, J.; Munlee, R.; Ruamcharoen, P. Eco-Friendly Bio-Based Composites of Cassava Starch and Natural Rubber Compatibilized with Nanoclays. Polym. Compos. 2023, 44, 1071–1082. [Google Scholar] [CrossRef]
  105. Cheviron, P.; Gouanvé, F.; Espuche, E. Preparation, Characterization and Barrier Properties of Silver/Montmorillonite/Starch Nanocomposite Films. J. Memb. Sci. 2016, 497, 162–171. [Google Scholar] [CrossRef]
  106. Park, H.; Lee, W.; Park, C. Environmentally Friendly Polymer Hybrids. J. Mater. Sci. 2003, 38, 909–915. [Google Scholar] [CrossRef]
  107. Sinharay, S.; Bousmina, M. Biodegradable Polymers and Their Layered Silicate Nanocomposites: In Greening the 21st Century Materials World. Prog. Mater. Sci. 2005, 50, 962–1079. [Google Scholar] [CrossRef]
  108. Llanos, J.H.R.; Tadini, C.C. Preparation and Characterization of Bio-Nanocomposite Films Based on Cassava Starch or Chitosan, Reinforced with Montmorillonite or Bamboo Nanofibers. Int. J. Biol. Macromol. 2018, 107, 371–382. [Google Scholar] [CrossRef]
  109. Huang, M.; Yu, J.; Ma, X. High Mechanical Performance MMT-Urea and Formamide-Plasticized Thermoplastic Cornstarch Biodegradable Nanocomposites. Carbohydr. Polym. 2006, 63, 393–399. [Google Scholar] [CrossRef]
  110. Mbey, J.A.; Hoppe, S.; Thomas, F. Cassava Starch–Kaolinite Composite Film. Effect of Clay Content and Clay Modification on Film Properties. Carbohydr. Polym. 2012, 88, 213–222. [Google Scholar] [CrossRef]
  111. Aimé Mbey, J.; Hoppe, S.; Thomas, F. Cassava Starch-Kaolinite Composite Films. Thermal and Mechanical Properties Related to Filler-Matrix Interactions. Polym. Compos. 2015, 36, 184–191. [Google Scholar] [CrossRef]
  112. Patakfalvi, R.; Oszkó, A.; Dékány, I. Synthesis and Characterization of Silver Nanoparticle/Kaolinite Composites. Colloids Surf. A Physicochem. Eng. Asp. 2003, 220, 45–54. [Google Scholar] [CrossRef]
  113. Girdthep, S.; Worajittiphon, P.; Molloy, R.; Lumyong, S.; Leejarkpai, T.; Punyodom, W. Biodegradable Nanocomposite Blown Fi Lms Based on Poly (Lactic Acid) Containing Silver-Loaded Kaolinite: A Route to Controlling Moisture Barrier Property and Silver Ion Release with a Prediction of Extended Shelf Life of Dried Longan. Polymer 2014, 55, 6776–6788. [Google Scholar] [CrossRef]
  114. Ren, J.; Dang, K.; Pollet, E.; Avérous, L. Preparation and Characterization of Thermoplastic Potato Starch/Halloysite Nano-Biocomposites: Effect of Plasticizer Nature and Nanoclay Content. Polymers 2018, 10, 808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Peponi, L.; Puglia, D.; Torre, L.; Valentini, L.; Kenny, J.M. Processing of Nanostructured Polymers and Advanced Polymeric Based Nanocomposites. Mater. Sci. Eng. R Rep. 2014, 85, 1–46. [Google Scholar] [CrossRef] [Green Version]
  116. Rafique, M.; Sadaf, I.; Rafique, M.S.; Tahir, M.B. A Review on Green Synthesis of Silver Nanoparticles and Their Applications. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1272–1291. [Google Scholar] [CrossRef]
  117. Ghaffari-Moghaddam, M.; Hadi-Dabanlou, R.; Khajeh, M.; Rakhshanipour, M.; Shameli, K. Green Synthesis of Silver Nanoparticles Using Plant Extracts. Korean J. Chem. Eng. 2014, 31, 548–557. [Google Scholar] [CrossRef]
  118. Ahmed, S.; Ahmad, M.; Swami, B.L.; Ikram, S. A Review on Plants Extract Mediated Synthesis of Silver Nanoparticles for Antimicrobial Applications: A Green Expertise. J. Adv. Res. 2016, 7, 17–28. [Google Scholar] [CrossRef] [Green Version]
  119. Klueh, U.; Wagner, V.; Kelly, S.; Johnson, A.; Bryers, J.D. Efficacy of Silver-Coated Fabric to Prevent Bacterial Colonization and Subsequent Device-Based Biofilm Formation. J. Biomed. Mater. Res. 2000, 53, 621–631. [Google Scholar] [CrossRef]
  120. Vieira, I.R.S.; de Carvalho, A.P.A.D.; Conte-Junior, C.A. Recent Advances in Biobased and Biodegradable Polymer Nanocomposites, Nanoparticles, and Natural Antioxidants for Antibacterial and Antioxidant Food Packaging Applications. Compr. Rev. Food Sci. Food Saf. 2022, 21, 3673–3716. [Google Scholar] [CrossRef]
  121. Sun, X.; Zhang, H.; Wang, J.; Dong, M.; Jia, P.; Bu, T.; Wang, Q.; Wang, L. Sodium Alginate-Based Nanocomposite Films with Strong Antioxidant and Antibacterial Properties Enhanced by Polyphenol-Rich Kiwi Peel Extracts Bio-Reduced Silver Nanoparticles. Food Packag. Shelf Life 2021, 29, 100741. [Google Scholar] [CrossRef]
  122. Hassan, A.H.A.; Omnia, A.; Ahmed, A.M.; Khalafalla, F.A.; Ali, F.H.M. Correspondence Antimicrobial Activity of Starch-Based Biodegradable Antimicrobial Films Incorporated with Biosynthesized Silver Nanoparticles Against Multiple Drug-Resistant Staphylococcus Aureus Food Isolates. J. Adv. Vet. Res. 2011, 13, 41–46. [Google Scholar]
  123. Mohseni, M.S.; Khalilzadeh, M.A.; Mohseni, M.; Hargalani, F.Z.; Getso, M.I.; Raissi, V.; Raiesi, O. Green Synthesis of Ag Nanoparticles from Pomegranate Seeds Extract and Synthesis of Ag-Starch Nanocomposite and Characterization of Mechanical Properties of the Films. Biocatal. Agric. Biotechnol. 2020, 25, 101569. [Google Scholar] [CrossRef]
  124. Khurshid, S.; Arif, S.; Ali, T.M.; Iqbal, H.M.; Shaikh, M.; Khurshid, H.; Akber, Q.; Yousaf, S. Effect of Silver Nanoparticles Prepared from Saraca Asoca Leaf Extract on Morphological, Functional, Mechanical, and Antibacterial Properties of Rice Starch Films. Starch-Stärke 2022, 74, 2100228. [Google Scholar] [CrossRef]
  125. Nechita, P. Active-Antimicrobial Coatings Based on Silver Nano-Particles and Natural Polymers for Paper Packaging Functionalization. Nord. Pulp Pap. Res. J. 2017, 32, 452–458. [Google Scholar] [CrossRef]
  126. Roy, S.; Rhim, J.-W. Starch/Agar-Based Functional Films Integrated with Enoki Mushroom-Mediated Silver Nanoparticles for Active Packaging Applications. Food Biosci. 2022, 49, 101867. [Google Scholar] [CrossRef]
  127. Rafeeq, C.M.; Paul, E.; Vidya Saagar, E.; Manzur Ali, P.P. Mycosynthesis of Zinc Oxide Nanoparticles Using Pleurotus Floridanus and Optimization of Process Parameters. Ceram. Int. 2021, 47, 12375–12380. [Google Scholar] [CrossRef]
  128. Yadav, N.; Singh, S.; Saini, O.; Srivastava, S. Technological Advancement in the Remediation of Heavy Metals Employing Engineered Nanoparticles: A Step towards Cleaner Water Process. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100757. [Google Scholar] [CrossRef]
  129. Kim, S.; Song, K. Bin Antimicrobial Activity of Buckwheat Starch Films Containing Zinc Oxide Nanoparticles against Listeria Monocytogenes on Mushrooms. Int. J. Food Sci. Technol. 2018, 53, 1549–1557. [Google Scholar] [CrossRef]
  130. Anugrahwidya, R.; Armynah, B.; Tahir, D. Composites Bioplastic Film for Various Concentration of Zinc Oxide (ZnO) Nanocrystals Towards Physical Properties for High Biodegradability in Soil and Seawater. J. Polym. Environ. 2022, 30, 2589–2601. [Google Scholar] [CrossRef]
  131. Babapour, H.; Jalali, H.; Mohammadi Nafchi, A. The Synergistic Effects of Zinc Oxide Nanoparticles and Fennel Essential Oil on Physicochemical, Mechanical, and Antibacterial Properties of Potato Starch Films. Food Sci. Nutr. 2021, 9, 3893–3905. [Google Scholar] [CrossRef]
  132. Ni, S.; Zhang, H.; Dai, H.; Xiao, H. Starch-Based Flexible Coating for Food Packaging Paper with Exceptional Hydrophobicity and Antimicrobial Activity. Polymers 2018, 10, 1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Rosman, N.F.; Nurfazianawatie, M.Z.; Omar, H.; Malek, N.S.A.; Hajar, N.; Buniyamin, I.; Abdullah, S.; Rusop, M.; Asli, N.A. Fungal Growth Physicochemical Properties Inhibition by Novel Zinc Oxide/Glutinous Tapioca Starch Composite. J. Adv. Res. Appl. Sci. Eng. Technol. 2022, 29, 76–89. [Google Scholar] [CrossRef]
  134. Solís-Gómez, A.; Neira-Velázquez, M.G.; Morales, J.; Sánchez-Castillo, M.A.; Pérez, E. Improving Stability of TiO2 Particles in Water by RF-Plasma Polymerization of Poly(Acrylic Acid) on the Particle Surface. Colloids Surf. A Physicochem. Eng. Asp. 2014, 451, 66–74. [Google Scholar] [CrossRef]
  135. Vallejo-Montesinos, J.; Gámez-Cordero, J.; Zarraga, R.; Pérez Pérez, M.C.; Gonzalez-Calderon, J.A. Influence of the Surface Modification of Titanium Dioxide Nanoparticles TiO2 under Efficiency of Silver Nanodots Deposition and Its Effect under the Properties of Starch–Chitosan (SC) Films. Polym. Bull. 2020, 77, 107–133. [Google Scholar] [CrossRef]
  136. Zhang, X.; Xiao, G.; Wang, Y.; Zhao, Y.; Su, H.; Tan, T. Preparation of Chitosan-TiO2 Composite Film with Efficient Antimicrobial Activities under Visible Light for Food Packaging Applications. Carbohydr. Polym. 2017, 169, 101–107. [Google Scholar] [CrossRef]
  137. Othman, S.H.; Abd Salam, N.R.; Zainal, N.; Kadir Basha, R.; Talib, R.A. Antimicrobial Activity of TiO2 Nanoparticle-Coated Film for Potential Food Packaging Applications. Int. J. Photoenergy 2014, 2014, 945930. [Google Scholar] [CrossRef] [Green Version]
  138. Tunma, S. Starch Based Nanocomposites in Active Packaging for Extended Shelf Life of Fresh Fruits. Walailak J. Sci. Technol. 2018, 15, 273–281. [Google Scholar] [CrossRef]
  139. Ostafińska, A.; Mikešová, J.; Krejčíková, S.; Nevoralová, M.; Šturcová, A.; Zhigunov, A.; Michálková, D.; Šlouf, M. Thermoplastic Starch Composites with TiO2 Particles: Preparation, Morphology, Rheology and Mechanical Properties. Int. J. Biol. Macromol. 2017, 101, 273–282. [Google Scholar] [CrossRef]
  140. Arezoo, E.; Mohammadreza, E.; Maryam, M.; Abdorreza, M.N. The Synergistic Effects of Cinnamon Essential Oil and Nano TiO2 on Antimicrobial and Functional Properties of Sago Starch Films. Int. J. Biol. Macromol. 2020, 157, 743–751. [Google Scholar] [CrossRef]
  141. Bumbudsanpharoke, N.; Harnkarnsujarit, N.; Chongcharoenyanon, B.; Kwon, S.; Ko, S. Enhanced Properties of PBAT/TPS Biopolymer Blend with CuO Nanoparticles for Promising Active Packaging. Food Packag. Shelf Life 2023, 37, 101072. [Google Scholar] [CrossRef]
  142. Almasi, H.; Jafarzadeh, P.; Mehryar, L. Fabrication of Novel Nanohybrids by Impregnation of CuO Nanoparticles into Bacterial Cellulose and Chitosan Nanofibers: Characterization, Antimicrobial and Release Properties. Carbohydr. Polym. 2018, 186, 273–281. [Google Scholar] [CrossRef] [PubMed]
  143. Francis, D.V.; Thaliyakattil, S.; Cherian, L.; Sood, N.; Gokhale, T. Metallic Nanoparticle Integrated Ternary Polymer Blend of PVA/Starch/Glycerol: A Promising Antimicrobial Food Packaging Material. Polymers 2022, 14, 1379. [Google Scholar] [CrossRef]
  144. Kubo, A.L.; Rausalu, K.; Savest, N.; Žusinaite, E.; Vasiliev, G.; Viirsalu, M.; Plamus, T.; Krumme, A.; Merits, A.; Bondarenko, O. Antibacterial and Antiviral Effects of Ag, Cu and Zn Metals, Respective Nanoparticles and Filter Materials Thereof against Coronavirus SARS-CoV-2 and Influenza A Virus. Pharmaceutics 2022, 14, 2549. [Google Scholar] [CrossRef]
  145. Wang, H.; Zhao, Y.; Yin, S.; Dai, Y.; Zhao, J.; Wang, Z.; Xing, B. Antagonism Toxicity of CuO Nanoparticles and Mild Ocean Acidification to Marine Algae. J. Hazard. Mater. 2023, 448, 130857. [Google Scholar] [CrossRef]
  146. Ramyadevi, J.; Jeyasubramanian, K.; Marikani, A.; Rajakumar, G.; Rahuman, A.A. Synthesis and Antimicrobial Activity of Copper Nanoparticles. Mater. Lett. 2012, 71, 114–116. [Google Scholar] [CrossRef]
  147. Garcia, C.V.; Shin, G.H.; Kim, J.T. Metal Oxide-Based Nanocomposites in Food Packaging: Applications, Migration, and Regulations. Trends Food Sci. Technol. 2018, 82, 21–31. [Google Scholar] [CrossRef]
  148. Rehana, D.; Mahendiran, D.; Kumar, R.S.; Rahiman, A.K. Evaluation of Antioxidant and Anticancer Activity of Copper Oxide Nanoparticles Synthesized Using Medicinally Important Plant Extracts. Biomed. Pharmacother. 2017, 89, 1067–1077. [Google Scholar] [CrossRef]
  149. Revathi, T.; Thambidurai, S. Cytotoxic, Antioxidant and Antibacterial Activities of Copper Oxide Incorporated Chitosan-Neem Seed Biocomposites. Int. J. Biol. Macromol. 2019, 139, 867–878. [Google Scholar] [CrossRef]
  150. Jadhav, M.S.; Kulkarni, S.; Raikar, P.; Barretto, D.A.; Vootla, S.K.; Raikar, U.S. Green Biosynthesis of CuO & Ag-CuO Nanoparticles from Malus Domestica Leaf Extract and Evaluation of Antibacterial, Antioxidant and DNA Cleavage Activities. New J. Chem. 2018, 42, 204–213. [Google Scholar] [CrossRef]
  151. Das, D.; Nath, B.C.; Phukon, P.; Dolui, S.K. Synthesis and Evaluation of Antioxidant and Antibacterial Behavior of CuO Nanoparticles. Colloids Surf. B Biointerfaces 2013, 101, 430–433. [Google Scholar] [CrossRef]
  152. Hosseini, S.N.; Pirsa, S.; Farzi, J. Biodegradable Nano Composite Film Based on Modified Starch-Albumin/MgO.; Antibacterial, Antioxidant and Structural Properties. Polym. Test. 2021, 97, 107182. [Google Scholar] [CrossRef]
  153. Luo, D.; Xie, Q.; Gu, S.; Xue, W. Potato Starch Films by Incorporating Tea Polyphenol and MgO Nanoparticles with Enhanced Physical, Functional and Preserved Properties. Int. J. Biol. Macromol. 2022, 221, 108–120. [Google Scholar] [CrossRef] [PubMed]
  154. Sani, I.K.; Geshlaghi, S.P.; Pirsa, S.; Asdagh, A. Composite Film Based on Potato Starch/Apple Peel Pectin/ZrO2 Nanoparticles/Microencapsulated Zataria Multiflora Essential Oil; Investigation of Physicochemical Properties and Use in Quail Meat Packaging. Food Hydrocoll. 2021, 117, 106719. [Google Scholar] [CrossRef]
  155. Sun, Q. Starch Nanoparticles. In Starch in Food; Elsevier: Amsterdam, The Netherlands, 2018; pp. 691–745. ISBN 9780081008683. [Google Scholar]
  156. Lin, N.; Tang, J.; Dufresne, A.; Tam, M.K.C. (Eds.) Advanced Functional Materials from Nanopolysaccharides; Springer Series in Biomaterials Science and Engineering; Springer: Singapore, 2019; Volume 15, ISBN 978-981-15-0912-4. [Google Scholar]
  157. Minakawa, A.F.K.; Faria-Tischer, P.C.S.; Mali, S. Simple Ultrasound Method to Obtain Starch Micro- and Nanoparticles from Cassava, Corn and Yam Starches. Food Chem. 2019, 283, 11–18. [Google Scholar] [CrossRef]
  158. Liu, C.; Li, K.; Li, X.; Zhang, M.; Li, J. Formation and Structural Evolution of Starch Nanocrystals from Waxy Maize Starch and Waxy Potato Starch. Int. J. Biol. Macromol. 2021, 180, 625–632. [Google Scholar] [CrossRef]
  159. Le Corre, D.; Angellier-Coussy, H. Preparation and Application of Starch Nanoparticles for Nanocomposites: A Review. React. Funct. Polym. 2014, 85, 97–120. [Google Scholar] [CrossRef]
  160. Bel Haaj, S.; Thielemans, W.; Magnin, A.; Boufi, S. Starch Nanocrystals and Starch Nanoparticles from Waxy Maize as Nanoreinforcement: A Comparative Study. Carbohydr. Polym. 2016, 143, 310–317. [Google Scholar] [CrossRef]
  161. Angellier, H.; Choisnard, L.; Molina-Boisseau, S.; Ozil, P.; Dufresne, A. Optimization of the Preparation of Aqueous Suspensions of Waxy Maize Starch Nanocrystals Using a Response Surface Methodology. Biomacromolecules 2004, 5, 1545–1551. [Google Scholar] [CrossRef]
  162. Zhou, J.; Tong, J.; Su, X.; Ren, L. Hydrophobic Starch Nanocrystals Preparations through Crosslinking Modification Using Citric Acid. Int. J. Biol. Macromol. 2016, 91, 1186–1193. [Google Scholar] [CrossRef]
  163. Wang, S.; Blazek, J.; Gilbert, E.; Copeland, L. New Insights on the Mechanism of Acid Degradation of Pea Starch. Carbohydr. Polym. 2012, 87, 1941–1949. [Google Scholar] [CrossRef]
  164. Dufresne, A.; Cavaillé, J.-Y.; Helbert, W. New Nanocomposite Materials: Microcrystalline Starch Reinforced Thermoplastic. Macromolecules 1996, 29, 7624–7626. [Google Scholar] [CrossRef]
  165. Putaux, J.L.; Molina-Boisseau, S.; Momaur, T.; Dufresne, A. Platelet Nanocrystals Resulting from the Disruption of Waxy Maize Starch Granules by Acid Hydrolysis. Biomacromolecules 2003, 4, 1198–1202. [Google Scholar] [CrossRef]
  166. Dufresne, A. Polysaccharide Nanocrystal Reinforced Nanocomposites. Can. J. Chem. 2008, 86, 484–494. [Google Scholar] [CrossRef]
  167. LeCorre, D.; Bras, J.; Dufresne, A. Influence of Botanic Origin and Amylose Content on the Morphology of Starch Nanocrystals. J. Nanopart. Res. 2011, 13, 7193–7208. [Google Scholar] [CrossRef]
  168. Kim, H.Y.; Park, D.J.; Kim, J.Y.; Lim, S.T. Preparation of Crystalline Starch Nanoparticles Using Cold Acid Hydrolysis and Ultrasonication. Carbohydr. Polym. 2013, 98, 295–301. [Google Scholar] [CrossRef]
  169. Sun, Q.; Li, G.; Dai, L.; Ji, N.; Xiong, L. Green Preparation and Characterisation of Waxy Maize Starch Nanoparticles through Enzymolysis and Recrystallisation. Food Chem. 2014, 162, 223–228. [Google Scholar] [CrossRef]
  170. Bel Haaj, S.; Magnin, A.; Pétrier, C.; Boufi, S. Starch Nanoparticles Formation via High Power Ultrasonication. Carbohydr. Polym. 2013, 92, 1625–1632. [Google Scholar] [CrossRef]
  171. Patel, C.M.; Chakraborty, M.; Murthy, Z.V.P. Fast and Scalable Preparation of Starch Nanoparticles by Stirred Media Milling. Adv. Powder Technol. 2016, 27, 1287–1294. [Google Scholar] [CrossRef]
  172. Lamanna, M.; Morales, N.J.; Garcia, N.L.; Goyanes, S. Development and Characterization of Starch Nanoparticles by Gamma Radiation: Potential Application as Starch Matrix Filler. Carbohydr. Polym. 2013, 97, 90–97. [Google Scholar] [CrossRef]
  173. Liu, D.; Wu, Q.; Chen, H.; Chang, P.R. Transitional Properties of Starch Colloid with Particle Size Reduction from Micro- to Nanometer. J. Colloid Interface Sci. 2009, 339, 117–124. [Google Scholar] [CrossRef]
  174. Apostolidis, E.; Mandala, I. Modification of Resistant Starch Nanoparticles Using High-Pressure Homogenization Treatment. Food Hydrocoll. 2020, 103, 105677. [Google Scholar] [CrossRef]
  175. Harsanto, B.W.; Pranoto, Y.; Supriyanto; Kartini, I. Breadfruit-Based Starch Nanoparticles Prepared Using Nanoprecipitation to Stabilize a Pickering Emulsion. J. Southwest Jiaotong Univ. 2021, 56, 372–383. [Google Scholar] [CrossRef]
  176. Bernardo, C.N.; Kling, I.C.S.; Ferreira, W.H.; Andrade, C.T.; Simao, R.A. Starch Films Containing Starch Nanoparticles as Produced in a Single Step Green Route. Ind. Crops Prod. 2022, 177, 114481. [Google Scholar] [CrossRef]
  177. Mohammad Amini, A.; Razavi, S.M.A. A Fast and Efficient Approach to Prepare Starch Nanocrystals from Normal Corn Starch. Food Hydrocoll. 2016, 57, 132–138. [Google Scholar] [CrossRef]
  178. Maryam; Kasim, A.; Novelina; Emriadi. Preparation and Characterization of Sago (Metroxylon Sp.) Starch Nanoparticles Using Hydrolysis-Precipitation Method. J. Phys. Conf. Ser. 2020, 1481, 012021. [Google Scholar] [CrossRef]
  179. Dai, L.; Li, C.; Zhang, J.; Cheng, F. Preparation and Characterization of Starch Nanocrystals Combining Ball Milling with Acid Hydrolysis. Carbohydr. Polym. 2018, 180, 122–127. [Google Scholar] [CrossRef]
  180. Shang, Y.; Chao, C.; Yu, J.; Copeland, L.; Wang, S.; Wang, S. Starch Spherulites Prepared by a Combination of Enzymatic and Acid Hydrolysis of Normal Corn Starch. J. Agric. Food Chem. 2018, 66, 6357–6363. [Google Scholar] [CrossRef]
  181. Yai, H. Mechanical, Thermal and Structural Properties of Rice Starch Films Reinforced with Rice Starch Nanocrystals. Int. Food Res. J. 2013, 20, 439–449. [Google Scholar]
  182. Oliveira, A.V.; da Silva, A.P.M.; Barros, M.O.; de sá, M.; Souza Filho, M.; Rosa, M.F.; Azeredo, H.M.C. Nanocomposite Films from Mango Kernel or Corn Starch with Starch Nanocrystals. Starch-Staerke 2018, 70, 1800028. [Google Scholar] [CrossRef]
  183. Xu, C.; Chen, C.; Wu, D. The Starch Nanocrystal Fi Lled Biodegradable Poly (ε-Caprolactone) Composite Membrane with Highly Improved Properties. Carbohydr. Polym. 2018, 182, 115–122. [Google Scholar] [CrossRef]
  184. Ma, X.; Jian, R.; Chang, P.R.; Yu, J. Fabrication and Characterization of Citric Acid-Modified Starch Nanoparticles/Plasticized-Starch Composites. Biomacromolecules 2008, 9, 3314–3320. [Google Scholar] [CrossRef]
  185. Klemm, D.; Heublein, B.; Fink, H.-P.; Bohn, A. Cellulose: Fascinating Biopolymer and Sustainable Raw Material. Angew. Chem. Int. Ed. 2005, 44, 3358–3393. [Google Scholar] [CrossRef]
  186. Shanks, R.A. Chemistry and Structure of Cellulosic Fibres as Reinforcements in Natural Fibre Composites. In Natural Fibre Composites; Elsevier: Amsterdam, The Netherlands, 2014; pp. 66–83. ISBN 9780857099228. [Google Scholar]
  187. Habibi, Y.; Lucia, L.A.; Rojas, O.J. Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications. Chem. Rev. 2010, 110, 3479–3500. [Google Scholar] [CrossRef]
  188. Gregory, D.A.; Tripathi, L.; Fricker, A.T.R.; Asare, E.; Orlando, I.; Raghavendran, V.; Roy, I. Materials Science & Engineering R Bacterial Cellulose: A Smart Biomaterial with Diverse Applications. Mater. Sci. Eng. R 2021, 145, 100623. [Google Scholar] [CrossRef]
  189. ISO/TS 20477; Nanotechnologies—Vocabulary for Cellulose Nanomaterial. ISO: Vernier, Switzerland, 2023; Volume 2023.
  190. Kumar, A.; Dixit, K.; Sinha, N. Fabrication and Characterization of Additively Manufactured CNT-Bioglass Composite Scaffolds Coated with Cellulose Nanowhiskers for Bone Tissue Engineering. Ceram. Int. 2023, 49, 17639–17649. [Google Scholar] [CrossRef]
  191. Yusefi, M.; Shameli, K.; Lee-Kiun, M.S.; Teow, S.Y.; Moeini, H.; Ali, R.R.; Kia, P.; Jie, C.J.; Abdullah, N.H. Chitosan Coated Magnetic Cellulose Nanowhisker as a Drug Delivery System for Potential Colorectal Cancer Treatment. Int. J. Biol. Macromol. 2023, 233, 123388. [Google Scholar] [CrossRef]
  192. Sayyed, A.J.; Pinjari, D.V.; Sonawane, S.H.; Bhanvase, B.A.; Sheikh, J.; Sillanpää, M. Cellulose-Based Nanomaterials for Water and Wastewater Treatments: A Review. J. Environ. Chem. Eng. 2021, 9, 106626. [Google Scholar] [CrossRef]
  193. Aramwit, P.; Bang, N. The Characteristics of Bacterial Nanocellulose Gel Releasing Silk Sericin for Facial Treatment. BMC Biotechnol. 2014, 14, 104. [Google Scholar] [CrossRef] [Green Version]
  194. Moon, R.J.; Schueneman, G.T.; Simonsen, J. Overview of Cellulose Nanomaterials, Their Capabilities and Applications. JOM 2016, 68, 2383–2394. [Google Scholar] [CrossRef]
  195. Jonoobi, M.; Oladi, R.; Davoudpour, Y.; Oksman, K.; Dufresne, A.; Hamzeh, Y.; Davoodi, R. Different Preparation Methods and Properties of Nanostructured Cellulose from Various Natural Resources and Residues: A Review. Cellulose 2015, 22, 935–969. [Google Scholar] [CrossRef]
  196. Michelin, M.; Gomes, D.G.; Romaní, A.; Polizeli, M.D.L.T.M.; Teixeira, J.A. Nanocellulose Production: Exploring the Enzymatic Route and Residues of Pulp and Paper Industry. Molecules 2020, 25, 3411. [Google Scholar] [CrossRef] [PubMed]
  197. Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: A New Family of Nature-Based Materials. Angew. Chem.—Int. Ed. 2011, 50, 5438–5466. [Google Scholar] [CrossRef] [PubMed]
  198. Eichhorn, S.J. Cellulose Nanowhiskers: Promising Materials for Advanced Applications. Soft Matter 2011, 7, 303–315. [Google Scholar] [CrossRef]
  199. Liu, A.; Wu, H.; Naeem, A.; Du, Q.; Ni, B.; Liu, H.; Li, Z.; Ming, L. Cellulose Nanocrystalline from Biomass Wastes: An Overview of Extraction, Functionalization and Applications in Drug Delivery. Int. J. Biol. Macromol. 2023, 241, 124557. [Google Scholar] [CrossRef] [PubMed]
  200. Brinchi, L.; Cotana, F.; Fortunati, E.; Kenny, J.M. Production of Nanocrystalline Cellulose from Lignocellulosic Biomass: Technology and Applications. Carbohydr. Polym. 2013, 94, 154–169. [Google Scholar] [CrossRef]
  201. Chang, C.; Hou, J.; Chang, P.R.; Huang, J. Structure and Properties of Cellulose Nanocrystals. In Nanocellulose: From Fundamentals to Advanced Materials; Wiley: Hoboken, NJ, USA, 2019; pp. 21–52. ISBN 9781618394408. [Google Scholar]
  202. Nessi, V.; Falourd, X.; Maigret, J.E.; Cahier, K.; D’Orlando, A.; Descamps, N.; Gaucher, V.; Chevigny, C.; Lourdin, D. Cellulose Nanocrystals-Starch Nanocomposites Produced by Extrusion: Structure and Behavior in Physiological Conditions. Carbohydr. Polym. 2019, 225, 115123. [Google Scholar] [CrossRef]
  203. Sucinda, E.F.; Abdul Majid, M.S.; Ridzuan, M.J.M.; Cheng, E.M.; Alshahrani, H.A.; Mamat, N. Development and Characterisation of Packaging Film from Napier Cellulose Nanowhisker Reinforced Polylactic Acid (PLA) Bionanocomposites. Int. J. Biol. Macromol. 2021, 187, 43–53. [Google Scholar] [CrossRef]
  204. Azeredo, H.M.C.; Rosa, M.F.; Mattoso, L.H.C. Nanocellulose in Bio-Based Food Packaging Applications. Ind. Crops Prod. 2017, 97, 664–671. [Google Scholar] [CrossRef] [Green Version]
  205. Trinh, B.M.; Smith, M.; Mekonnen, T.H. A Nanomaterial-Stabilized Starch-Beeswax Pickering Emulsion Coating to Extend Produce Shelf-Life. Chem. Eng. J. 2022, 431, 133905. [Google Scholar] [CrossRef]
  206. Chen, Q.; Zhou, L.; Zou, J.; Gao, X. The Preparation and Characterization of Nanocomposite Film Reinforced by Modified Cellulose Nanocrystals. Int. J. Biol. Macromol. 2019, 132, 1155–1162. [Google Scholar] [CrossRef]
  207. Dhali, K.; Daver, F.; Cass, P.; Adhikari, B. Surface Modification of the Cellulose Nanocrystals through Vinyl Silane Grafting. Int. J. Biol. Macromol. 2022, 200, 397–408. [Google Scholar] [CrossRef]
  208. Zhang, K.; Zhou, M.; Cheng, F.; Lin, Y.; Zhu, P.; Li, J.; Tang, K. Preparation and Characterization of Starch-Based Nanocomposites Reinforced by Graphene Oxide Self-Assembled on the Surface of Silane Coupling Agent Modified Cellulose Nanocrystals. Int. J. Biol. Macromol. 2022, 198, 187–193. [Google Scholar] [CrossRef]
  209. Spence, K.L.; Venditti, R.A.; Rojas, O.J.; Habibi, Y.; Pawlak, J.J. A Comparative Study of Energy Consumption and Physical Properties of Microfibrillated Cellulose Produced by Different Processing Methods. Cellulose 2011, 18, 1097–1111. [Google Scholar] [CrossRef]
  210. Rol, F.; Karakashov, B.; Nechyporchuk, O.; Terrien, M.; Meyer, V.; Dufresne, A.; Belgacem, M.N.; Bras, J. Pilot-Scale Twin Screw Extrusion and Chemical Pretreatment as an Energy-Efficient Method for the Production of Nanofibrillated Cellulose at High Solid Content. ACS Sustain. Chem. Eng. 2017, 5, 6524–6531. [Google Scholar] [CrossRef]
  211. Hassan, M.L.; Bras, J.; Hassan, E.A.; Silard, C.; Mauret, E. Enzyme-Assisted Isolation of Microfibrillated Cellulose from Date Palm Fruit Stalks. Ind. Crops Prod. 2014, 55, 102–108. [Google Scholar] [CrossRef]
  212. Janardhnan, S.; Sain, M.M. Isolation of Cellulose Microfibrils—An Enzymatic Approach. Bioresources 2006, 1, 176–188. [Google Scholar] [CrossRef]
  213. Eyholzer, C.; Bordeanu, N.; Lopez-Suevos, F.; Rentsch, D.; Zimmermann, T.; Oksman, K. Preparation and Characterization of Water-Redispersible Nanofibrillated Cellulose in Powder Form. Cellulose 2010, 17, 19–30. [Google Scholar] [CrossRef] [Green Version]
  214. Saito, T.; Kimura, S.; Nishiyama, Y.; Isogai, A. Cellulose Nanofibers Prepared by TEMPO-Mediated Oxidation of Native Cellulose. Biomacromolecules 2007, 8, 2485–2491. [Google Scholar] [CrossRef]
  215. Hassan, M.L.; Bras, J.; Mauret, E.; Fadel, S.M.; Hassan, E.A.; El-wakil, N.A. Palm Rachis Microfibrillated Cellulose and Oxidized-Microfibrillated Cellulose for Improving Paper Sheets Properties of Unbeaten Softwood and Bagasse Pulps. Ind. Crops Prod. 2015, 64, 9–15. [Google Scholar] [CrossRef]
  216. Pennells, J.; Godwin, I.D.; Martin, D.J. Trends in the Production of Cellulose Nanofibers from Non- Wood Sources. Cellulose 2020, 27, 575–593. [Google Scholar] [CrossRef]
  217. Khawas, P.; Deka, S.C. Isolation and Characterization of Cellulose Nanofibers from Culinary Banana Peel Using High-Intensity Ultrasonication Combined with Chemical Treatment. Carbohydr. Polym. 2016, 137, 608–616. [Google Scholar] [CrossRef] [PubMed]
  218. Siqueira, G.; Oksman, K.; Tadokoro, S.K.; Mathew, A.P. Re-Dispersible Carrot Nano Fi Bers with High Mechanical Properties and Reinforcing Capacity for Use in Composite Materials. Compos. Sci. Technol. 2016, 123, 49–56. [Google Scholar] [CrossRef] [Green Version]
  219. Balakrishnan, P.; Gopi, S.; MS, S.; Thomas, S. UV Resistant Transparent Bionanocomposite Films Based on Potato Starch/Cellulose for Sustainable Packaging. Starch-Stärke 2018, 70, 1700139. [Google Scholar] [CrossRef]
  220. Fazeli, M.; Keley, M.; Biazar, E. Preparation and Characterization of Starch-Based Composite Films Reinforced by Cellulose Nanofibers. Int. J. Biol. Macromol. 2018, 116, 272–280. [Google Scholar] [CrossRef] [PubMed]
  221. Hietala, M.; Mathew, A.P.; Oksman, K. Bionanocomposites of Thermoplastic Starch and Cellulose Nanofibers Manufactured Using Twin-Screw Extrusion. Eur. Polym. J. 2013, 49, 950–956. [Google Scholar] [CrossRef]
  222. Chen, Q.; Liu, Y.; Chen, G. A Comparative Study on the Starch-Based Biocomposite Films Reinforced by Nanocellulose Prepared from Different Non-Wood Fibers. Cellulose 2019, 26, 2425–2435. [Google Scholar] [CrossRef]
  223. Zhang, L.; Zhao, J.; Zhang, Y.; Li, F.; Jiao, X.; Li, Q. The Effects of Cellulose Nanocrystal and Cellulose Nanofiber on the Properties of Pumpkin Starch-Based Composite Films. Int. J. Biol. Macromol. 2021, 192, 444–451. [Google Scholar] [CrossRef]
  224. Punia, S.; Scott, W.; Ozogul, F.; Dunno, K.D.; Armstrong, G.; Dawson, P. Food Bioscience Development of Starch-Based Films Reinforced with Cellulosic Nanocrystals and Essential Oil to Extend the Shelf Life of Red Grapes. Food Biosci. 2022, 47, 101621. [Google Scholar] [CrossRef]
  225. Kümmerer, K.; Menz, J.; Schubert, T.; Thielemans, W. Biodegradability of Organic Nanoparticles in the Aqueous Environment. Chemosphere 2011, 82, 1387–1392. [Google Scholar] [CrossRef]
  226. Ullah, H.; Santos, H.A.; Khan, T. Applications of Bacterial Cellulose in Food, Cosmetics and Drug Delivery. Cellulose 2016, 23, 2291–2314. [Google Scholar] [CrossRef]
  227. Dourado, F.; Gama, M.; Rodrigues, A.C. A Review on the Toxicology and Dietetic Role of Bacterial Cellulose. Toxicol. Rep. 2017, 4, 543–553. [Google Scholar] [CrossRef] [Green Version]
  228. Azeredo, H.M.C.; Barud, H.; Farinas, C.S.; Vasconcellos, V.M.; Claro, A.M. Bacterial Cellulose as a Raw Material for Food and Food Packaging Applications. Front. Sustain. Food Syst. 2019, 3, 7. [Google Scholar] [CrossRef] [Green Version]
  229. Shi, Z.; Zhang, Y.; Phillips, G.O.; Yang, G. Utilization of Bacterial Cellulose in Food. Food Hydrocoll. 2014, 35, 539–545. [Google Scholar] [CrossRef]
  230. Mohammadkazemi, F.; Azin, M.; Ashori, A. Production of Bacterial Cellulose Using Different Carbon Sources and Culture Media. Carbohydr. Polym. 2015, 117, 518–523. [Google Scholar] [CrossRef]
  231. Avcioglu, N.H. Bacterial Cellulose: Recent Progress in Production and Industrial Applications. World J. Microbiol. Biotechnol. 2022, 38, 86. [Google Scholar] [CrossRef]
  232. Esa, F.; Tasirin, S.M.; Rahman, N.A. Overview of Bacterial Cellulose Production and Application. Agric. Agric. Sci. Procedia 2014, 2, 113–119. [Google Scholar] [CrossRef] [Green Version]
  233. Fabra, M.J.; López-Rubio, A.; Ambrosio-Martín, J.; Lagaron, J.M. Improving the Barrier Properties of Thermoplastic Corn Starch-Based Films Containing Bacterial Cellulose Nanowhiskers by Means of PHA Electrospun Coatings of Interest in Food Packaging. Food Hydrocoll. 2016, 61, 261–268. [Google Scholar] [CrossRef] [Green Version]
  234. Grande, C.J.; Torres, F.G.; Gomez, C.M.; Troncoso, O.P.; Canet-Ferrer, J.; Martínez-Pastor, J. Development of Self-Assembled Bacterial Cellulose-Starch Nanocomposites. Mater. Sci. Eng. C 2009, 29, 1098–1104. [Google Scholar] [CrossRef]
  235. Paillet, M.; Dufresne, A. Chitin Whisker Reinforced Thermoplastic Nanocomposites. Macromolecules 2001, 34, 6527–6530. [Google Scholar] [CrossRef]
  236. Muñoz-Núñez, C.; Fernández-García, M.; Muñoz-Bonilla, A. Chitin Nanocrystals: Environmentally Friendly Materials for the Development of Bioactive Films. Coatings 2022, 12, 144. [Google Scholar] [CrossRef]
  237. Dutta, P.K.; Tripathi, S.; Mehrotra, G.K.; Dutta, J. Perspectives for Chitosan Based Antimicrobial Films in Food Applications. Food Chem. 2009, 114, 1173–1182. [Google Scholar] [CrossRef]
  238. 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, 115602. [Google Scholar] [CrossRef] [PubMed]
  239. Shu, X.Z.; Zhu, K.J. A Novel Approach to Prepare Tripolyphosphate/Chitosan Complex Beads for Controlled Release Drug Delivery. Int. J. Pharm. 2000, 201, 51–58. [Google Scholar] [CrossRef] [PubMed]
  240. Rampino, A.; Borgogna, M.; Blasi, P.; Bellich, B.; Cesàro, A. Chitosan Nanoparticles: Preparation, Size Evolution and Stability. Int. J. Pharm. 2013, 455, 219–228. [Google Scholar] [CrossRef]
  241. Chang, P.R.; Jian, R.; Yu, J.; Ma, X. Fabrication and Characterisation of Chitosan Nanoparticles/Plasticised-Starch Composites. Food Chem. 2010, 120, 736–740. [Google Scholar] [CrossRef]
  242. Chang, P.R.; Jian, R.; Yu, J.; Ma, X. Starch-Based Composites Reinforced with Novel Chitin Nanoparticles. Carbohydr. Polym. 2010, 80, 420–425. [Google Scholar] [CrossRef]
  243. Antoniou, J.; Liu, F.; Majeed, H.; Zhong, F. Characterization of Tara Gum Edible Films Incorporated with Bulk Chitosan and Chitosan Nanoparticles: A Comparative Study. Food Hydrocoll. 2015, 44, 309–319. [Google Scholar] [CrossRef]
  244. Li, R.; Zhuang, D.; Feng, H.; Wang, S.; Zhu, J. Novel “All-in-One” Multifunctional Gelatin-Based Film for Beef Freshness Maintaining and Monitoring. Food Chem. 2023, 418, 136003. [Google Scholar] [CrossRef]
  245. McKeen, L.W. Introduction to Plastics and Polymers. In Film Properties of Plastics and Elastomers; William Andrew: Norwich, NY, USA, 2012; pp. 1–18. [Google Scholar] [CrossRef]
  246. Hazrati, K.Z.; Sapuan, S.M.; Zuhri, M.Y.M.; Jumaidin, R. Effect of Plasticizers on Physical, Thermal, and Tensile Properties of Thermoplastic Films Based on Dioscorea Hispida Starch. Int. J. Biol. Macromol. 2021, 185, 219–228. [Google Scholar] [CrossRef]
  247. Mali, S.; Sakanaka, L.S.; Yamashita, F.; Grossmann, M.V.E. Water Sorption and Mechanical Properties of Cassava Starch Films and Their Relation to Plasticizing Effect. Carbohydr. Polym. 2005, 60, 283–289. [Google Scholar] [CrossRef]
  248. Maruddin, F.; Malaka, R.; Baba, S.; Amqam, H.; Taufik, M.; Sabil, S. Brightness, Elongation and Thickness of Edible Film with Caseinate Sodium Using a Type of Plasticizer. IOP Conf. Ser. Earth Environ. Sci. 2020, 492, 012043. [Google Scholar] [CrossRef]
  249. Sanyang, M.L.; Sapuan, S.M.; Jawaid, M.; Ishak, M.R.; Sahari, J. Effect of Plasticizer Type and Concentration on Physical Properties of Biodegradable Films Based on Sugar Palm (Arenga pinnata) Starch for Food Packaging. J. Food Sci. Technol. 2016, 53, 326–336. [Google Scholar] [CrossRef] [Green Version]
  250. Ma, C.; Tan, C.; Xie, J.; Yuan, F.; Tao, H.; Guo, L.; Cui, B.; Yuan, C.; Gao, W.; Zou, F.; et al. Effects of Different Ratios of Mannitol to Sorbitol on the Functional Properties of Sweet Potato Starch Films. Int. J. Biol. Macromol. 2023, 242, 124914. [Google Scholar] [CrossRef]
  251. Chillo, S.; Flores, S.; Mastromatteo, M.; Conte, A.; Gerschenson, L.; Del Nobile, M.A. Influence of Glycerol and Chitosan on Tapioca Starch-Based Edible Film Properties. J. Food Eng. 2008, 88, 159–168. [Google Scholar] [CrossRef]
  252. Sahari, J.; Sapuan, S.M.; El-Shekeil, Y.A.; Ishak, M.R.; Akhtar, R. Natural Fibre-Reinforced Thermoplastic Starch Composites. In Starch-Based Blends, Composites and Nanocomposites; The Royal Society of Chemistry: London, UK, 2015; pp. 109–142. [Google Scholar]
  253. Geleta, T.T.; Habtegebreil, S.A.; Tolesa, G.N. Physical, Mechanical, and Optical Properties of Enset Starch from Bulla Films Influenced by Different Glycerol Concentrations and Temperatures. J. Food Process. Preserv. 2020, 44, e14586. [Google Scholar] [CrossRef]
  254. Ben, Z.Y.; Samsudin, H.; Yhaya, M.F. Glycerol: Its Properties, Polymer Synthesis, and Applications in Starch Based Films. Eur. Polym. J. 2022, 175, 111377. [Google Scholar] [CrossRef]
  255. Jiménez, A.; Fabra, M.J.; Talens, P.; Chiralt, A. Edible and Biodegradable Starch Films: A Review. Food Bioprocess Technol. 2012, 5, 2058–2076. [Google Scholar] [CrossRef]
  256. Zahiruddin, S.M.M.; Othman, S.H.; Tawakkal, I.S.M.A.; Talib, R.A. Mechanical and Thermal Properties of Tapioca Starch Films Plasticized with Glycerol and Sorbitol. Food Res. 2019, 3, 157–163. [Google Scholar] [CrossRef]
  257. Maniglia, B.C.; Tessaro, L.; Ramos, A.P.; Tapia-Blácido, D.R. Which Plasticizer Is Suitable for Films Based on Babassu Starch Isolated by Different Methods? Food Hydrocoll. 2019, 89, 143–152. [Google Scholar] [CrossRef]
  258. Benelli, G.; Pavela, R.; Giordani, C.; Casettari, L.; Curzi, G.; Cappellacci, L.; Petrelli, R.; Maggi, F. Acute and Sub-Lethal Toxicity of Eight Essential Oils of Commercial Interest against the Filariasis Mosquito Culex Quinquefasciatus and the Housefly Musca Domestica. Ind. Crops Prod. 2018, 112, 668–680. [Google Scholar] [CrossRef]
  259. Angulo-Milhem, S.; Verriele, M.; Nicolas, M.; Thevenet, F. Indoor Use of Essential Oils: Emission Rates, Exposure Time and Impact on Air Quality. Atmos. Environ. 2021, 244, 117863. [Google Scholar] [CrossRef]
  260. Mishra, A.P.; Devkota, H.P.; Nigam, M.; Adetunji, C.O.; Srivastava, N.; Saklani, S.; Shukla, I.; Azmi, L.; Shariati, M.A.; Melo Coutinho, H.D.; et al. Combination of Essential Oils in Dairy Products: A Review of Their Functions and Potential Benefits. LWT 2020, 133, 110116. [Google Scholar] [CrossRef]
  261. Mahmood, K.; Kamilah, H.; Karim, A.A.; Ariffin, F. Enhancing the Functional Properties of Fish Gelatin Mats by Dual Encapsulation of Essential Oils in β-Cyclodextrins/Fish Gelatin Matrix via Coaxial Electrospinning. Food Hydrocoll. 2023, 137, 108324. [Google Scholar] [CrossRef]
  262. Sharma, S.; Barkauskaite, S.; Jaiswal, A.K.; Jaiswal, S. Essential Oils as Additives in Active Food Packaging. Food Chem. 2021, 343, 128403. [Google Scholar] [CrossRef]
  263. Shlosman, K.; Rein, D.M.; Shemesh, R.; Koifman, N.; Caspi, A.; Cohen, Y. Encapsulation of Thymol and Eugenol Essential Oils Using Unmodified Cellulose: Preparation and Characterization. Polymers 2023, 15, 95. [Google Scholar] [CrossRef]
  264. Detoni, C.B.; De Oliveira, D.M.; Santo, I.E.; Pedro, A.S.; El-Bacha, R.; Da Silva Velozo, E.; Ferreira, D.; Sarmento, B.; De Magalhães Cabral-Albuquerque, E.C. Evaluation of Thermal-Oxidative Stability and Antiglioma Activity of Zanthoxylum tingoassuiba Essential Oil Entrapped into Multi- and Unilamellar Liposomes. J. Liposome Res. 2012, 22, 1–7. [Google Scholar] [CrossRef]
  265. Shi, F.; Zhao, J.H.; Liu, Y.; Wang, Z.; Zhang, Y.T.; Feng, N.P. Preparation and Characterization of Solid Lipid Nanoparticles Loaded with Frankincense and Myrrh Oil. Int. J. Nanomed. 2012, 7, 2033–2043. [Google Scholar] [CrossRef] [Green Version]
  266. Loi, M.; Paciolla, C.; Logrieco, A.F.; Mulè, G. Plant Bioactive Compounds in Pre- and Postharvest Management for Aflatoxins Reduction. Front. Microbiol. 2020, 11, 243. [Google Scholar] [CrossRef] [Green Version]
  267. Castillo, F.; Hernández, D.; Gallegos, G.; Rodríguez, R.; Aguilar, C.N. Antifungal Properties of Bioactive Compounds from Plants; INTECH Open Access Publisher: London, UK, 2012. [Google Scholar]
  268. Gershenzon, J.; Dudareva, N. The Function of Terpene Natural Products in the Natural World. Nat. Chem. Biol. 2007, 3, 408–414. [Google Scholar] [CrossRef]
  269. Gutiérrez-Del-río, I.; López-Ibáñez, S.; Magadán-Corpas, P.; Fernández-Calleja, L.; Pérez-Valero, Á.; Tuñón-Granda, M.; Miguélez, E.M.; Villar, C.J.; Lombó, F. Terpenoids and Polyphenols as Natural Antioxidant Agents in Food Preservation. Antioxidants 2021, 10, 1264. [Google Scholar] [CrossRef]
  270. Lafay, S.; Gil-Izquierdo, A. Bioavailability of Phenolic Acids. Phytochem. Rev. 2008, 7, 301–311. [Google Scholar] [CrossRef]
  271. Durazzo, A.; Lucarini, M.; Souto, E.B.; Cicala, C.; Caiazzo, E.; Izzo, A.A.; Novellino, E.; Santini, A. Polyphenols: A Concise Overview on the Chemistry, Occurrence, and Human Health. Phytother. Res. 2019, 33, 2221–2243. [Google Scholar] [CrossRef] [Green Version]
  272. Pateiro, M.; Gómez-Salazar, J.A.; Jaime-Patlán, M.; Sosa-Morales, M.E.; Lorenzo, J.M. Plant Extracts Obtained with Green Solvents as Natural Antioxidants in Fresh Meat Products. Antioxidants 2021, 10, 181. [Google Scholar] [CrossRef]
  273. Pateiro, M.; Munekata, P.E.S.; Sant’Ana, A.S.; Domínguez, R.; Rodríguez-Lázaro, D.; Lorenzo, J.M. Application of Essential Oils as Antimicrobial Agents against Spoilage and Pathogenic Microorganisms in Meat Products. Int. J. Food Microbiol. 2021, 337, 108966. [Google Scholar] [CrossRef]
  274. Jayasena, D.D.; Jo, C. Essential Oils as Potential Antimicrobial Agents in Meat and Meat Products: A Review. Trends Food Sci. Technol. 2013, 34, 96–108. [Google Scholar] [CrossRef]
  275. Fiore, A.; Park, S.; Volpe, S.; Torrieri, E.; Masi, P. Active Packaging Based on PLA and Chitosan-Caseinate Enriched Rosemary Essential Oil Coating for Fresh Minced Chicken Breast Application. Food Packag. Shelf Life 2021, 29, 100708. [Google Scholar] [CrossRef]
  276. Alizadeh-Sani, M.; Moghaddas Kia, E.; Ghasempour, Z.; Ehsani, A. Preparation of Active Nanocomposite Film Consisting of Sodium Caseinate, ZnO Nanoparticles and Rosemary Essential Oil for Food Packaging Applications. J. Polym. Environ. 2021, 29, 588–598. [Google Scholar] [CrossRef]
  277. Souza, V.G.L.; Rodrigues, C.; Ferreira, L.; Pires, J.R.A.; Duarte, M.P.; Coelhoso, I.; Fernando, A.L. In Vitro Bioactivity of Novel Chitosan Bionanocomposites Incorporated with Different Essential Oils. Ind. Crops Prod. 2019, 140, 111563. [Google Scholar] [CrossRef]
  278. Restrepo, A.E.; Rojas, J.D.; García, O.R.; Sánchez, L.T.; Pinzón, M.I.; Villa, C.C. Mechanical, Barrier, and Color Properties of Banana Starch Edible Films Incorporated with Nanoemulsions of Lemongrass (Cymbopogon citratus) and Rosemary (Rosmarinus officinalis) Essential Oils. Food Sci. Technol. Int. 2018, 24, 705–712. [Google Scholar] [CrossRef]
  279. Roy, S.; Rhim, J.W. Genipin-Crosslinked Gelatin/Chitosan-Based Functional Films Incorporated with Rosemary Essential Oil and Quercetin. Materials 2022, 15, 3769. [Google Scholar] [CrossRef]
  280. Jiang, Y.; Wu, N.; Fu, Y.J.; Wang, W.; Luo, M.; Zhao, C.J.; Zu, Y.G.; Liu, X.L. Chemical Composition and Antimicrobial Activity of the Essential Oil of Rosemary. Environ. Toxicol. Pharmacol. 2011, 32, 63–68. [Google Scholar] [CrossRef] [PubMed]
  281. Tafa, T.G.; Engida, A.M. Preparation of Green Film with Improved Physicochemical Properties and Enhanced Antimicrobial Activity Using Ingredients from Cassava Peel, Bamboo Leaf and Rosemary Leaf. Heliyon 2022, 8, e10130. [Google Scholar] [CrossRef] [PubMed]
  282. Rivera Leiva, A.F.; Hernández-Fernández, J.; Ortega Toro, R. Active Films Based on Starch and Wheat Gluten (Triticum vulgare) for Shelf-Life Extension of Carrots. Polymers 2022, 14, 5077. [Google Scholar] [CrossRef]
  283. Syafiq, R.; Sapuan, S.M.; Zuhri, M.R.M. Antimicrobial Activity, Physical, Mechanical and Barrier Properties of Sugar Palm Based Nanocellulose/Starch Biocomposite Films Incorporated with Cinnamon Essential Oil. J. Mater. Res. Technol. 2021, 11, 144–157. [Google Scholar] [CrossRef]
  284. Zhou, Y.; Wu, X.; Chen, J.; He, J. Effects of Cinnamon Essential Oil on the Physical, Mechanical, Structural and Thermal Properties of Cassava Starch-Based Edible Films. Int. J. Biol. Macromol. 2021, 184, 574–583. [Google Scholar] [CrossRef]
  285. Lu, F.; Ding, Y.C.; Ye, X.Q.; Ding, Y.T. Antibacterial Effect of Cinnamon Oil Combined with Thyme or Clove Oil. Agric. Sci. China 2011, 10, 1482–1487. [Google Scholar] [CrossRef]
  286. Singh, G.; Maurya, S.; de Lampasona, M.P.; Catalan, C.A.N. A Comparison of Chemical, Antioxidant and Anti-microbial Studies of Cinnamon Leaf and Bark Volatile Oils, Oleoresins and Their Constituents. Food Chem. Toxicol. 2007, 45, 1650–1661. [Google Scholar] [CrossRef]
  287. Trifan, A.; Zengin, G.; Brebu, M.; Skalicka-woźniak, K.; Luca, S.V. Phytochemical Characterization and Evaluation of the Antioxidant and Anti-enzymatic Activity of Five Common Spices: Focus on Their Essential Oils and Spent Material Extractives. Plants 2021, 10, 2692. [Google Scholar] [CrossRef]
  288. Xie, Y.; Huang, Q.; Wang, Z.; Cao, H.; Zhang, D. Structure-Activity Relationships of Cinnamaldehyde and Eugenol Derivatives against Plant Pathogenic Fungi. Ind. Crops Prod. 2017, 97, 388–394. [Google Scholar] [CrossRef]
  289. Clemente, I.; Aznar, M.; Silva, F.; Nerín, C. Antimicrobial Properties and Mode of Action of Mustard and Cinnamon Essential Oils and Their Combination against Foodborne Bacteria. Innov. Food Sci. Emerg. Technol. 2016, 36, 26–33. [Google Scholar] [CrossRef]
  290. Gutierrez, J.; Barry-Ryan, C.; Bourke, P. Antimicrobial Activity of Plant Essential Oils Using Food Model Media: Efficacy, Synergistic Potential and Interactions with Food Components. Food Microbiol. 2009, 26, 142–150. [Google Scholar] [CrossRef] [PubMed]
  291. Nieto, G.; Huvaere, K.; Skibsted, L.H. Antioxidant Activity of Rosemary and Thyme By-Products and Synergism with Added Antioxidant in a Liposome System. Eur. Food Res. Technol. 2011, 233, 11–18. [Google Scholar] [CrossRef]
  292. Cai, C.; Ma, R.; Duan, M.; Deng, Y.; Liu, T.; Lu, D. Effect of Starch Film Containing Thyme Essential Oil Microcapsules on Physicochemical Activity of Mango. LWT 2020, 131, 109700. [Google Scholar] [CrossRef]
  293. Issa, A.; Ibrahim, S.A.; Tahergorabi, R. Impact of Sweet Potato Starch-Based Nanocomposite Films Activated with Thyme Essential Oil on the Shelf-Life of Baby Spinach Leaves. Foods 2017, 6, 43. [Google Scholar] [CrossRef] [Green Version]
  294. Cui, Y.; Cheng, M.; Han, M.; Zhang, R.; Wang, X. Characterization and Release Kinetics Study of Potato Starch Nanocomposite Films Containing Mesoporous Nano-Silica Incorporated with Thyme Essential Oil. Int. J. Biol. Macromol. 2021, 184, 566–573. [Google Scholar] [CrossRef]
  295. Park, S.Y.; Raka, R.N.; Hui, X.L.; Song, Y.; Sun, J.L.; Xiang, J.; Wang, J.; Jin, J.M.; Li, X.K.; Xiao, J.S.; et al. Six Spain Thymus Essential Oils Composition Analysis and Their in Vitro and in Silico Study against Streptococcus Mutans. BMC Complement. Med. Ther. 2023, 23, 106. [Google Scholar] [CrossRef]
  296. Sivaram, S.; Somanathan, H.; Kumaresan, S.M.; Muthuraman, M.S. The Beneficial Role of Plant Based Thymol in Food Packaging Application: A Comprehensive Review. Appl. Food Res. 2022, 2, 100214. [Google Scholar] [CrossRef]
  297. Basavegowda, N.; Baek, K.H. Synergistic Antioxidant and Antibacterial Advantages of Essential Oils for Food Packaging Applications. Biomolecules 2021, 11, 1267. [Google Scholar] [CrossRef]
  298. López-Gómez, A.; Navarro-Martínez, A.; Garre, A.; Iguaz, A.; Maldonado-Guzmán, P.; Martínez-Hernández, G.B. Kinetics of Carvacrol Release from Active Paper Packaging for Fresh Fruits and Vegetables under Conditions of Open and Closed Package. Food Packag. Shelf Life 2023, 37, 101081. [Google Scholar] [CrossRef]
  299. Xu, Y.; Chen, L.; Zhang, Y.; Huang, Y.; Cao, J.; Jiang, W. Antimicrobial and Controlled Release Properties of Nanocomposite Film Containing Thymol and Carvacrol Loaded UiO-66-NH2 for Active Food Packaging. Food Chem. 2023, 404, 134427. [Google Scholar] [CrossRef]
  300. El Abdali, Y.; Mahraz, A.M.; Beniaich, G.; Mssillou, I.; Chebaibi, M.; Bin Jardan, Y.A.; Lahkimi, A.; Nafidi, H.A.; Aboul-Soud, M.A.M.; Bourhia, M.; et al. Essential Oils of Origanum Compactum Benth: Chemical Characterization, in Vitro, in Silico, Antioxidant, and Antibacterial Activities. Open Chem. 2023, 21, 20220282. [Google Scholar] [CrossRef]
  301. Bora, L.; Burkard, T.; Juan, M.H.S.; Radeke, H.H.; Muț, A.M.; Vlaia, L.L.; Magyari-Pavel, I.Z.; Diaconeasa, Z.; Socaci, S.; Borcan, F.; et al. Phytochemical Characterization and Biological Evaluation of Origanum Vulgare L. Essential Oil Formulated as Polymeric Micelles Drug Delivery Systems. Pharmaceutics 2022, 14, 2413. [Google Scholar] [CrossRef]
  302. Song, X.; Zuo, G.; Chen, F. Effect of Essential Oil and Surfactant on the Physical and Antimicrobial Properties of Corn and Wheat Starch Films. Int. J. Biol. Macromol. 2018, 107, 1302–1309. [Google Scholar] [CrossRef] [PubMed]
  303. Erdoğan Eliuz, E.; Bahadırlı, N.P. Investigation of Antibacterial Activity and Mechanism of T. Spicata Essential Oil, and Activation of the Hydrosol Formed as a by-Product with UV. Biologia 2023, 78, 1161–1170. [Google Scholar] [CrossRef]
  304. Ochoa-Velasco, C.E.; Pérez-Pérez, J.C.; Varillas-Torres, J.M.; Navarro-Cruz, A.R.; Hernández-Carranza, P.; Munguía-Pérez, R.; Cid-Pérez, T.S.; Avila-Sosa, R. Starch Edible Films/Coatings Added with Carvacrol and Thymol: In Vitro and in Vivo Evaluation against Colletotrichum Gloeosporioides. Foods 2021, 10, 175. [Google Scholar] [CrossRef]
  305. Mao, S.; Li, F.; Zhou, X.; Lu, C.; Zhang, T. Characterization and Sustained Release Study of Starch-Based Films Loaded with Carvacrol: A Promising UV-Shielding and Bioactive Nanocomposite Film. LWT 2023, 180, 114719. [Google Scholar] [CrossRef]
  306. Razzouk, S.; Mazri, M.A.; Jeldi, L.; Mnasri, B.; Ouahmane, L.; Alfeddy, M.N. Chemical Composition and Antimicrobial Activity of Essential Oils from Three Mediterranean Plants against Eighteen Pathogenic Bacteria and Fungi. Pharmaceutics 2022, 14, 1608. [Google Scholar] [CrossRef]
  307. Kiki, M.J. In Vitro Antiviral Potential, Antioxidant, and Chemical Composition of Clove (Syzygium aromaticum) Essential Oil. Molecules 2023, 28, 2421. [Google Scholar] [CrossRef]
  308. Devi, K.P.; Sakthivel, R.; Nisha, S.A.; Suganthy, N.; Pandian, S.K. Eugenol Alters the Integrity of Cell Membrane and Acts against the Nosocomial Pathogen Proteus Mirabilis. Arch. Pharm. Res. 2013, 36, 282–292. [Google Scholar] [CrossRef]
  309. Nazzaro, F.; Fratianni, F.; De Martino, L.; Coppola, R.; De Feo, V. Effect of Essential Oils on Pathogenic Bacteria. Pharmaceuticals 2013, 6, 1451–1474. [Google Scholar] [CrossRef]
  310. Devi, K.P.; Nisha, S.A.; Sakthivel, R.; Pandian, S.K. Eugenol (an Essential Oil of Clove) Acts as an Antibacterial Agent against Salmonella Typhi by Disrupting the Cellular Membrane. J. Ethnopharmacol. 2010, 130, 107–115. [Google Scholar] [CrossRef] [PubMed]
  311. Lin, R.; Chen, H.; Xu, R.; Liu, B.; Yuan, C.; Guo, L.; Liu, P.; Fang, Y.; Cui, B. Green Preparation of 3D Micronetwork Eugenol-Encapsuled Porous Starch for Improving the Performance of Starch-Based Antibacterial Film. Int. J. Biol. Macromol. 2023, 241, 124593. [Google Scholar] [CrossRef]
  312. Chen, Y.; Wei, F.; Mu, W.; Han, X. Antioxidant and Antibacterial Starch-Based Edible Films Composed of Eugenol/Gelatin Microspheres. New J. Chem. 2023, 47, 4228–4238. [Google Scholar] [CrossRef]
  313. Cheng, J.; Wang, H.; Kang, S.; Xia, L.; Jiang, S.; Chen, M.; Jiang, S. An Active Packaging Film Based on Yam Starch with Eugenol and Its Application for Pork Preservation. Food Hydrocoll. 2019, 96, 546–554. [Google Scholar] [CrossRef]
  314. Haddou, S.; Hassania Loukili, E.; Hbika, A.; Chahine, A.; Hammouti, B. Phytochemical Study Using HPLC-UV/GC–MS of Different of Cannabis sativa L. Seeds Extracts from Morocco. Mater. Today Proc. 2023, 72, 3896–3903. [Google Scholar] [CrossRef]
  315. Benoutman, A.; Erbiai, E.H.; Edderdaki, F.Z.; Cherif, E.K.; Saidi, R.; Lamrani, Z.; Pintado, M.; Pinto, E.; Esteves da Silva, J.C.G.; Maouni, A. Phytochemical Composition, Antioxidant and Antifungal Activity of Thymus Capitatus, a Medicinal Plant Collected from Northern Morocco. Antibiotics 2022, 11, 681. [Google Scholar] [CrossRef]
  316. Autor, E.; Cornejo, A.; Bimbela, F.; Maisterra, M.; Gandía, L.M.; Martínez-Merino, V. Extraction of Phenolic Compounds from Populus Salicaceae Bark. Biomolecules 2022, 12, 539. [Google Scholar] [CrossRef]
  317. Anwar, A.; Siddiqui, R.; Shah, M.R.; Khan, N.A. Gold Nanoparticle-Conjugated Cinnamic Acid Exhibits Antiacanthamoebic and Antibacterial Properties. Antimicrob. Agents Chemother. 2018, 62, 10–1128. [Google Scholar] [CrossRef] [Green Version]
  318. Schmidt, E.; Bail, S.; Friedl, S.M.; Jirovetz, L.; Buchbauer, G.; Wanner, J.; Denkova, Z.; Slavchev, A.; Stoyanova, A.; Geissler, M. Antimicrobial Activities of Single Aroma Compounds 1. Nat. Prod. Commun. 2010, 5, 1934578X1000500906. [Google Scholar] [CrossRef] [Green Version]
  319. Ordoñez, R.; Atarés, L.; Chiralt, A. Antibacterial Properties of Cinnamic and Ferulic Acids Incorporated to Starch and PLA Monolayer and Multilayer Films. Food Control 2022, 136, 108878. [Google Scholar] [CrossRef]
  320. Che Hamzah, N.H.; Khairuddin, N.; Muhamad, I.I.; Hassan, M.A.; Ngaini, Z.; Sarbini, S.R. Characterisation and Colour Response of Smart Sago Starch-Based Packaging Films Incorporated with Brassica Oleracea Anthocyanin. Membranes 2022, 12, 913. [Google Scholar] [CrossRef] [PubMed]
  321. Priyadarshi, R.; Ezati, P.; Rhim, J.-W. Recent Advances in Intelligent Food Packaging Applications Using Natural Food Colorants. ACS Food Sci. Technol. 2021, 1, 124–138. [Google Scholar] [CrossRef]
  322. Kong, J.; Ge, X.; Sun, Y.; Mao, M.; Yu, H.; Chu, R.; Wang, Y. Multi-Functional PH-Sensitive Active and Intelligent Packaging Based on Highly Cross-Linked Zein for the Monitoring of Pork Freshness. Food Chem. 2023, 404, 134754. [Google Scholar] [CrossRef] [PubMed]
  323. Qin, S.; Sun, H.; Wan, X.; Wu, Y.; Lin, X.; Kan, H.; Hou, D.; Zheng, Z.; He, X.; Liu, C. Carboxymethylcellulose Reinforced Starch Films and Rapid Detection of Spoiled Beverages. Front. Bioeng. Biotechnol. 2023, 10, 1099118. [Google Scholar] [CrossRef]
  324. Huang, H.L.; Tsai, I.L.; Lin, C.; Hang, Y.H.; Ho, Y.C.; Tsai, M.L.; Mi, F.L. Intelligent Films of Marine Polysaccharides and Purple Cauliflower Extract for Food Packaging and Spoilage Monitoring. Carbohydr. Polym. 2023, 299, 120133. [Google Scholar] [CrossRef]
  325. De Pascual-Teresa, S.; Sanchez-Ballesta, M.T. Anthocyanins: From Plant to Health. Phytochem. Rev. 2008, 7, 281–299. [Google Scholar] [CrossRef]
  326. Li, L.; Wang, W.; Zheng, M.; Sun, J.; Chen, Z.; Wang, J.; Ma, Q. Nanocellulose-Enhanced Smart Film for the Accurate Monitoring of Shrimp Freshness via Anthocyanin-Induced Color Changes. Carbohydr. Polym. 2023, 301, 120352. [Google Scholar] [CrossRef]
  327. Roy, S.; Rhim, J.W. Anthocyanin Food Colorant and Its Application in PH-Responsive Color Change Indicator Films. Crit. Rev. Food Sci. Nutr. 2021, 61, 2297–2325. [Google Scholar] [CrossRef]
  328. Fotouhi, M.; Nasihatkon, B.; Solouki, S.; Seidi, S. A Card Instead of a Lab: A Ligand Embedded in a Bio-Composite of Starch/Gelatin Intelligent Film for Milk Quality Test Followed by Colorimetric Analysis. Int. J. Biol. Macromol. 2023, 228, 426–434. [Google Scholar] [CrossRef]
  329. Qin, Y.; Yun, D.; Xu, F.; Chen, D.; Kan, J.; Liu, J. Smart Packaging Films Based on Starch/Polyvinyl Alcohol and Lycium Ruthenicum Anthocyanins-Loaded Nano-Complexes: Functionality, Stability and Application. Food Hydrocoll. 2021, 119, 106850. [Google Scholar] [CrossRef]
  330. Vedove, T.M.A.R.D.; Maniglia, B.C.; Tadini, C.C. Production of Sustainable Smart Packaging Based on Cassava Starch and Anthocyanin by an Extrusion Process. J. Food Eng. 2021, 289, 110274. [Google Scholar] [CrossRef]
  331. Zhai, X.; Shi, J.; Zou, X.; Wang, S.; Jiang, C.; Zhang, J.; Huang, X.; Zhang, W.; Holmes, M. Novel Colorimetric Films Based on Starch/Polyvinyl Alcohol Incorporated with Roselle Anthocyanins for Fish Freshness Monitoring. Food Hydrocoll. 2017, 69, 308–317. [Google Scholar] [CrossRef] [Green Version]
  332. Choi, I.; Lee, J.Y.; Lacroix, M.; Han, J. Intelligent PH Indicator Film Composed of Agar/Potato Starch and Anthocyanin Extracts from Purple Sweet Potato. Food Chem. 2017, 218, 122–128. [Google Scholar] [CrossRef]
  333. Zong, Z.; Liu, M.; Chen, H.; Farag, M.A.; Wu, W.; Fang, X.; Niu, B.; Gao, H. Preparation and Characterization of a Novel Intelligent Starch/Gelatin Binary Film Containing Purple Sweet Potato Anthocyanins for Flammulina Velutipes Mushroom Freshness Monitoring. Food Chem. 2023, 405, 134839. [Google Scholar] [CrossRef]
  334. Ceballos, R.L.; Ochoa-Yepes, O.; Goyanes, S.; Bernal, C.; Famá, L. Effect of Yerba Mate Extract on the Performance of Starch Films Obtained by Extrusion and Compression Molding as Active and Smart Packaging. Carbohydr. Polym. 2020, 244, 116495. [Google Scholar] [CrossRef]
  335. Kavas, E.; Terzioğlu, P.; Sıcak, Y. Betanin and Nano-Calcium Carbonate Incorporated Polyvinyl Alcohol/Starch Films for Active and Intelligent Packaging Applications. J. Polym. Environ. 2023. [Google Scholar] [CrossRef]
  336. Naghdi, S.; Rezaei, M.; Abdollahi, M. A Starch-Based PH-Sensing and Ammonia Detector Film Containing Betacyanin of Paperflower for Application in Intelligent Packaging of Fish. Int. J. Biol. Macromol. 2021, 191, 161–170. [Google Scholar] [CrossRef]
  337. Liu, D.; Dang, S.; Zhang, L.; Munsop, K.; Li, X. Corn Starch/Polyvinyl Alcohol Based Films Incorporated with Curcumin-Loaded Pickering Emulsion for Application in Intelligent Packaging. Int. J. Biol. Macromol. 2021, 188, 974–982. [Google Scholar] [CrossRef]
  338. Hazer, S.; Aytac, A. Monitoring Food Quality—Effect of Curcumin in the Development of Polyethylene/Thermoplastic Starch Based Smart Packaging. J. Vinyl Addit. Technol. 2023, 1–14. [Google Scholar] [CrossRef]
  339. Ezati, P.; Rhim, J.-W. Starch and Agar-Based Color-Indicator Films Integrated with Shikonin for Smart Packaging Application of Shrimp. ACS Food Sci. Technol. 2021, 1, 1963–1969. [Google Scholar] [CrossRef]
  340. Ranjeth Kumar Reddy, T.; Kim, H.J. Mechanical, Optical, Thermal, and Barrier Properties of Poly (Lactic Acid)/Curcumin Composite Films Prepared Using Twin-Screw Extruder. Food Biophys. 2019, 14, 22–29. [Google Scholar] [CrossRef]
Figure 1. Analytics related to keywords: starch, active, and packaging (source: Scopus).
Figure 1. Analytics related to keywords: starch, active, and packaging (source: Scopus).
Polymers 15 02972 g001
Figure 2. Chemical structure of starch-constituting natural polymers amylopectin and amylose [23].
Figure 2. Chemical structure of starch-constituting natural polymers amylopectin and amylose [23].
Polymers 15 02972 g002
Figure 4. Illustrative scheme of starch granule disruption [42].
Figure 4. Illustrative scheme of starch granule disruption [42].
Polymers 15 02972 g004
Figure 5. Representation of a tape casting setup used for ceramic manufacturing [67].
Figure 5. Representation of a tape casting setup used for ceramic manufacturing [67].
Polymers 15 02972 g005
Figure 6. Tortuosity pathway in polymeric matrices with different geometric nanoparticles [87].
Figure 6. Tortuosity pathway in polymeric matrices with different geometric nanoparticles [87].
Polymers 15 02972 g006
Figure 7. Possible dispersion states of nanoclay in composite structures [100].
Figure 7. Possible dispersion states of nanoclay in composite structures [100].
Polymers 15 02972 g007
Figure 8. Antimicrobial mechanisms of Ag nanoparticles [120].
Figure 8. Antimicrobial mechanisms of Ag nanoparticles [120].
Polymers 15 02972 g008
Figure 9. Acid hydrolysis with sulfuric acid of waxy maize and starch [158].
Figure 9. Acid hydrolysis with sulfuric acid of waxy maize and starch [158].
Polymers 15 02972 g009
Figure 12. Scheme of main steps for CNC preparation [200].
Figure 12. Scheme of main steps for CNC preparation [200].
Polymers 15 02972 g012
Figure 13. Scheme of the self-assembly BC–starch film: (a) Starch granules are in suspension in the culture medium; (b) After autoclaving, starch is partially gelatinized, amylose leaches and granules swell; (c) BC nanofibrils grow in presence of the partially gelatinized starch; (d) After hot pressing, the nanocomposite shows interpenetrating networks of amylose and cellulose. [234].
Figure 13. Scheme of the self-assembly BC–starch film: (a) Starch granules are in suspension in the culture medium; (b) After autoclaving, starch is partially gelatinized, amylose leaches and granules swell; (c) BC nanofibrils grow in presence of the partially gelatinized starch; (d) After hot pressing, the nanocomposite shows interpenetrating networks of amylose and cellulose. [234].
Polymers 15 02972 g013
Figure 14. Chemical structures of chitin and chitosan via deacetylation.
Figure 14. Chemical structures of chitin and chitosan via deacetylation.
Polymers 15 02972 g014
Figure 16. Oxidation mechanism [296].
Figure 16. Oxidation mechanism [296].
Polymers 15 02972 g016
Figure 17. Carvacrol structure.
Figure 17. Carvacrol structure.
Polymers 15 02972 g017
Figure 18. Eugenol structure.
Figure 18. Eugenol structure.
Polymers 15 02972 g018
Figure 19. Cinnamic acid and its derivate acids.
Figure 19. Cinnamic acid and its derivate acids.
Polymers 15 02972 g019
Table 1. Corn, potato, and rice starch content.
Table 1. Corn, potato, and rice starch content.
Corn (Maize) Starch
ContentAmountUnits
Density [26,27,28]1.356–1.4029g/cm3
Amylose [26,27,28,29]24.64–29.4g/100 g
Amylopectin [26,27]72–75.36g/100 g
Crude fats [26,27,28]0.32–7.13g/100 g
Crude proteins [26,27,28]7.70–0.38g/100 g
Moisture contents [26,27,28,29]10.45–12.7%
Potato starch
ContentAmountUnits
Amylose [30,31]17–38.8g/100 g
Amylopectin [32]73.42g/100 g
Crude fats [31]0.06–0.20g/100 g
Crude proteins [31]0.01–0.09g/100 g
Moisture contents [33]9–11%
Rice starch
ContentAmountUnits
Amylose [34,35,36]4.1–32g/100 g
Amylopectin [32]74.56g/100 g
Crude fats [34]1.9–28.1g/100 g
Crude proteins [34] 0–3.0g/100 g
Moisture contents [34]13%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Muñoz-Gimena, P.F.; Oliver-Cuenca, V.; Peponi, L.; López, D. A Review on Reinforcements and Additives in Starch-Based Composites for Food Packaging. Polymers 2023, 15, 2972. https://doi.org/10.3390/polym15132972

AMA Style

Muñoz-Gimena PF, Oliver-Cuenca V, Peponi L, López D. A Review on Reinforcements and Additives in Starch-Based Composites for Food Packaging. Polymers. 2023; 15(13):2972. https://doi.org/10.3390/polym15132972

Chicago/Turabian Style

Muñoz-Gimena, Pedro Francisco, Víctor Oliver-Cuenca, Laura Peponi, and Daniel López. 2023. "A Review on Reinforcements and Additives in Starch-Based Composites for Food Packaging" Polymers 15, no. 13: 2972. https://doi.org/10.3390/polym15132972

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