26

The use of biomass for packaging films and coatings

H.M.C. De Azeredo,    Institute of Food Research, UK

M.F. Rosa, M. De Sá and M. Souza Filho,    Embrapa Tropical Agroindustry, Brazil

K.W. Waldron,    Institute of Food Research, UK

Abstract:

Because of concerns involving the continual disposal of huge volumes of non-biodegradable food packaging materials, there has been an increasing tendency to replace petroleum-derived polymers with bio-based, environmentally friendly biodegradable macromolecules. There are a variety of biomass-derived structures which can be used as packaging materials, especially as films and coatings. Moreover, there are several biomass-derived compounds which can be used as additives for these materials such as plasticizers, crosslinking agents, and reinforcements, which can enhance physical properties and applicability of the materials for food packaging purposes. This review focuses on biomaterials which can be used to develop food packaging structures (especially films and coatings), their properties and interactions, and how they influence the packaging performance.

Key words

polysaccharides; proteins; lignocellulose; nanocomposites

26.1 Introduction

People have been using natural polymeric materials such as silk, wool, cotton, wood and leather for centuries, but the advent of the petroleum-derived plastics at the beginning of the twentieth century provided the food industry with an increasingly wide variety of synthetic materials to be used as food packaging. However, fossil fuels are limited and non-renewable, and recycling is limited because of technical as well as economic difficulties. Less than 3% of the waste plastic worldwide gets recycled, compared with recycling rates of 30% for paper, 35% for metals and 18% for glass, according to Helmut Kaiser Consultancy (2012). Moreover, once discarded, petroleum-based plastics are generally non-biodegradable, in that they are resistant to microbial attack. This is due to their water insolubility, and the problem that evolutionary processes have not been sufficiently rapid to create new enzymes capable of degrading synthetic polymers during their relatively short existence in the natural environment (Mueller, 2006). In particular, the accumulated waste generated by the continuous and extensive disposal of food packaging has raised considerable concerns over their deleterious effects on wildlife and the environment. Even their incineration can produce toxic compounds such as furans and dioxins produced from burning polyvinylchloride (PVC) (Jayasekara et al., 2005). In recent decades, the concerns surrounding conventional petroleum-based plastics have stimulated a focus of attention on natural macromolecules such as polysaccharides and proteins, because of their sustainable supply and biodegradability (Verbeek and van den Berg, 2010).

In 2011, the global use of biodegradable plastics was 0.85 million metric tons. BCC Research expects that the use of bioplastics will increase up to 3.7 million metric tons by 2016, a compound annual growth rate of 34.3% (BCC Research, 2012). According to studies by Helmut Kaiser Consultancy (2012), bioplastics are expected to cover approximately 25–30% of the total plastics market by 2020, and the market itself is estimated to reach over US$10 billion by 2020. Europe is one of the most important markets, partly due to an increasing environmental awareness, but also due to the limited amount of crude oil reserves.

Biodegradable materials and particularly renewable materials have been promoted as materials for use in food packaging, especially as flexible packaging, although applications for rigid packaging materials have also been mentioned (Mohareb and Mittal, 2007; Stepto, 2009).

Some biodegradable materials (actually most proteins and polysaccharides) are also edible, and can be used to develop edible packaging (films and coatings). Edible packaging materials are intended to be integral parts of foods and to be eaten with the products, thus they are also inherently biodegradable (Krochta, 2002). Edible films (or sheets) are stand-alone structures that are preformed separately from the food and then applied on the food surface or between food components, or sealed into edible pouches. Edible coatings, on the other hand, are formed directly onto the surface of the food products (Krochta and De Mulder-Johnston, 1997). Gel capsules, microcapsules and tablet coatings made from edible materials could also be considered as edible packaging (Janjarasskul and Krochta, 2010). Although edible packaging is not expected to replace conventional plastic packaging, they have an important role to play in the whole development of renewable packaging. This is because they can be used to extend food stability by reducing exchange of moisture, O2, CO2, lipids and flavour compounds between the food and the surrounding environment, thus increasing food stability. So, they help to improve the efficiency of food packaging and therefore reduce the amount of conventional packaging materials required for each application. Hence they have been included in this review.

The barrier requirements of an edible film or coating depend on their application and the properties of the food they are supposed to protect. For example, fresh fruits and vegetables are alive, respiring foods. Films or coatings intended for use on them should have low water vapour permeability so they reduce the desiccation rates, while the O2 permeability should be low enough to reduce the respiration rates, extending the produce shelf life, but not low enough to create anaerobic conditions, leading to ethanol production and off-flavor formation. Nuts are especially susceptible to oxidation, thus the barrier against O2 is the most important factor to provide an extended shelf life; barrier against UV light is also helpful to reduce oxidation rates, and moisture barrier reduces the absorption of water, which can lead to loss of the crunchy texture. Some edible coatings can be applied to food to be fried, so that the coatings act as a barrier to frying oil, reducing the oil absorption by the product and consequently its final fat content; in this case, coatings should be highly hydrophilic to have a good barrier against the hydrophobic oils. Edible films and coatings can also be applied between food components, such as between the crust and the sauce and toppings of a pizza, minimizing the moisture transfer from the sauce and toppings to the crust.

Another requirement for edible films and coatings is that they are tasteless, and do not interfere with the sensory characteristics of the food product. But there are cases where the edible films or coatings are supposed to have a characteristic desirable flavour, which should be compatible with the food to be coated. This is the case, for instance, for fruit purée-based edible films (Azeredo et al., 2009; Mild et al., 2011; Sothornvit and Pitak, 2007) and coatings (Azeredo et al., 2012b; Sothornvit and Rodsamran, 2008), fruit pomace-based edible films (Park and Zhao, 2006), and vegetable purée edible films (Wang et al., 2011).

There are a number of methods for evaluating the physical properties of the resulting film and coatings materials, such as tensile test analyses and dynamic mechanical thermal analysis (DMTA) for mechanical properties (Moates et al., 2001; Siracusa et al., 2008), determination of barrier properties (Siracusa et al., 2008), differential scanning calorimetry (DSC) for thermal properties (Abdorreza et al., 2011), FTIR spectroscopy for interactions between components (Paes et al., 2010; Yakimets et al., 2007), scanning electron microscopy for ultrastructural analyses (Bilbao-Sáinz et al., 2010), NMR spectroscopy for studying crosslinking mechanism (Zhang et al., 2010a), polymer–permeant interactions and their effects on polymer organization (Karbowiak et al., 2008, 2009), and dielectric thermal analysis (DETA) for dielectric behaviour of the films (Moates et al., 2001). However, those techniques have not been covered in this review, which is rather focused on the biomaterials which can be used for food packaging (especially films and coatings), their basic properties and interactions, and how they affect the end materials.

26.2 Components of packaging films and coatings from the biomass

All packaging films and coatings should have at least two components: a matrix, which usually consists of a macromolecule able to form a cohesive structure, and a plasticizer, usually required for reducing rigidity and brittleness inherent to most matrices. Additionally, some other components can be incorporated to improve the physical properties of films and coatings, such as their barrier and mechanical properties, or their resistance to moisture. Figure 26.1 presents a scheme of the basic and optional components of biomass-derived films and coatings.

image
26.1 Basic and optional components for films and coatings.

26.3 Processes for producing bio-based films

Two basic technological approaches comprising wet and dry processes are generally used to produce biomaterials for packaging purposes.

The wet processes are based on separation of macromolecules from a solvent phase, usually by evaporation of the solvent. Usually the wet processes consist of casting a previously homogenized and vacuum degassed film forming dispersion (containing at least a biopolymer matrix, a solvent and usually a plasticizer) on a suitable base material (from which the film can be easily peeled off) and later drying to a moisture content (usually 5–8%) which is optimal for peeling the film away (Lagrain et al., 2010; Tharanathan, 2003), as indicated in Fig. 26.2. Film formation generally involves inter- and intramolecular associations or crosslinking of polymer chains forming a network that entraps and immobilizes the solvent. The degree of cohesion depends on polymer structure, solvent used, temperature and the presence of other molecules such as plasticizers, reinforcements, etc. (Tharanathan, 2003).

image
26.2 Basic scheme of the steps of a casting process to produce films.

Two continuous film casting methods are typically used to manufacture biopolymer films by wet processes: (a) casting on steel belt conveyors and (b) casting on a disposable substrate on a coating line. In (a), solutions are spread uniformly on a continuous steel belt that passes through a drying chamber. The dry film is then stripped from the steel belt and wound into mill rolls for later conversion. One of the advantages of steel belt conveyors is the ability to cast aqueous solutions directly onto the belt surface, optimizing uniformity, heat transfer and drying efficiency, while eliminating expense of a separate substrate such as polyester film or coated paper. In (b), known as web coating, solutions are spread uniformly onto a carrier web or substrate, usually a polyester film or coated paper, and the coated substrate is passed through a drying chamber. The dry film is wound into rolls while still adhering to the substrate, and is usually separated in a secondary operation (Rossman, 2009).

In contrast, the dry processes use the thermoplastic properties of the macromolecules under low moisture conditions. The biomaterials can be shaped by existing plastic processing techniques (the so-called thermoplastic processing technologies), including thermoforming, compression moulding, extrusion, roller milling, or extrusion coating and lamination (Lagrain et al., 2010). Heating amorphous polymers above their glass transition temperature (Tg) changes them into a rubbery state, making it possible to form films after cooling. Although the dry processes require more extensive and advanced equipment, they are efficient in large-scale production due to the low moisture contents, high temperatures, high pressures and short process times (Hernandez-Izquierdo and Krochta, 2008). Furthermore, the materials obtained are likely to exhibit more robust tensile characteristics compared with films cast with the use of plasticizers (Mangavel et al., 2004) and lower water solubility, due to the creation of a highly crosslinked film network (Rhim and Ng, 2007).

Most conventional plastics, such as low density polyethylene films, are produced by extrusion. An extruder consists of a heated, fixed metal barrel containing one or two screws which convey the raw material through the heated barrel, from the feed end to the die (Fig. 26.3). The screws induce shear forces and increasing pressure along the barrel (Verbeek and van den Berg, 2010). Extrusion is one of the most important polymer processing techniques, offering several advantages over solution casting (Hernandez-Izquierdo and Krochta, 2008). However, many bio-based films are more difficult to produce by dry processes when compared to petroleum-based polymers, as they do not usually have defined melting points (due to their heterogeneous nature) and undergo decomposition upon heating (Tharanathan, 2003). There is a delicate balance required so the formulations resist the process conditions and at the same time achieve the desired film performance (Rossman, 2009). Starch (Pushpadass et al., 2008; Thunwall et al., 2008) and protein films (Hochstetter et al., 2006; Hernandez-Izquierdo et al., 2008; Kumar et al., 2010) have been produced by extrusion. Some proteins which exhibit thermoplastic behaviour can be processed without further treatment, but other proteins and starch should be plasticized before processing (Rhim and Ng, 2007).

image
26.3 Schematic representation of an extruder. (adapted from Li et al., 2011b)

26.4 Processes for producing edible coatings

The processes described in the literature for producing and applying edible coatings are restricted to laboratory discontinuous and small-scale techniques. Similarly to a basic casting method, the coating dispersion is produced by homogenization and vacuum degassing of a film forming dispersion (or by melting, for lipid-based coatings). The dispersion is then applied directly on a food surface (or between food components) by dipping or spraying (Fig. 26.4). Dipping is more adequate when an irregular surface has to be coated. After dipping, excess coating material is allowed to drain from the product. Spraying is more adequate for thinner coatings materials and/or when only one side of the product is supposed to be coated (such as in pizza crusts, as exemplified before). Both spraying and dipping are followed by drying for polysaccharide or protein coatings, or by cooling for lipid-based coatings.

image
26.4 Basic steps to prepare and apply a coating to a food surface.

26.5 Products from biomass as film and/or coating matrices

There has been increasing interest in the search for new uses of biomass byproducts from the food industry. Many of these byproducts contain potential film-forming macromolecules such as polysaccharides or proteins which present opportunities for the design of bioplastics to be used as packaging materials.

A downside to the use of biomaterials for packaging purposes is that their inherent properties are usually inferior to those of petrochemical-based systems. However, unlike the conventional polymers, they are bio-degradable. This means that they can either be disposed of through, for example, composting or anaerobic digestion, or might even be exploited as a source of fermentable sugars after enzymatic digestion. There is therefore increasing support for their use in order to reduce the huge volume of plastic waste continually generated by food packaging disposal. According to Van der Zee (2005), the correlations between polymer structure and biodegradability have been proved challenging, since interplays between different factors occur simultaneously, often making it difficult to establish correlations, and creating exceptions when an apparent rule was expected to be followed. For instance, since the first step in biodegradation involves the action of extracellular water-borne enzymes, hydrophilicity favours the biodegradation, and the semicrystalline nature tends to limit it to amorphous regions, although highly crystalline starch materials and bacterial polyesters are rapidly hydrolysed. Some chemical properties that are important include chemical bonds in the polymer backbone, position and chemical activity of side groups, and chemical activity of end groups. Linkages involving hetero atoms, such as ester and amide (or peptide) bonds are considered susceptible to enzymatic degradation, although there are exceptions such as polyamides and aromatic polyesters.

This chapter overviews some of the macromolecules that can be obtained from biomass byproducts which have been used as matrices for food packaging materials, as well as other biomass-derived compounds which can be used as additives, crosslinking agents or reinforcements to these matrices, improving their properties and potential applicability as food packaging materials.

26.5.1 Lignocellulosic biomass

Cellulose and its derivatives

Cellulose, the most abundant biopolymer, is formed by the repeated connection of D-glucose building blocks. Adjacent cellulose chains form a framework of aggregates (elementary fibrils) containing crystalline and amorphous regions; the crystalline regions are maintained by inter- and intramolecular hydrogen bonding. Several elementary fibrils can associate with each other to form cellulose crystallites, which are then held together by a monolayer of hemicelluloses, generating thread-like structures which are enclosed in a matrix of hemicellulose and protolignin, forming a natural composite referred to as cellulose microfibril (Ramos, 2003).

Cellulose represents about a third of the plant cell wall composition, and it is also produced by a family of sea animals called tunicates (sea squirts), by several species of algae, and by some species of bacteria and fungi (Charreau et al., 2013). Cellulose is an important structural component characterized by its hydrophilicity, chirality, biodegradability, broad capacity for chemical modification, and its formation of versatile semicrystalline fibre morphologies.

Together with starch, cellulose and its derivatives (such as ethers and esters) are the most important raw materials for elaboration of biodegradable and edible films (Peressini et al., 2003). Cellulose is an essentially linear natural polymer of (1 → 4)-β-D-glucopyranosyl units. Its tightly packed polymer chains and highly crystalline structure makes it insoluble in water. Water solubility can be conferred by etherification; the water-soluble cellulose ethers, including methyl cellulose (MC), hydroxypropyl cellulose (HPC), hydroxypropylmethyl cellulose (HPMC), and carboxymethyl cellulose (CMC), have good film-forming properties (Cha and Chinnan, 2004; Janjarasskul and Krochta, 2010).

Cellulose films prepared from aqueous alkali/urea solutions were reported to exhibit better oxygen barrier properties when compared to those of conventional cellophane and PVC (Yang et al., 2011). Cellulose derivatives have been used as coatings, extending the shelf life of avocados (Maftoonazad and Ramaswamy, 2005; Maftoonazad et al., 2008) and fresh eggs (Suppakul et al., 2010). Cellulose-based films have been produced by several companies such as Innovia (UK), FKuR (Germany) and Daicel Polymer (Japan).

Hemicelluloses

Hemicelluloses (HC) are heteropolysaccharides closely associated with cellulose in the plant cell walls (Mikkonen and Tenkanen, 2012). They are usually defined as the alkali-soluble material after the removal of pectic substances from plant cell walls (Sun et al., 2004). According to Scheller and Ulvskov (2010), HC are polysaccharides in plant cell walls that have β-(1 → 4)-linked backbones with an equatorial configuration. They have different structures which may contain glucose, xylose, mannose, galactose, arabinose, fucose, as well as glucuronic and galacturonic acids in different proportions, depending on the source (Ebringerová and Heinze, 2000). Hemicelluloses (as well as lignin) cover cellulose microfibrils. The structural similarity between HC and cellulose generates a conformational homology leading to a hydrogen-bonded network between HC and cellulose microfibrils (O’Neill and York, 2003; Rose and Bennett, 1999).

According to their primary structure, hemicelluloses can be categorized into four main groups: xyloglycans (xylans), mannoglycans (mannans), β-glucans and xyloglucans (Ebringerová et al., 2005). Xylans usually consist of a β(1 → 4)-D-xylopyranose backbone with side groups in position 2 or 3. Non-branched homoxylans occur in certain seaweeds, and heteroxylans include glucuronoxylans, arabinoxylans and more complex structures (Hansen and Plackett, 2008). Mannans comprise galactomannans and glucomannans. Galactomannans consist of a β(1 → 4)-linked mannopyranose backbone highly substituted with β(1 → 6)-linked galactopyranose residues (Wyman et al., 2005), while glucomannans consist of alternating β-D-glucopyranosyl and β-D-mannopyranosyl units attached with (1 → 4) bonds (Mikkonen and Tenkanen, 2012). β-glucans have a D-glucopyranose backbone with mixed β linkages (1 → 3, 1 → 4) in different ratios (Ebringerová et al., 2005). Finally, xyloglucans have a backbone of β(1 → 4)-linked D-glycopyranose residues with a distribution of D-xylopyranose in position 6 (Hansen and Plackett, 2008).

Some studies have been conducted on arabinoxylan-based films. Höije et al. (2005) obtained strong but highly hygroscopic arabinoxylan films from barley husks. β-glucan films were reported to be more compact than arabinoxylan films, with smaller nanopores, favouring the barrier properties (Ying et al., 2011). Zhang et al. (2011) reported that the properties of arabinoxylan films were well correlated with the arabinose/xylose (Ara/Xyl) ratios. More crystalline films with lower water uptake resulted from lower arabinose contents. On the other hand, a lower chain mobility was observed in the amorphous parts for highly substituted xylans.

Galactomannan films have also been the subject of several studies. Mikkonen et al. (2007) reported that galactomannans with lower galactose content produced films with higher elongation at break and tensile strength, probably because galactomannans with fewer side chains can interact with other polysaccharides due to their long blocks of unsubstituted mannose units (Srivastava and Kapoor, 2005). Cerqueira et al. (2009) successfully used galactomannans to coat different tropical fruits, choosing the formulations taking into account parameters such as wettability, barrier to gases and mechanical properties. Cheeses have been reported to have their shelf life extended by application of galactomannan-based coatings (Cerqueira et al., 2010; Martins et al., 2010).

Pectins

Pectins are water-soluble anionic heteropolysaccharides composed mainly of (1 → 4)-α-D-galactopyranosyluronic acid units, in which some carboxyl groups of galacturonic acid are esterified with methanol (Fig. 26.5). They are extracted from citrus peels and apple pomace by hot dilute mineral acid. Short hydrolysis times produce pectinic acids and high-methoxyl pectins (HMP), while extended acid treatment de-esterifies the methyl esters to pectic acids and generates low-methoxyl pectins (LMP). Commercial pectins are categorized according to their degree of esterification (DE), defined as the ratio of esterified to total galacturonic acid groups (Sriamornsak, 2003). HMP have a DE > 50 (usually > 69), whereas LMP have a DE < 50 (Farris et al., 2009). The ratio of esterified to non-esterified galacturonic acid determines the behaviour of pectin in food applications, since it affects solubility and gelation properties of pectin. HMP form gels with sugar and acid, whereas LMP form gels in the presence of divalent cations such as Ca2 +, which links adjacent LMP chains via ionic interactions, forming a tridimensional network (Janjarasskul and Krochta, 2010).

image
26.5 A fragment of pectin containing esterified and non-esterified carboxyl groups of galacturonic acid.

Pectin films, similarly to other polysaccharide films, have poor water resistance, and have been proposed for potential industrial uses where water binding either is not a problem or can provide specific advantages, such as edible bags for soup ingredients (Fishman et al., 2000). LMP films, on the other hand, when crosslinked with calcium ions, have not only improved water resistance, but also improved mechanical and barrier properties (Kang et al., 2005). Moreover, the presence of carboxyl groups carrying a negative charge at pH > pKa enables exploiting electrostatic interactions of pectin with positively charged counterparts (Farris et al., 2009).

Lignin

Lignin is the second most abundant terrestrial biopolymer after cellulose, accounting for approximately 30% of the organic carbon in the biosphere (Boerjan et al., 2003). It is associated with cellulose and hemicelluloses in plant cell walls, and is an abundant waste product in the pulp and paper industry.

Lignins are complex aromatic heteropolymers derived mainly from three hydroxycinnamyl alcohol monomers (monolignols) differing in their degree of methoxylation (Fig. 26.6): p-coumaryl, coniferyl and sinapyl alcohols (Boerjan et al., 2003; Buranov and Mazza, 2008).

image
26.6 Lignin monolignols. (from Buranov and Mazza, 2008)

Lignin has some interesting properties to be used for packaging films, such as its small particle size, hydrophobicity and ability to form stable mixtures (Park et al., 2008). Moreover, lignins have been shown to have efficient antibacterial and antioxidant properties (Ugartondo et al., 2009). Lignin has been used as a film component in composites with gelatin (Núñez-Flores et al., 2013; Ojagh et al., 2011; Vengal and Srikumar, 2005), starch (Baumberger et al., 1998; Vengal and Srikumar, 2005) and chitosan (Chen et al., 2009). Baumberger et al. (1998) observed that starch films incorporated with up to 30% lignin presented higher elongation and water resistance than control starch films. On the other hand, lignin impaired the tensile strength of the films at high relative humidity (71%), reflecting the incompatibility between the hydrophilic starch and the hydrophobic lignin, which was bolstered by the water. Indeed, microscopic observations confirmed that the material consisted of two phases – a hydrophilic starch matrix filled with hydrophobic lignin aggregates. On the other hand, Chen et al. (2009) reported a good dispersion of lignin (up to 20%) in a chitosan matrix, evidenced by SEM, which was corroborated by FTIR results, indicating the existence of hydrogen bonding between chitosan and lignin. An FTIR spectroscopy study on gelatin-lignin films revealed strong protein conformational changes induced by lignin, producing a plasticizing effect, which was reflected in the mechanical and thermal properties (Núñez-Flores et al., 2013).

A drawback from incorporating lignin in films is that they acquire a brownish colour (Mishra et al., 2007; Núñez-Flores et al., 2013). On the other hand, they acquire better light barrier properties (Núñez-Flores et al., 2013), which could be of interest in food applications when ultraviolet-induced lipid oxidation is a problem.

26.5.2 Polysaccharides (other than cell wall polysaccharides)

Polysaccharides are hydrophilic polymers and therefore exhibit very low moisture barrier properties. They are of prime interest as matrices for biodegradable film formation because of their availability and rather low cost.

Most polysaccharides are neutral, although some gums are negatively charged. As a consequence of the large number of hydroxyl and other polar groups in their structure, hydrogen bonds play important roles in film formation and characteristics. Negatively charged gums, such as alginate, pectin and carboxymethyl cellulose, tend to present some different properties depending on the pH (Han and Gennadios, 2005).

Polysaccharide films are usually formed by disrupting interactions among polymer segments during coacervation and forming new inter-molecular hydrophilic and hydrogen bonds upon evaporation of the solvent (Janjarasskul and Krochta, 2010). Because of their hydrophilicity, polysaccharide films provide a good barrier to CO2 and O2, hence they retard respiration and ripening of fruits (Cha and Chinnan, 2004). On the other hand, similarly to other hydrophilic materials, their high polarity determines their poor barrier to water vapour (Park and Chinnan, 1995) as well as their sensitivity to moisture, which may affect their functional properties (Janjarasskul and Krochta, 2010).

Starches

Starches are polymers of D-glucopyranosyl, consisting of a mixture of the predominantly linear amylose and the highly branched amylopectin (Fig. 26.7). Native starch molecules arrange themselves in semi-crystalline granules in which amylose and amylopectin are linked by hydrogen bonding. When heat is applied to native starch in the presence of plasticizers such as water and glycerol, the granules swell and hydrate, which triggers the gelatinization process, characterized by loss of crystallinity and molecular order, followed by a dramatic increase in viscosity (Kramer, 2009). This transformation is named gelatinization and leads to the so-called thermo-plastic starch (TPS) (Huneault and Li, 2007).

image
26.7 Chemical structure of a starch fragment.

Amylose responds to the film-forming capacity of starches, since linear chains of amylose in solution tend to interact by hydrogen bonds and, consequently, amylose gels and films are stiff, cohesive and relatively strong. On the other hand, amylopectin films are brittle and non-continuous, since branched amylopectin chains in solution present little tendency to interact (Peressini et al., 2003).

The first studies using starch in biodegradable food packaging were focused on substituting part of the synthetic matrix (usually polyethylene) by starch, but there were difficulties ascribed to chemical incompatibility between the polymers. Recently, studies on pure starch-based materials have predominated, usually focusing on two major drawbacks related to application of starch films. The first one is that native starch commonly exists as granules with about 15–45% crystallinity, and starch-based materials are susceptible to ageing and starch re-crystallization (retrogradation) (Forssell et al., 1999; Ma et al., 2006), which makes starch rigid and brittle during long-term storage, restricting its applications (Huang et al., 2005). The second is their high hydrophilicity, causing its barrier properties to decrease with increasing relative humidity. Plasticizers are used to overcome the first drawback and improve material flexibility (Moates et al., 2001; Peressini et al., 2003; Mali et al., 2004), but, since they are usually highly hydrophilic, presenting hydroxyl groups capable of interacting with water by hydrogen bonds, they tend to increase the moisture affinity of the films (Mali et al., 2005).

Some companies have commercially produced starch-based packaging materials, such as Eco-Go (Thailand), Plantic (Australia) in a joint venture with Du Pont (USA), JMP (Australia), StarchTech (USA), Biome (UK), and BASF (Germany).

Alginates

Alginates, which are extracted from brown seaweeds, are salts of alginic acid, a linear co-polymer of D-mannuronic and L-guluronic acid monomers (Fig. 26.8), containing homogeneous poly-mannuronic and poly-guluronic acid blocks (M and G blocks, respectively) and MG blocks containing both uronic acids. The presence of carboxyl groups in each constituent residue (Ikeda et al., 2000) enables sodium alginate to crosslink with di- or trivalent metal cations, especially calcium ions (Ca2 +), to produce strong gels or films (Cha and Chinnan, 2004). Calcium ions pull alginate chains together via ionic interactions, after which interchain hydrogen bonding occurs (Kester and Fennema, 1986). Films can be formed either from evaporating water from an alginate gel or by a two-step procedure involving drying of alginate solution followed by treatment with a calcium salt solution to induce crosslinking (Janjarasskul and Krochta, 2010). The strength and permeability of films may be altered by changing calcium concentration and temperature, among other factors (Kester and Fennema, 1986). Alginate films have been studied as edible coatings to be applied to a variety of foods such as fruits/vegetables (Fan et al., 2009; Fayaz et al., 2009) and meat products (Marcos et al., 2008; Chidanandaiah et al., 2009).

image
26.8 Basic structure of alginates, containing units of mannuronic (M) and guluronic (G) acids.

Chitosan

Chitosan is a linear polysaccharide consisting of β-(1 → 4)-linked residues of N-acetyl-2-amino-2-deoxy-D-glucose (glucosamine) and 2-amino-2-deoxy-D-glucose (N-acetyl-glucosamine) (Fig. 26.9). It is produced from partial deacetylation of chitin, which is considered as the second most abundant polysaccharide in nature after cellulose (Dutta et al., 2009; Aranaz et al., 2010). Chitin is present in the exoskeleton of crustacea and insects, and can also be found in the cell wall of certain groups of fungi, particularly zygomycetes (Chatterjee et al., 2005). It is usually extracted from crab and shrimp shells as a byproduct of the seafood industry. Since the deacetylation of chitin is usually incomplete, chitosan is a copolymer comprising D-glucosamine and N-acetyl-D-glucosamine with various fractions of acetylated units (Aranaz et al., 2010).

image
26.9 Chemical structure of chitosan.

Chitosan is soluble in diluted aqueous acidic solutions due to the protonation of –NH2 groups at the C2 position (Aranaz et al., 2010). The cationic character confers unique properties to the polymer, such as antimicrobial activity and the ability to carry and slow-release functional ingredients (Coma et al., 2002). The charge density depends on the degree of deacetylation as well as the pH. Quaternization of the nitrogen atoms of amino groups has been a usual chitosan modification, whose objective is to introduce permanent positive charges along the polymer chains, providing the molecule with a cationic character independent of the aqueous medium pH (Curti et al., 2003; Aranaz et al., 2010).

Chitosan films have been proven effective in extending the shelf life of fruits (Hernández-Muñoz et al., 2006; Chien et al., 2007; Lin et al., 2011; Ali et al., 2011) and to have retarded microbial growth on fruit surfaces (Hernández-Muñoz et al., 2006; Chien et al., 2007; Campaniello et al., 2008). The polycationic structure of chitosan probably interacts with the predominantly anionic components (lipopolysaccharides, proteins) of microbial cell membranes, especially Gram-negative bacteria (Helander et al., 2001).

Less conventional polysaccharides

Several other polysaccharides are found in nature, and it is virtually impossible to mention all of them. But some examples are given in this section of less common polysaccharides which have been tested or suggested as food packaging materials.

Carrageenans

Some red algae species (Rhodophyta) have a family of polysaccharides called carrageenans as cell wall polysaccharides (Van de Velde et al., 2002). They are hydrophilic linear sulfated galactans consisting of alternating (1 → 3)-β-D-galactopyranose (G-units) and (1 → 4)-α-D-galactopyranose (D-units), forming a disaccharide repeating unit (Campo et al., 2009). Carrageenan-based coatings have been proven to be efficient to increased stability of fresh-cut (Bico et al., 2009) and fresh whole fruits (McGuire and Baldwin, 1998; Ribeiro et al., 2007). Ribeiro et al. (2007) observed that carrageenan coatings resulted in lower weight loss and lower loss of firmness of strawberries when compared to starch coatings, probably reflecting a better moisture barrier of carrageenan coatings. Moreover, carrageenan films presented a significantly lower oxygen permeability than starch films.

Tree exudates

Natural gums are obtained as exudates from different tree species, including gum acacia, cashew tree gum, and mesquite gum. The tree gums have been grouped into three types based on the nature of polysaccharide type, namely, arabinogalactans, substituted glucuronomannans or substituted rhamnogalacturonans (Sims and Furneaux, 2003). Cashew tree gum (CTG) is a heteropolysaccharide exudated from the cashew tree (Anacardium occidentale) bark (Miranda, 2009), whose composition is made up of galactose, glucose, arabinose, rhamnose and glucuronic acid (De Paula et al., 1998). The greatest cultivation of cashew trees can be found in Brazil, and it is mainly focused on cashew nut production (Bezerra et al., 2007). CTG films have been obtained by Carneiro-da-Cunha et al. (2009) and suggested to be applied as apple coatings. Azeredo et al. (2012a) obtained alginate–CTG blend films crosslinked with CaCl2. CTG reduced tensile strength and barrier properties of the films, but favoured film extensibility. Some studies have described the use of mesquite gum as film-forming matrices (Osés et al., 2009; Bosquez-Molina et al., 2010). Gum acacia coatings have been proven to extend the shelf life of tomatoes (Ali et al., 2010) and shiitake mushrooms (Jiang et al., 2013).

Cactus mucilages

Some hetero-polysaccharides can be obtained from cactus stems, which is waste from cactus pruning. The mucilage extracted from stems of prickly pear cactus (Opuntia ficus-indica), which constitutes about 14% of the cladode dry weight (Ginestra et al., 2009), has been reported to contain residues of D-galactose,D-xylose,L-arabinose,L-rhamnose and D-galacturonic acid (McGarvie and Parolis, 1979). Del-Valle et al. (2005) studied the use of prickly pear mucilage as an edible coating to strawberries, and reported that coated strawberries presented extended physical and sensory stability when compared to uncoated ones.

Bacterial cellulose

Whilst ‘cellulose’ is a word originally given to the substance which constitutes a key load-bearing component of the cell wall of higher plants, bacterial cellulose (BC) is an extracellular product synthesized by bacteria belonging to some genera, its most efficient producers being the Gram-negative, acetic acid bacteria Gluconacetobacter xylinum (Iguchi et al., 2000; Retegi et al., 2010). These microfibril bundles have excellent intrinsic properties due to their high crystallinity, including a reported elastic modulus of 78 GPa (Guhados et al., 2005). Compared with cellulose from plants, BC has important structural differences and it also possesses higher water-holding capacity, higher degree of polymerization (up to 8,000), and a finer web-like network (Klemm et al., 2005). It is produced as a gel, and, although its solid portion is less than 1%, it is almost pure cellulose containing no lignin and other foreign substances (Iguchi et al., 2000). Despite the identical chemical composition, BC is superior to plant cellulose owing to its purity, high elasticity, and nano-morphology with a large surface area (S. Chang et al., 2012). It is an interesting biomaterial thanks to its fine network, excellent mechanical properties, high water-holding capacity, crystallinity and bio-compatibility (Yan et al., 2008; Putra et al., 2008).

Retegi et al. (2010) obtained compression moulded bacterial cellulose (BC) films with different porosities, generated by different compression pressures. Higher pressures were found to produce films with better final mechanical properties. This behaviour was ascribed to the higher densification, reducing interfibrillar spaces, thus increasing the possibility of interfibrillar bonding zones.

26.5.3 Proteins

Proteins are widely available as biomass byproducts from plants (wheat gluten, maize zein, soybean proteins) and animals (collagen, gelatin, keratin, casein, whey proteins).

Proteins are linear, random copolymers built from up to 20 different monomers. The main mechanism of formation of protein films involves denaturation of the protein initiated by heat, solvents, or change in pH, followed by association of peptide chains through new intermolecular interactions (Janjarasskul and Krochta, 2010).

Proteins are distinguished from polysaccharides in that they have approximately 20 different amino acid monomers, rather than just a few or even one monomer, such as glucose in cellulose and starch. The amino acids are similar in that they contain an amino group (–NH2) and a carboxyl group (–COOH) attached to a central carbon atom, but each amino acid has unique properties conferred by a different side group attached to the central carbon, which can be non-polar, polar uncharged, or polar (positively or negatively) charged at pH 7 (Cheftel et al., 1985; Krochta, 2002). While hydroxyl is the only reactive group in polysaccharides, proteins may be involved in several possible interactions and chemical reactions (Hernandez-Izquierdo and Krochta, 2008), such as chemical reactions through covalent (peptide and disulfide) bonds and non-covalent (ionic, hydrogen, and van der Waals) interactions. Moreover, hydrophobic interactions may occur between non-polar groups of amino acid chains (Kokini et al., 1994). In addition, protein-based films are considered to have high UV barrier properties, owing to their high content of aromatic amino acids which absorb UV light (Mu et al., 2012).

The most unique properties of proteins compared to other film-forming materials are heat denaturation, electrostatic charges, and amphiphilic character. Protein conformation can be affected by many factors, such as charge density and hydrophilic–hydrophobic balance (Han and Gennadios, 2005). Proteins have good film-forming properties and good adherence to hydrophilic surfaces. Protein-derived films provide good barriers to O2 and CO2 but not to water (Cha and Chinnan, 2004). Their barrier and mechanical properties are impaired by moisture owing to their inherent hydrophilic nature (Janjarasskul and Krochta, 2010).

The processing of protein films or other protein-based materials mostly requires three main steps: breaking of intermolecular bonds (non-covalent and covalent, if necessary) by chemical or physical rupturing agents; arranging and orienting polymer chains in the desired conformation; and allowing the formation of new intermolecular bonds and interactions stabilizing the film network (Jerez et al., 2007). Globular proteins are required to unfold and realign before a new three-dimensional network can be formed, and stabilized by new inter- and intra-molecular interactions (Verbeek and van den Berg, 2010).

A material to be extruded receives a considerable amount of mechanical energy, which may affect the characteristics of the final products. Extrusion requires the formation of a melt, which implies processing the protein above its softening point. Proteins have many different functional groups, and consequently a great variety of possible chain interactions that reduce molecular mobility and increase viscosity, resulting in a softening temperature which is often above decomposition temperature. Plasticizers can be useful to reduce the softening temperature, but protein extrusion is usually only possible within a limited range of processing conditions (Verbeek and van den Berg, 2010).

Gelatin

Not a naturally occurring protein, gelatin is produced by partial hydrolysis of collagen, which is the main constituent of animal skin, bone, and connective tissue. The insoluble collagen is treated with dilute acid or alkali, resulting in partial cleavage of the crosslinks, and the structure is broken down to such an extent that the soluble gelatin is formed (Karim and Bhat, 2009).

Gelatin is obtained mainly from pigskin and other mammalian sources, but the marine sources (fish skin and bone) have been increasingly looked upon as possible alternatives to bovine and porcine gelatin, as they are not associated with the risks of bovine spongiform encephalopathy (BSE, or ‘mad cow disease’) outbreaks, and because they are acceptable for religious groups which have restrictions on pig and cow derivatives (Karim and Bhat, 2009). Moreover, the fish industry sector has tried to find new outlets for their skin and bone byproducts. However, gelatins from cold water species, representing the majority of the industrial fisheries, present inferior physical properties (such as lower gelling and melting temperatures and lower modulus) when compared to mammalian gelatin (Haug et al., 2004), which has been largely related to the lower contents of hydroxyproline in collagen from cold water fish species, since hydroxyproline is involved in interchain hydrogen bonding, which stabilizes the triple helical structure of collagen (Wasswa et al., 2007). Gelatins from warm water fish species, like tilapia, on the other hand, have physical properties more similar to those of mammalian gelatins (Sarabia et al., 2000). Moreover, fish gelatins contain more hydrophobic amino acids, so their films show significantly lower water vapour permeability when compared to films produced from mammalian gelatins (Avena-Bustillos et al., 2006).

Gelatin contains a large amount of proline, hydroxyproline, lysine, and hydroxylysine, which can react in an aldol-condensation reaction to form intra- and intermolecular crosslinks among the protein chains (Dangaran et al., 2009). On cooling and dehydration, gelatin films are formed with irreversible conformational changes (Badii and Howell, 2006; Dangaran et al., 2009), with an increased concentration of triple helical structures. The triple helix structure is the basic unit of collagen, from which gelatin is derived. Thus, gelatin molecules partly revert back to the collagen structure during gelation (Gómez-Guillén et al., 2002; Chiou et al., 2006).

Several studies have described the successful use of gelatin to form films by casting (Chiou et al., 2006; Jongjareonrak et al., 2006; Hanani et al., 2012b) as well as dry methods (Hanani et al., 2012a; Krishna et al., 2012).

Milk proteins

Total bovine milk proteins consist of about 80% casein and 20% whey proteins. Whey proteins are those which remain in milk serum after casein coagulation during cheese or casein manufacture. It is a mixture of proteins with diverse functional properties, the main ones being α-lactalbumins, β-lactoglobulins, bovine serum albumin, immunoglobulins, and proteose-peptones (Pérez-Gago and Krochta, 2002). Native whey proteins are globular complexes but become random coils upon denaturation and can form three-dimensional networks to produce biodegradable films (Zhou et al., 2009). Since whey proteins have a high proportion of hydrophilic amino acid in their structure, whey protein films have low tensile strength and high water vapour permeability (McHugh et al., 1994), but these properties might be improved by combining whey protein with materials with better tensile strength and hydrophobic properties, such as zein (Ghanbarzadeh and Oromiehi, 2008). The water vapour permeability of whey protein films was decreased by addition of beeswax, by both increasing the hydrophobic character of the film and decreasing the amount of hydrophilic plasticizer (glycerol) required (Talens and Krochta, 2005). Whey protein coatings have been used to increase the stability of fruits (Pérez-Gago et al., 2003; Reinoso et al., 2008), meat products (Shon and Chin, 2008), nuts (Min and Krochta, 2007) and eggs (Caner, 2005).

Casein is a phosphoprotein, which can be separated into various electrophoretic fractions such as α-(s1)- and α-(s2)-caseins, β-casein, and κ-casein, all of them having low solubility at low pH (pI = 4.6) (Müller-Buschbaum et al., 2006). Caseins form films from aqueous solutions without further treatment due to their random-coil nature and their ability to form extensive hydrogen bonds (Lacroix and Cooksey, 2005), as well as electrostatic interactions (Gennadios et al., 1994). Casein film solubility can be decreased by buffer treatments at the isoelectric point (Chen, 2002), by physical crosslinking using irradiation (Vachon et al., 2000) and by chemical crosslinking using aldehydes (Ghosh et al., 2009). Moreover, casein precipitated with high pressure CO2 was reported to present lower water solubility than acid-precipitated casein (Dangaran et al., 2006). Casein films have been used to increase stability of bread (Schou et al., 2005).

Zein

Zein is the name given to the prolamin (alcohol-soluble) fraction of the maize proteins (Ghanbarzadeh and Oromiehi, 2008), representing about 50% of the maize endosperm proteins (Biswas et al., 2005). Zein is commercially produced from maize gluten, a co-product of the maize starch production (Ghanbarzadeh and Oromiehi, 2008). Moreover, the worldwide growth of the bioethanol industry has resulted in a huge increase in zein availability, which has motivated the development of new applications for this protein (Biswas et al., 2009).

Zein films are usually prepared by dissolving zein in aqueous ethyl alcohol (Ghanbarzadeh et al., 2007). Zein is rich in non-polar amino acids, with low proportions of basic and acidic amino acids, which confers lower water vapour permeability and better water resistance to zein films when compared to other protein films (Dangaran et al., 2009; Ghanbarzadeh and Oromiehi, 2008).

Gluten

Wheat gluten proteins are a byproduct of starch extraction from wheat flour, which is commonly used as a functional ingredient, especially in bakery products. Wheat gluten consists of a mixture of proteins that can be classified into two types: the water insoluble glutenins, which comprise proteins with multiple peptide chains linked via interchain disulfide bonds, forming a continuous network that provides strength and elasticity; and the water soluble gliadins, which consist of single polypeptide chains associated via hydrogen bonding, hydrophobic interactions and intramolecular disulfide bonds, acting as a plasticizer to glutenin network and conferring viscosity to gluten (Balaguer et al., 2011a; Goesaert et al., 2005; Hernández-Muñoz et al., 2003).

The low quality gluten is unsuitable for flour improvement in breadmaking, but can be used for preparing plastic films, presenting adequate viscoelasticity, adhesiveness, thermoplasticity, and good film-forming properties, although the glutenins tend to aggregate upon shearing and heating (Balaguer et al., 2011a).

Gluten shows very low solubility in water, because of its low content of amino acids with ionizable side chains and high contents of non-polar amino acids and glutamine, which has a high hydrogen-bonding potential (Lagrain et al., 2010). A complex solvent system with basic or acidic conditions in the presence of alcohol and disulfide bond-reducing agents is required to prepare casting solutions (Cuq et al., 1998). Some aspects regarding gluten extrusion were approached by Lagrain et al. (2010). In contrast to what happens to thermoplastic materials, gluten viscosity does not decrease upon heating, but rather levels off or even increases due to crosslinking reactions. Therefore, gluten extrusion is only possible in a limited window of operating conditions ranging from the onset of protein flow to aggregation and eventually extensive depolymerization (Redl et al., 1999; Zhang et al., 2005), which occurs at rather low temperatures when compared to most synthetic polymers. Gluten materials are thus usually extruded between 80 and 130°C. The formation of a gluten network involves the dissociation and unraveling of the gluten proteins, which allows both glutenin and gliadin to recombine and crosslink through specific linkages in an oriented pattern (Hernandez-Izquierdo and Krochta, 2008), predominant reactions including SH oxidation and SH/SS interchange reactions leading to the formation of SS crosslinks (Lagrain et al., 2010). Overall, gluten-based materials are stable three-dimensional macromolecular networks stabilized by low-energy interactions and strengthened by covalent bonds such as SS bonds between cysteine residues (Lagrain et al., 2010).

26.5.4 Lipids

Unlike other macromolecules, lipids are not biopolymers, being unable to form cohesive, self-supporting films for packaging. So they are used either as coatings directly applied to food to provide a moisture barrier, or as components in stand-alone composite emulsion films, in which proteins or polysaccharides provide structural integrity and lipids respond for hydrophobic character, improving the water vapour barrier (Krochta, 2002). Moreover, the presence of lipids in composite formulations provides an appealing glassy finish to the material.

Polarity of lipids depends on the distribution of chemical groups, the length of aliphatic chains and presence and degree of unsaturation (Morillon et al., 2002). Lipids with longer chains and lower unsaturation and branching degrees have better barrier against water vapour, because of their lower polarity (Rhim and Shellhammer, 2005; Janjarasskul and Krochta, 2010). On the other hand, more polar lipids have more affinity for proteins and polysaccharides when in emulsions, producing more homogeneous films and avoiding the separation of a lipid phase (Fabra et al., 2011b).

Several types of lipids can be used for food coating or as film components, such as waxes, triglycerides, as well as di- and monoglycerides. Waxes, which are esters of long-chain aliphatic acids with long-chain aliphatic alcohols (Rhim and Shellhammer, 2005), are especially resistant to water diffusion, because of their very low content of polar groups (Kester and Fennema, 1986) and their high content in long-chain fatty alcohols and alkanes (Morillon et al., 2002). There are a variety of naturally occurring waxes, derived from vegetables (e.g., carnauba, candelilla and sugar cane waxes), minerals (e.g., paraffin and microcrystalline waxes), or animals – including insects (e.g., beeswax, lanolin and wool grease) (Rhim and Shellhammer, 2005). Some studies reported extension of shelf life of fruits resulting from application of wax-based coatings, including carnauba wax (Dang et al., 2008; Gonçalves et al., 2010) and candelilla wax (Saucedo-Pompa et al., 2009).

Most commonly, lipids have been used as hydrophobic components of emulsion films. Lipid-polysaccharide or lipid-protein emulsion films combine the complementary advantages of each component. Polysaccharides and proteins act as matrices, since they have the mechanical properties to form self-supporting films, and have good barrier against gases such as O2 and CO2, while lipids are useful to reduce water vapour permeability. García et al. (2000) demonstrated that lipid addition to starch films decreased their crystalline-amorphous ratio, which is expected to increase film diffusivity and consequently permeability. On the other hand, lipid addition also reduces the hydrophilic–hydrophobic ratio of films, which decreases their water solubility and therefore water vapour permeability. Emulsion films have been reported to enhance stability of fresh-cut (Chiumarelli and Hubinger, 2012) and whole fruits (Bosquez-Molina et al., 2003; Maftoonazad et al., 2007; Navarro-Tarazaga et al., 2011), vegetables (Conforti and Zinck, 2002), nuts (Mehyar et al., 2012), eggs (Wardy et al., 2011) and bakery products (Bravin et al., 2006).

26.6 Products from biomass as film plasticizers

Brittleness is an inherent quality attributed to most biopolymer films, due to extensive intermolecular forces such as hydrogen bonding, electrostatic forces, hydrophobic bonding, and disulfide bonding (Sothornvit and Krochta, 2005; Srinivasa et al., 2007). Plasticizers are required to break polymer–polymer interactions (such as hydrogen bonds and van der Waals forces), sometimes forming secondary bonds to polymer chains, causing the distance between adjacent chains to increase (Fig. 26.10), thus lowering the glass transition temperature (Tg) and reducing film rigidity and brittleness. Moreover, plasticizers act as processing aids, since they lower the processing temperature, reduce sticking in moulds, and enhance wetting (Sothornvit and Krochta, 2005). On the other hand, plasticizers increase film permeability, and the decrease in cohesion negatively affects mechanical properties (Sothornvit and Krochta, 2005; Vieira et al., 2011).

image
26.10 Representative scheme of the effect of plasticizers.

Thanks to the low molecular size of plasticizers, they occupy spaces between polymer chains, reducing secondary forces. Most plasticizers contain hydroxyl groups which form hydrogen bonds with biopolymers, changing the three-dimensional polymer organization, reducing the energy required for mobility and the degree of hydrogen bonding between chains, resulting in increasing free volume and molecular mobility (Sothornvit and Krochta, 2005; Vieira et al., 2011).

The increased interest in bio-based packaging materials has been followed by a search for natural-based plasticizers, similarly biodegradable and of low toxicity. Commonly used plasticizers in biodegradable food packaging are polyols, mono-, di- or oligosaccharides, and fatty acids and other lipids.

26.6.1 Polyols

Glycerol can be produced either by microbial fermentation or synthesized chemically from petrochemical feedstock. Additionally, glycerol is a major byproduct of the increasing biodiesel production, thus creating a significant surplus resulting in a sharp decrease in glycerol prices. In this context, application of glycerol for value-added products is a necessity as well as an opportunity for the biodiesel industry (Johnson and Taconi, 2007; Yang et al., 2012). The use of glycerol as plasticizer in biopolymer-based films can be a way to help solve the existing surplus of this byproduct from biodiesel production.

Glycerol as well as other polyols, including sorbitol, propylene glycol, and polypropylene glycol, have great affinity for polysaccharide and protein films, because of their hydrophilicity. Several studies have demonstrated the effectiveness of polyols as plasticizers for polysaccharide-based and protein-based films and coatings (Bergo and Sobral, 2007; Jouki et al., 2013; Ramos et al., 2013;Tapia-Blácido et al., 2013). Qiao et al. (2011) used polyol mixtures including mixtures of glycerol and higher molecular weight polyols (HP) such as xylitol, sorbitol and maltitol. The increase of the molecular weight and the content of HP in the polyol mixture enhanced the thermal stability and mechanical strength of the resulting materials.

26.6.2 Mono- and disaccharides

It is not only molecular size, configuration and total number of functional hydroxyl groups that are important characteristics to be considered for an effective plasticizer, but also its compatibility with the film-forming polymer. Polymer–plasticizer compatibility is necessary to generate a homogeneous mixture without phase separation. It has been suggested that some monosaccharides work as plasticizers in polysaccharide films more effectively than polyols, because of the structure similarity between monosaccharides and polysaccharides (Zhang and Han, 2006). Indeed, Zhang and Han (2006) observed that starch films plasticized with monosaccharides (glucose, mannose and fructose) presented better overall physical properties when compared to starch films plasticized with polyols (glycerol and sorbitol). On the other hand, polyols (especially glycerol) presented better ability to lower Tg of the films, indicating that polyols are more effective in terms of thermomechanical properties. Other studies have indicated monosaccharides and disaccharides as effective plasticizers in different polysaccharide-based films (Olivas and Barbosa-Cánovas, 2008; Piermaria et al., 2011).Whey protein films plasticized with sucrose presented excellent oxygen barrier properties, but sucrose tended to crystallize with time (Dangaran and Krochta, 2007).

26.6.3 Lipids

Generally, the purpose of adding lipids to films is to reduce their water vapour permeability and/or to provide an attractive gloss. Moreover, incorporating lipids in protein- or polysaccharide-based films may interfere with polymer chain-to-chain interactions and provide flexible domains within the film. Fabra et al. (2008) observed that oleic acid apparently interacted with the protein (sodium caseinate) matrix forming bonds through polar groups, modifiying the interaction balances in the protein network. The result can be a plasticizing effect, including reduction of film strength and increase of film flexibility, as described for whey protein films plasticized with beeswax (Talens and Krochta, 2005), caseinate films with oleic acid (Fabra et al., 2008), and gelatin films with stearic and oleic acids (Limpisophon et al., 2010).

The major drawback of using lipids as plasticizers is their low compatibility with most biopolymer matrices, because of their hydrophobicity. On the other hand, this same characteristic of lipids represents an advantage in which they reduce the water vapour permeability and moisture sensitivity of biopolymer materials.

26.7 Products from biomass as crosslinking agents for packaging materials

Chemical crosslinking is the process of linking polymer chains by covalent bondings, forming tridimensional networks (Fig. 26.11) which reduce the mobility of the structure and usually enhance its mechanical and barrier properties and its water resistance. Chemical crosslinking provides a mechanism for enhancing the performance of biopolymers. The most common crosslinking reagents are symmetrical bifunctional compounds with reactive groups with specificity for functional groups present on the matrix macromolecules (Balaguer et al., 2011a). Low toxicity crosslinking agents have been explored nowadays for use in food packaging materials, such as phenolic compounds and genipin.

image
26.11 Schematic representation of crosslinking of polymer chains.

26.7.1 Phenolic compounds

Several natural phenolic compounds derived from plants have been used as crosslinkers to modify biopolymer films. Several potential interactions may be involved, such as hydrogen bonding, ionic, hydrophobic interactions, and covalent bonding, although covalent bonds are more rigid and thermally stable than other interactions (Zhang et al., 2010a). The postulated chemical pathway involves oxidization of diphenol moieties of phenolic acids or other polyphenols, under alkaline conditions, producing quinone intermediates which react with nucleophiles from reactive amino acid groups such as sulfhydryl groups of cystine, amine groups of lysine and arginine, and amide groups of asparagine and glutamine, forming covalent C–N or C–S bonds between the phenolic ring and proteins. The thus regenerated hydroquinone can be reoxidized and bind a second protein chain, resulting in a crosslink (Strauss and Gibson, 2004; Zhang et al., 2010a).

Caffeic and tannic acids were used as crosslinkers for gelatin (Zhang et al., 2010a), and the use of high-resolution NMR technique confirmed the occurrence of chemical reactions between the phenolic groups in phenolic compounds and amino groups in gelatin, forming covalent C–N bonds; the crosslinking decreased the mobility of the gelatin matrix. A similar study by the same group (Zhang et al., 2010b) indicated that the structure of a gelatin film crosslinked by tannic acid (3 wt%) was stable even under boiling, and that the crosslinking modification enhanced the mechanical properties of the protein.

Several other studies have reported beneficial effects from crosslinking protein matrices with phenolic compounds, such as tannic acid and sorghum condensed tannins (Emmambux et al., 2004), procyanidin (He et al., 2011) and ferulic acid (Ou et al., 2005; Fabra et al., 2011a).

Mathew and Abraham (2008) and Cao et al. (2007) reported significant increases in tensile strength and decreases in elongation at break resulting from adding ferulic acid to starch/chitosan films and gelatin-based films, respectively, which was ascribed to the crosslinking between ferulic acid and the polysaccharides or protein used. Ferulic acid is found as a naturally occurring component in cell walls, crosslinking polysaccharides (especially hemicelluloses) with each other and with other cell wall components such as lignin (Ng et al., 1997; Parker et al., 2005).

26.7.2 Genipin

Genipin (Fig. 26.12) is a hydrolysis product from geniposide, which is a component of traditional Chinese medicine, isolated from gardenia fruits (Gardenial jasminoides Ellis). According to studies by Sung et al. (1999), genipin is about 10,000 times less cytotoxic than glutaraldehyde.

image
26.12 Chemical structure of genipin. (from Yuan et al., 2007)

Genipin reacts with nucleophilic groups such as amino groups, being an adequate crosslinking agent for protein films as well as chitosan. Mi et al. (2005) studied the reaction mechanism of chitosan with genipin, and found that genipin undertakes a ring-opening reaction to form an intermediate aldehyde group resulting from the nucleophilic attack by chitosan amino groups. The genipin molecules reacting with a nucleophilic reagent may further undergo polymerization (Fernandes et al., 2013; Jin et al., 2004; Yuan et al., 2007). A dark-blue coloration appears in crosslinked materials exposed to air, which is associated with the oxygen radical-induced polymerization of genipin as well as its reaction with amino groups (Muzzarelli, 2009). Mi et al. (2005) observed that the colour of genipin crosslinked chitosan membranes varied from original transparent to bluish or brownish, depending on the pH value upon crosslinking. These authors ascribed the colour changes to establishment of different structures of crosslinked chitosan resulting from reaction of original genipin or polymerized genipin with primary amino groups on chitosan. The degrees of crosslinking of the genipin-crosslinked chitosan membranes depended significantly on their crosslinking pH values, being higher around pH 7.

Crosslinking with genipin has been reported to improve mechanical properties, water resistance of chitosan films, although they have turned dark bluish (Jin et al., 2004), which may be a market hindrance to many food packaging applications. Bigi et al. (2002) reported that gelatin films crosslinked with genipin presented higher Young’s modulus, better thermal stability (reflected in higher denaturation temperature), and better water resistance than uncrosslinked films. After 1 month of storage in buffer solution, a small gelatin amount (about 2%) was lost from the films, but their mechanical, thermal and swelling properties were very close to those of gelatin films previously crosslinked by glutaraldehyde (Bigi et al., 2001).

Another geniposide derivative, aglycone geniposidic acid, has been used to crosslink chitosan films, improving their tensile strength and water vapour barrier, although reducing their elongation (Mi et al., 2006).

26.7.3 Aldehydes

Aldehydes such as glutaraldehyde, glyoxal or formaldehyde have been used as crosslinking agents to improve mechanical and barrier properties of protein films. However, due to concerns about possible toxic effects of such aldehydes, naturally occurring, less toxic aldehydes and other crosslinking agents have been explored for use in food packaging.

Cinnamaldehyde is an aromatic unsaturated aldehyde derived from cinnamon, consisting of a phenyl group attached to an unsaturated aldehyde (Fig. 26.13). It has been used as an antimicrobial agent in active packaging applications (Becerril et al., 2007), and it can also act as a crosslinking agent for proteins. Balaguer et al. (2011b) demonstrated the crosslinking effect of cinnamaldehyde for gliadin films, which was ascribed to the formation of intermolecular covalent bonds between polypeptide chains, polymerizing gliadins and reticulating the protein matrix. However, it is still uncertain which functional groups of proteins or other macromolecules have more potential to react with cinnamaldehyde, but Balaguer et al. (2011a) proposed a crosslinking mechanism involving the amino groups of proteins, although not ruling out the participation of other reactive groups.

image
26.13 Chemical structure of cinnamaldehyde. (from Balaguer et al., 2011b)

Gliadin films presented significant improvements in their tensile strength and elastic modulus upon crosslinking with cinnamaldehyde. Such effects, as well as water resistance, were reported to be proportional to the cinnamaldehyde concentration (Balaguer et al., 2011b). The crosslinked films did not disintegrate upon a 5-month immersion in water, although a weight loss was reported, indicating that part of the material was solubilized (Balaguer et al., 2011a).

Dialdehyde polysaccharides have received attention as crosslinking agents of protein films. The oxidation of polysaccharides by periodate is characterized by the cleavage of the C2–C3 bond of glucose residues, resulting in the formation of two aldehyde groups per glucose unit, forming 2,3-dialdehyde polysaccharides (Li et al., 2011a). The aldehyde groups can crosslink with ε-amino groups by C = N bonds, as in lysine or hydroxylysine side groups of gelatin (Fig. 26.14) to improve the properties of protein films (Dawlee et al., 2005; Mu et al., 2012). Dialdehyde starch (DAS) was used as a crosslinking agent for gelatin films (Martucci and Ruseckaite, 2009). Crosslinking with DAS up to 10 wt% enhanced moisture resistance and barrier properties of the films, but higher amounts of DAS conducted to phase separation, impairing transparency and tensile properties. Mu et al. (2012) reported that the addition of dialdehyde carboxymethylcellulose (DCMC) to gelatin films increased their tensile strength and thermal stability and reduced their water sensitivity, while keeping their transparency.

image
26.14 Periodate oxidization of a glucose residue in a polysaccharide and the Schiff’s reaction between gelatin and the dialdehyde polysaccharide. (from Mu et al., 2012)

26.8 Products from biomass as reinforcements for packaging materials

Polymer composites are mixtures of polymers with inorganic or organic fillers with certain geometries (fibres, flakes, spheres, particulates). When the fillers are nanoparticles, that is to say, when they have at least one nanosized dimension (up to 100 nm), the resulting material is a nanocomposite (Alexandre and Dubois, 2000). Polysaccharides are good candidates for renewable and biodegradable nanofillers, because of their partly crystalline structures, conferring good reinforcement effects (Le Corre et al., 2010).

26.8.1 From micro to nanoscale

The change of filler dimensions from micro to nanoscale brings about important advantages concerning the resulting composite materials. Because of their size, nanoparticles have larger surface area-to-volume ratio than their microscale counterparts. A uniform dispersion of nanoparticles leads to a very large matrix/filler interfacial area, which changes the molecular mobility, improving thermal, barrier and mechanical properties of the material. Fillers with a high ratio of the largest to the smallest dimension (i.e., aspect ratio), such as nanofibres, are particularly interesting because of their high specific surface area, providing better reinforcing effects (Azizi Samir et al., 2005; Dalmas et al., 2007). In addition, an interphase region of altered mobility surrounding each nanoparticle is induced by well-dispersed nanoparticles, resulting in a percolating interphase network playing an important role in improving the nanocomposite properties (Qiao and Brinson, 2009). According to Jordan et al. (2005), for a constant filler content, a reduction in particle size increases the number of filler particles, bringing them closer to one another, and causing the interface layers from adjacent particles to overlap, altering the bulk properties significantly.

De Moura et al. (2009) observed that the water vapour permeability of hydroxypropyl methylcellulose (HPMC) films reinforced with chitosan nanoparticles was affected by the size of the CsN – the smaller the particles, the lower the permeability. According to the authors, CsN tended to occupy the empty spaces in the pores of the HPMC matrix, thereby improving tensile and barrier properties.

26.8.2 Cellulosic reinforcements

Cellulose fibres are built up by smaller and mechanically stronger entities (cellulose nanoparticles) which can be extracted under proper conditions. Cellulose nanoparticles (i.e., cellulose elements having at least one dimension in the 1–100 nm range, here referred to as nanocellulose), are inherently a low cost and widely available material. Moreover, they are environmentally friendly, easy to recycle by combustion, and require low energy consumption in manufacturing (Klemm et al., 2005; Charreau et al., 2013).

A uniform dispersion of cellulose nanoparticles on a polymer matrix reduces the molecular mobility, changes the relaxation behaviour, and improves the overall thermal and mechanical properties of the material (Azizi Samir et al., 2005; Charreau et al., 2013; Qiu and Hu, 2013). In this context, nanocellulose has been presented as a promising reinforcing and barrier component for elaboration of low cost, lightweight, and high-strength bionanocomposites for food packaging purposes mainly due to its compatibility with the biopolymers (Helbert et al., 1996; Podsiadlo et al., 2005; Moon et al., 2011). Basically two different classes of nanocellulose can be obtained – whiskers and nanofibrils (Azizi Samir et al., 2005). The term ‘whiskers’ (or ‘nanocrystals’) is used to designate elongated crystalline rod-like nanoparticles, whereas the designation ‘nanofibrils’ is used for long flexible nanoparticles consisting of alternating crystalline and amorphous strings (Abdul Khalil et al., 2012).

Different approaches have been introduced to produce nanocellulose: top-down (nanometric structures are obtained by size reduction of bulk materials) and bottom-up approaches (nanostructures are built from individual atoms or molecules capable of self-assembling) (Yousefi et al., 2013). Any cellulosic material can be virtually considered as a potential source for top-down approaches, including crop residues and agroindustry non-food feedstock. However, variations in cellulose source and its preparation conditions lead to a broad spectrum of structures, properties and applicability, which affect performance of cellulose nanoreinforcements (Kvien and Oksman, 2007; Azeredo, 2009; Dufresne, 2012).

Acid hydrolysis has been the primary method for isolating nanocellulose, consisting basically in removing the amorphous regions present in the fibrils leaving the crystalline regions intact; the dimensions of the whiskers after hydrolysis depend on the percentage of amorphous regions in the bulk fibrils, which varies for each organism (Gardner et al., 2008). The morphology of the obtained nanowhiskers is influenced by acid-to-pulp ratio, reaction time, temperature and cellulose source. In spite of the widely varied dimensions (of 3–70 nm widths and 35–3000 nm lengths) reported from different cellulose sources and hydrolysis conditions, cellulose nanowhiskers typically consist of structures 200–400 nm in length and with an aspect ratio of about 10 (Beck-Candanedo et al., 2005; Elazzouzi-Hafraoui et al., 2008; Rosa et al., 2010). Other methods can be used to extract nanocellulose from the lignocellulosic sources, usually based on successive chemical and mechanical treatments, including high-pressure homogenization (Zimmermann et al., 2010), electrospinning (Konwarh et al., 2013), enzymatic hydrolysis (de Campos et al., 2013), TEMPO-mediated oxidation (Isogai et al., 2011), solvent-based isolation (Yousefi et al., 2011), chemi-mechanical forces (Yousefi et al., 2013), ultrasonication (Chen et al., 2011; de Campos et al., 2013), cryo crushing (Alemdar and Sain, 2008) and steam explosion (Deepa et al., 2011). Such processes usually produce longer nanostructures (typically several micrometers in length), but with less uniform width (5–100 nm) (Siró and Plackett, 2010) and lower crystallinity (Iwamoto et al., 2007).

Although the most important industrial source of cellulosic fibres is wood, crops such as flax, cotton, hemp, sisal and others, especially from by-products of these different plants (corn, wheat, rice, sorghum, barley, sugar cane, pineapple, bananas and coconut crops) are likely to become of increasing interest as sources of nanocellulose. These non-wood plants generally contain less lignin than wood and therefore pre-treatment processes are less demanding (Siró and Plackett, 2010; Rosa et al., 2010).

In contrast to celluloses from plants, which require mechanical or chemomechanical processes to produce nanosized structures, the aforementioned bacterial cellulose (BC) is produced already as a nanomaterial by bacteria through cellulose biosynthesis and the building up of bundles of microfibrils (Nakagaito and Yano, 2005). In addition, BC is produced as a highly hydrated and relatively pure cellulose membrane and therefore no chemical treatments are needed to remove lignin and hemicelluloses, as is the case for plant celluloses (Siró and Plackett, 2010).

Nanocomposites for food packaging purposes have been developed by adding nanocellulose to polymers to enhance their physical and mechanical properties (Paula et al., 2011; Azeredo et al., 2012b; Abdollahi et al., 2013; Zainuddin et al., 2013). Nanocellulose has been reported to have a great effect in improving tensile strength and elastic modulus of polymers, especially at temperatures above the Tg of the matrix polymer (Wu et al., 2007; Azeredo et al., 2012b; Zainuddin et al., 2013; Abdollahi et al., 2013). This effect is ascribed not only to the geometry and stiffness of the nanocellulose, but also to the formation of a fibril network within the polymer matrix, the cellulose fibres probably being linked through hydrogen bonds. Barrier properties of polymer films have also been observed to be improved by cellulose nanostructures (Sanchez-Garcia et al., 2008; Paula et al., 2011; Azeredo et al., 2012b; Abdollahi et al., 2013). The presence of crystalline fibres is thought to increase the tortuosity in the materials leading to slower diffusion processes and, hence, to lower permeability (Sanchez-Garcia et al., 2008). Nanocellulose has also been reported to improve thermal properties of polymers (Petersson et al., 2007; Paula et al., 2011). The performance of nanocellulose has been reported to be strongly related to the content, dimensions and consequent aspect ratios of the nanostructures, as well as to the degree of matrix–cellulose interaction and percolation effects (Petersson and Oksman, 2006; Hubbe et al., 2008; Tang and Liu, 2008; Kim et al., 2009).

Since 2011, pilot plants have been opened for the production of nanocellulose in Sweden, Canada and the United States. These facilities make it possible to produce nanocellulose on a large scale for the first time, and it was an important step towards the industrialization of this technology.

26.8.3 Starch nanoreinforcement

Starch granules can be submitted to an extended-time hydrolysis at temperatures below that of gelatinization, when the amorphous regions are hydrolysed before the crystalline lamellae, which are more resistant to hydrolysis. The nanocrystals thus separated show platelet morphology with thicknesses of 6–8 nm (Kristo and Biliaderis, 2007). To prepare starch nanoparticles instead of nanocrystals, Ma et al. (2008) precipitated a starch solution within ethanol as the precipitant.

Similarly to cellulose nanocrystals, the reinforcing effect of starch nanocrystals (SNC) is usually ascribed to the formation of a percolating network maintained by hydrogen bonds above a given filler content (the percolation threshold) (Le Corre et al., 2010). Although not proven, this phenomenon was evidenced from experiments which indicated a changing behaviour above certain filler concentration (Angellier et al., 2005). It has been suggested that starch nanocrystals, similarly to nanoclays (which have a platelet morphology as well), create a tortuous diffusion pathway for permeant molecules through the nanocomposite materials, improving their barrier properties (Le Corre et al., 2010).

Kristo and Biliaderis (2007) reported that the addition of SNC to pullulan films resulted in improved tensile strength and modulus and decreased water vapour permeability. Moreover, the SNC promoted an increase in Tg, probably because of strong interactions of nanocrystals with one another and with the matrix, restricting chain mobility. Chen et al. (2008) observed that the tensile strength and elongation of polyvinyl alcohol (PVA) were only slightly improved by addition of SNC up to 10 wt% and, above this content, such properties were impaired by SNC. On the other hand, the properties of PVA nanocomposite with SNC were better than those of the composites with native starch, indicating that SNC presented a better dispersion and stronger interactions with the matrix than native starch granules.

26.8.4 Chitin/chitosan nanostructures

Chitin whiskers can be prepared by acid hydrolysis of chitin (Lu et al., 2004; Sriupayo et al., 2005), and have been successfully prepared from different chitin sources such as squid pens (Paillet and Dufresne, 2001), crab shells (Nair and Dufresne, 2003), and shrimp shells (Sriupayo et al., 2005). When the acid hydrolysis is followed by mechanical ultrasonication/disruption, chitin nanoparticles (nanospheres) can be formed rather than chitin whiskers (Chang et al., 2010b). Chitosan nanoparticles can be obtained by ionic gelation of chitosan, where the cationic amino groups of chitosan form electrostatic interactions with polyanionic crosslinking agents, such as tripolyphosphate (López-León et al., 2005).

Some authors reported beneficial effects from adding chitin whiskers to biopolymer films. The chitin whiskers added by Lu et al. (2004) to soy protein isolate (SPI) greatly improved the tensile properties and water resistance of the matrix. Sriupayo et al. (2005) reported that the whiskers improved the water resistance of chitosan films, and enhanced their tensile strength until a content of 2.96%, but impaired the strength when at higher contents. Similarly, Chang et al. (2010b) reported that chitin nanoparticles were uniformly dispersed and presented good interaction with a starch matrix when at low loading levels (up to 5%), improving its mechanical, thermal and barrier properties. However, aggregation of nanoparticles occurred at higher loading, impairing the performance of the matrix.

Chitosan nanoparticles have also been proven to be effective as nanoreinforcement to bio-based films. The incorporation of chitosan-tripolyphosphate (CS-TPP) nanoparticles significantly improved mechanical and barrier properties of hydroxypropyl methylcellulose (HPMC) films (De Moura et al., 2009). The authors attributed such effects to the nanoparticles filling discontinuities in the HPMC matrix. Chang et al. (2010a) added chitosan nanoparticles to a starch matrix. The tensile, barrier and thermal properties of the matrix were improved by low contents of the nanoparticles, when they were well dispersed in the matrix. Such effects were ascribed to close interactions between the nanoparticles and the matrix, due to their chemical similarities. However, higher nanoparticle loads (8% w/w) resulted in their aggregation in the nanocomposites, impairing the physical properties of the materials.

26.9 Future trends

26.9.1 Polymer surface modifications

Since most biopolymers have limitations for their applications as packaging materials, such as water sensitivity, brittleness or poor mechanical performance, chemical modification techniques can sometimes be used to generate new biomaterials with improved properties. Shi et al. (2011) grafted lauryl chloride groups onto zein molecule through an acylation reaction, and obtained seven-fold increase in elongation at break of modified zein. Moreover, the modification increased the zein hydrophobicity, suggesting that the end material presented probably decreased moisture sensitivity and water vapour permeability.

The combination between ultraviolet light and ozone (UV-O treatment) has been suggested as an effective method to modify polymer surfaces, leading to oxidation reactions at the polymer surface, while the bulk properties, such as thermal, barrier and mechanical properties of the polymer, may not be altered (Shi et al., 2009). Shi et al. (2009) used UV-O treatment to control hydrophilicity of zein films. The treatment converted some of the surface methyl groups mainly to carbonyl groups, decreasing the water contact angles and increasing the surface hydrophilicity of zein films. Moreover, the authors suggested that, once the surface has been modified, several active compounds could have been linked to the functional groups formed, and the polymer would be not only a barrier material but a carrier to active compounds, constituting an active packaging material.

26.9.2 Active and bioactive biopackaging

Conventional food packaging systems are supposed to passively protect the food, acting as barriers between the food and the surrounding environment. On the other hand, an active food packaging may be defined as a system that not only acts as a barrier but also interacts with the food in some desirable way, for example by releasing desirable compounds (such as antimicrobial or antioxidant agents), or by removing a detrimental factor (such as oxygen or water vapour), usually to improve food stability, that is to say, to better maintain food quality and safety. The compounds added more frequently are antimicrobials, such as chitosan (Shen et al., 2010), acids (Guillard et al., 2009), phenolic compounds (Cerisuelo et al., 2012; Arrieta et al., 2013) and antimicrobial peptides (Sanjurjo et al., 2006; Gómez-Guillén et al., 2011).

More recently, the concept of bioactive packaging, in which a food packaging or coating has a potential to enhance food impact over consumer health, has been proposed by Lopez-Rubio et al. (2006). Enclosing bioactive compounds within packaging instead of directly incorporating them into food presents some industrial benefits, such as increasing retention of bioactives, reducing some incompatibility problems between the bioactive and the food matrix, and reducing changes to food sensory properties (Lopez-Rubio et al., 2006). Antioxidant packaging materials could be included in both active and bioactive packaging concepts, since antioxidant compounds usually have alleged benefits both for food stability and for consumer health. Some studies have described the incorporation of antioxidant compounds or extracts to biopolymer films and coatings, such as phenolic compounds to zein films (Arcan and Yemenicioglu, 2011), α-tocopherol to chitosan films (Martins et al., 2012), curcumin and ascorbyl dipalmitate to cellulose-based films (Sonkaew et al., 2012), and ferulic acid and α-tocopherol to sodium caseinate films (Fabra et al., 2011a). Other studies have described the development of probiotic films in coatings, based on the incorporation of lactic acid bacteria, such as Bifidobacterium lactis incorporated to alginate and gellan films for coating of fresh fruits (Tapia et al., 2007), Lactobacillus sakei added to caseinate films to control Listeria monocytogenes in fresh beef (Gialamas et al.,2010),Lactobacillus acidophilus in starch-based coatings for breads (Altamirano-Fortoul et al., 2012) and L. acidophilus and Bifidobacterium bifidum incorporated to gelatin coatings for fish (López de Lacey et al., 2012).

The active and bioactive compounds are most frequently simply added to the film-forming formulation, constituting one of the formulation components, but not being chemically linked to the matrix polymer. However, some studies have suggested the covalent immobilization of active compounds onto functionalized polymer surfaces, with many possible applications for food packaging purposes. Biopolymer films may be functionalized in two basic steps, namely, the treatment of the polymer surface in order to produce reactive functional groups, and the reaction of such groups with an active compound (Kugel et al., 2011). Sometimes the functionalization can be favoured by an intermediary step, such as grafting a polyfunctional agent onto the polymer surface, increasing the density of available functional groups, or by a spacer molecule to reduce steric hindrances (Goddard and Hotchkiss, 2007). An active compound which is covalently immobilized onto the packaging material is not supposed to be released, but becomes effective when in contact with the food surface (Han, 2003). The active compounds to be tethered can be enzymes, antimicrobials, biosensors, bioreactors, etc. Some studies have described the covalent attachment of active compounds onto the surface of biopolymer materials to be applied as food packaging or coatings, such as lysozyme onto chitosan coatings (Lian et al., 2012) and N-halamine onto chitosan films (Li et al., 2013).

26.10 Conclusion

The world biomass is rich in macromolecules with film-forming abilities which can be explored to develop materials for food packaging and coating purposes. Several studies have been conducted describing the development of food packaging materials from biomaterials, especially polysaccharides and proteins. However, to convert macromolecules from biomass into materials with both processability and performance compatible with petrochemical-based ones, research needs to address several challenges. Important gaps in knowledge remain on the structure and functionality of biomaterials. Moreover, chemical and physical changes during processing of biomaterials need to be better understood, so that the processing conditions are improved. Another major issue is to minimize the impact of environmental conditions (especially humidity) on material performance. Finally, biomaterials need to be compatible with their petrochemical counterparts in terms of price, in order to penetrate the market. One of the main challenges to make biomaterials commercially viable is their processing by conventional processing techniques used for petroleum-based polymers such as extrusion and injection moulding. The industrial application of edible coatings to food surfaces also requires developments of techniques and equipment adequate to each kind of food to be coated.

Moreover, nanotechnology has demonstrated a great potential to expand the use of biodegradable polymers, since the addition of nanofillers has led to improvements in overall performance of biopolymers, making them more competitive in a market dominated by non-biodegradable materials. However, there are still important safety concerns about nanotechnology applications to food contact materials. Considering the tiny dimensions of nanofillers, it is reasonable to assume that they might migrate from the packaging material to food. Although the properties and safety of most starting materials in their bulk form are usually well known, their nano-sized counterparts frequently exhibit different properties, because their small sizes would allow them to cross more barriers through the body, while their high surface area increases their reactivity. Hence, detailed information is still required to assess the potential toxicity and environmental effects of nanofillers to be incorporated to food packaging materials.

Finally, the biopackaging field has adopted technologies to improve the performance and/or to add functionalities to biopolymer-based materials, so their range of applications can be widened, and they can become more competitive in the polymer market.

26.11 Acknowledgements

The authors wish to acknowledge financial support from EMBRAPA (Brazil) and the BBSRC (UK).

26.12 References

1. Abdollahi M, Alboofetileh M, Rezaei M, Behrooz R. Comparing alginate nanocomposite films reinforced with organic and/or inorganic nanofillers for food packaging. Food Hydrocoll. 2013;32:416–424.

2. Abdorreza MN, Cheng LH, Karim AA. Effects of plasticizers on thermal properties and heat sealability of sago starch films. Food Hydrocoll. 2011;25:56–60.

3. Abdul Khalil HPS, Bhat AH, Ireana Yusra AF. Green composites from sustainable cellulose nanofibrils: a review. Carboh Polym. 2012;87:963–979.

4. Alemdar A, Sain M. Biocomposites from wheat straw nanofibers: morphology, thermal and mechanical properties. Compos Sci Technol. 2008;68:557–565.

5. Alexandre M, Dubois P. Polymer-layered silicate nanocomposites: preparation, properties and uses of a new class of materials. Mater Sci Eng. 2000;28:1–63.

6. Ali A, Maqbool M, Ramachandran S, Alderson PG. Gum arabic as a novel edible coating for enhancing shelf-life and improving postharvest quality of tomato (Solanum lycopersicum L.) fruit. Postharv Biol Technol. 2010;58:42–47.

7. Ali A, Muhammad MT, Sijam K, Siddiqui Y. Effect of chitosan coatings on the physicochemical characteristics of Eksotika II papaya (Carica papaya L.) fruit during cold storage. Food Chem. 2011;124:620–626.

8. Altamirano-Fortoul R, Moreno-Terrazas R, Quezada-Gallo A, Rosell CM. Viability of some probiotic coatings in bread and its effect on the crust mechanical properties. Food Hydrocoll. 2012;29:166–174.

9. Angellier H, Molina-Boisseau S, Dufresne A. Mechanical properties of waxy maize starch nanocrystals reinforced natural rubber. Macromol. 2005;38:9161–9170.

10. Aranaz I, Harris R, Heras A. Chitosan amphiphilic derivatives: chemistry and applications. Curr Org Chem. 2010;14:308–330.

11. Arcan I, Yemenicioglu A. Incorporating phenolic compounds opens a new perspective to use zein films as flexible bioactive packaging materials. Food Res Int. 2011;44:550–556.

12. Arrieta MP, Peltzer MA, Garrigós MC, Jiménez A. Structure and mechanical properties of sodium and calcium caseinate edible active films with carvacrol. J Food Eng. 2013;114:486–494.

13. Avena-Bustillos RJ, Olsen CW, Olson DA, et al. Water vapor permeability of mammalian and fish gelatin films. J Food Sci. 2006;71:202–207.

14. Azeredo HMC. Nanocomposites for food packaging applications. Food Res Int. 2009;42:1240–1253.

15. Azeredo HMC, Mattoso LHC, Wood D, Williams TG, Avena-Bustillos RJ, McHugh TH. Nanocomposites edible films from mango puree reinforced with cellulose nanofibers. J Food Sci. 2009;74:N31–N35.

16. Azeredo HMC, Magalhães US, Oliveira SA, Ribeiro HL, Brito ES, De Moura MR. Tensile and water vapour properties of calcium-crosslinked alginate-cashew tree gum films. Int J Food Sci Technol. 2012a;47:710–715.

17. Azeredo HMC, Miranda KWE, Rosa MF, Nascimento DM, De Moura MR. Edible films from alginate-acerola puree reinforced with cellulose whiskers. LWT – Food Sci Technol. 2012b;46:294–297.

18. Azizi Samir MAS, Alloin F, Dufresne A. Review of recent research into cellulosic whiskers, their properties and their application in nanocomposite field. Biomacromol. 2005;6:612–626.

19. Badii F, Howell NK. Fish gelatin: structure, gelling properties and interaction with egg albumen proteins. Food Hydrocoll. 2006;20:630–640.

20. Balaguer MP, Gómez-Estaca J, Gavara R, Hernandez-Munoz P. Biochemical properties of bioplastics made from wheat gliadins cross-linked with cinnamaldehyde. J Agric Food Chem. 2011a;59:13212–13220.

21. Balaguer MP, Gómez-Estaca J, Gavara R, Hernandez-Munoz P. Functional properties of bioplastics made from wheat gliadins modified with cinnamaldehyde. J Agric Food Chem. 2011b;59:6689–6695.

22. Baumberger S, Lapierre C, Monties B, Della Valle G. Use of kraft lignin as filler for starch films. Polym Degrad Stabil. 1998;59:273–277.

23. BCC Research. Global markets and technologies for bioplastics. Report code PLS050B, available at http://www.bccresearch.com/report/bioplastics-markets-technologies-pls050b.html; 2012.

24. Becerril R, Gomez-Lus R, Goni P, Lopez P, Nerin C. Combination of analytical and microbiological techniques to study the antimicrobial activity of a new active food packaging containing cinnamon or oregano against E. coli and S. aureus. Anal Bioanal Chem. 2007;388:1003–1011.

25. Beck-Candanedo S, Roman M, Gray DG. Effect of reaction conditions on the properties and behavior of wood cellulose nanocrystal suspensions. Biomacromol. 2005;6:1048–1054.

26. Bergo P, Sobral PJA. Effects of plasticizer on physical properties of pigskin gelatin films. Food Hydrocoll. 2007;21:1285–1289.

27. Bezerra MA, Lacerda CF, Gomes Filho E, Abreu CEB, Prisco JT. Physiology of cashew plants grown under adverse conditions. Brazil J Plant Physiol. 2007;19:449–461.

28. Bico SLS, Raposo MFJ, Morais RMSC, Morais AMMB. Combined effects of chemical dip and/or carrageenan coating and/or controlled atmosphere on quality of fresh-cut banana. Food Control. 2009;20:508–514.

29. Bigi A, Cojazzi G, Panzavolta S, Roveri N, Rubini K. Mechanical and thermal properties of gelatin films at different degrees of glutaraldehyde crosslinking. Biomater. 2001;22:763–768.

30. Bigi A, Cojazzi G, Panzavolta S, Roveri N, Rubini K. Stabilization of gelatin films by crosslinking with genipin. Biomater. 2002;23:4827–4832.

31. Bilbao-Sáinz C, Avena-Bustillos RJ, Wood DF, Williams TG, McHugh TH. Composite edible films based on hydroxypropyl methylcellulose reinforced with microcrystalline cellulose nanoparticles. J Agric Food Chem. 2010;58:3753–3760.

32. Biswas A, Sessa DJ, Lawton JW, Gordon SH, Willett JL. Microwaveassisted rapid modification of zein by octenyl succinic anhydride. Cereal Chemistry. 2005;82:1–3.

33. Biswas A, Selling GW, Woods KK, Evans K. Surface modification of zein films. Ind Crops Prod. 2009;30:168–171.

34. Boerjan W, Ralph J, Baucher M. Lignin biosynthesis. Annu Rev Plant Biol. 2003;54:519–546.

35. Bosquez-Molina E, Guerrero-Legarreta I, Vernon-Carter EJ. Moisture barrier properties and morphology of mesquite gum–candelilla wax based edible emulsion coatings. Food Res Int. 2003;36:885–893.

36. Bosquez-Molina E, Tomás SA, Rodríguez-Huezo ME. Influence of CaCl2 on the water vapor permeability and the surface morphology of mesquite gum based edible films. LWT – Food Sci Technol. 2010;43:1419–1425.

37. Bravin B, Peressini D, Sensidoni A. Development and application of polysaccharide–lipid edible coating to extend shelf-life of dry bakery products. J Food Eng. 2006;76:280–290.

38. Buranov AU, Mazza G. Lignin in straw of herbaceous crops. Ind Crops Prod. 2008;28:237–259.

39. Campaniello D, Bevilacqua A, Sinigaglia M, Corbo MR. Chitosan: antimicrobial activity and potential applications for preserving minimally processed strawberries. Food Microbiol. 2008;25:992–1000.

40. Campo VL, Kawano DF, Silva Jr DB, Carvalho I. Carrageenans: biological properties, chemical modifications and structural analysis – a review. Carboh Polym. 2009;77:167–180.

41. Caner C. Whey protein isolate coating and concentration effects on egg shelf life. J Sci Food Agric. 2005;85:2143–2148.

42. Cao N, Fu Y, He J. Mechanical properties of gelatin films cross-linked, respectively, by ferulic acid and tannin acid. Food Hydrocoll. 2007;21:575–584.

43. Carneiro-da-Cunha MG, Cerqueira MA, Souza BWS, Souza MP, Teixeira JA, Vicente AA. Physical properties of edible coatings and films made with a polysaccharide from Anacardium occidentale L. J Food Eng. 2009;95:379–385.

44. Cerisuelo JP, Alonso J, Aucejo S, Gavara R, Hernández-Muñoz P. Modifications induced by the addition of a nanoclay in the functional and active properties of an EVOH film containing carvacrol for food packaging. J Membr Sci. 2012;423–424:247–256.

45. Cerqueira MA, Lima AM, Teixeira JA, Moreira RA, Vicente AA. Suitability of novel galactomannans as edible coatings for tropical fruits. J Food Eng. 2009;94:372–378.

46. Cerqueira MA, Sousa-Gallagher MJ, Macedo I, et al. Use of galactomannan edible coating application and storage temperature for prolonging shelf-life of “Regional” cheese. J Food Eng. 2010;97:87–94.

47. Cha DS, Chinnan MS. Biopolymer-based antimicrobial packaging – a review. Crit Rev Food Sci Nutr. 2004;44:223–237.

48. Chang PR, Jian R, Yu J, Ma X. Fabrication and characterisation of chitosan nanoparticles/plasticized-starch composites. Food Chem. 2010a;120:736–740.

49. Chang PR, Jian R, Yu J, Ma X. Starch-based composites reinforced with novel chitin nanoparticles. Carboh Polym. 2010b;80:420–425.

50. Chang ST, Chen LC, Lin SB, Chen HH. Nano-biomaterials application: morphology and physical properties of bacterial cellulose/gelatin composites via crosslinking. Food Hydrocoll. 2012;27:137–144.

51. Charreau H, Foresti ML, Vázquez A. Nanocellulose patents trends: a comprehensive review on patents on cellulose nanocrystals, microfibrillated and bacterial cellulose. Recent Pat Nanotechnol. 2013;7:56–80.

52. Chatterjee S, Adhya M, Guha AK, Chatterjee BP. Chitosan from Mucor rouxii: production and physico-chemical characterization. Proc Biochem. 2005;40:395–400.

53. Cheftel JC, Cuq JL, Lorient D. Amino acids, peptides, and proteins. In: Fennema OR, ed. Food Chemistry. New York: Marcel Dekker; 1985;245–369.

54. Chen H. Formation and properties of casein films and coatings. In: Gennadios A, ed. Protein-Based Films and Coatings. Boca Raton, FL: CRC Press; 2002;181–211.

55. Chen L, Tang C, Ning N, Wang C, Fu Q, Zhang Q. Preparation and properties of chitosan/lignin composite films. Chin J Polym Sci. 2009;27:739–746.

56. Chen WS, Yu HP, Liu YX. Preparation of millimeter-long cellulos I nanofibers with diameters of 30–80 nm from bamboo fibers. Carboh Polym. 2011;86:453–461.

57. Chen Y, Cao X, Chang PR, Huneault MA. Comparative study on the films of poly(vinyl alcohol)/pea starch nanocrystals and poly(vinyl alcohol)/native pea starch. Carbohy Polym. 2008;73:8–17.

58. Chidanandaiah, Keshri RC, Sanyal MK. Effect of sodium alginate coating with preservatives on the quality of meat patties during refrigerated (4 ± 1°C) storage. J Muscle Foods. 2009;20:275–292.

59. Chien PJ, Sheu F, Yang FH. Effects of edible chitosan coating on quality and shelf life of sliced mango fruit. J Food Eng. 2007;78:225–229.

60. Chiou BS, Avena-Bustillos RJ, Shey J, et al. Rheological and mechanical properties of cross-linked fish gelatins. Polym. 2006;47:6379–6386.

61. Chiumarelli M, Hubinger MD. Stability, solubility, mechanical and barrier properties of cassava starch – Carnauba wax edible coatings to preserve fresh-cut apples. Food Hydrocoll. 2012;28:59–67.

62. Coma V, Martial-Gros A, Garreau S, Copinet A, Salin F, Deschamps A. Edible antimicrobial films based on chitosan matrix. J Food Sci. 2002;67:1162–1169.

63. Conforti FD, Zinck JB. Hydrocolloid-lipid coating affect on weight loss, pectin content, and textural quality of green bell peppers. J Food Sci. 2002;67:1360–1363.

64. Cuq B, Gontard N, Guilbert S. Proteins as agricultural polymers for packaging production. Cereal Chem. 1998;75:1–9.

65. Curti E, Britto D, Campana-Filho SP. Methylation of chitosan with iodomethane: effect of reaction conditions on chemoselectivity and degree of substitution. Macromol Biosci. 2003;3:571–576.

66. Dalmas F, Cavaillé JY, Gauthier C, Chazeau L, Dendievel R. Viscoelastic behavior and electrical properties of flexible nanofiber filled polymer nanocomposites: influence of processing conditions. Compos Sci Technol. 2007;67:829–839.

67. Dang KTH, Singh Z, Swinny EE. Edible coatings influence fruit ripening, quality, and aroma biosynthesis in mango fruit. J Agric Food Chem. 2008;56:1361–1370.

68. Dangaran KL, Krochta JM. Preventing the loss of tensile, barrier and appearance properties caused by plasticizer crystallization in whey protein films. Int J Food Sci Technol. 2007;42:1094–1100.

69. Dangaran KL, Cooke P, Tomasula PM. The effect of protein particle size reduction on the physical properties of CO2-precipitated casein films. J Food Sci. 2006;71:E196–E201.

70. Dangaran K, Tomasula PM, Qi P. Structure and formation of proteinbased edible films and coatings. In: Embuscado ME, Huber KC, eds. Edible Films and Coatings for Food Applications. New York: Springer; 2009;25–56.

71. Dawlee S, Sugandhi A, Balakrishnan B, Labarre D, Jayakrishnan A. Oxidized chondroitin sulfate-cross-linked gelatin matrixes: a new class of hydrogels. Biomacromol. 2005;6:2040–2048.

72. de Campos A, Correa AC, Cannella D, et al. Obtaining nanofibers from curauá and sugarcane bagasse fibers using enzymatic hydrolysis followed by sonication. Cellulose. 2013;20:1491–1500.

73. De Moura MR, Aouada FA, Avena-Bustillos RJ, McHugh TH, Krochta JM, Mattoso LHC. Improved barrier and mechanical properties of novel hydroxypropyl methylcellulose edible films with chitosan/tripolyphosphate nanoparticles. J Food Eng. 2009;92:448–453.

74. De Paula RCM, Heatley F, Budd PM. Characterisation of Anacardium occidentale exudate polysaccharide. Polym Int. 1998;45:27–35.

75. Deepa B, Abraham E, Cherian BM, et al. Structure, morphology and thermal characteristics of banana nano fibers obtained by steam explosion. Bioresource Technol. 2011;102:1988–1997.

76. Del-Valle V, Hernández-Muñoz P, Guarda A, Galotto MJ. Development of a cactus-mucilage edible coating (Opuntia ficus indica) and its application to extend strawberry (Fragaria ananassa) shelf-life. Food Chem. 2005;91:751–756.

77. Dufresne A. Nanocellulose: From Nature to High Performance Tailored Materials Berlin: Walter de Gruyter; 2012.

78. Dutta PK, Tripathi S, Mehrotra GK, Dutta J. Perspectives for chitosan based antimicrobial films in food applications. Food Chem. 2009;114:1173–1182.

79. Ebringerová A, Heinze T. Xylan and xylan derivatives – biopolymers with valuable properties, 1: Naturally occurring xylans structures, isolation procedures and properties. Macromol Rapid Commun. 2000;21:542–556.

80. Ebringerová A, Hromadkova Z, Heinze T. Hemicellulose. Adv Polym Sci. 2005;186:1–67.

81. Elazzouzi-Hafraoui S, Nishiyama Y, Putaux JL, Heux L, Dubreuil F, Rochas C. The shape and size distribution of crystalline nanoparticles prepared by acid hydrolysis of native cellulose. Biomacromol. 2008;9:57–65.

82. Emmambux MN, Stading M, Taylor JRN. Sorghum kafirin film property modification with hydrolysable and condensed tannins. J Cereal Sci. 2004;40:127–135.

83. Fabra MJ, Talens P, Chiralt A. Tensile properties and water vapor permeability of sodium caseinate films containing oleic acid–beeswax mixtures. J Food Eng. 2008;85:393–400.

84. Fabra MJ, Hambleton A, Talens P, Debeaufort F, Chiralt A. Effect of ferulic acid and α-tocopherol antioxidants on properties of sodium caseinate edible films. Food Hydrocoll. 2011a;25:1441–1447.

85. Fabra MJ, Pérez-Masiá R, Talens P, Chiralt A. Influence of the homogenization conditions and lipid self-association on properties of sodium caseinate based films containing oleic and stearic acids. Food Hydrocoll. 2011b;25:1112–1121.

86. Fan Y, Xu Y, Wang D, Zhang L, Sun L, Zhang B. Effect of alginate coating combined with yeast antagonist on strawberry (Fragaria × ananassa) preservation quality. Postharv Biol Technol. 2009;53:84–90.

87. Farris S, Schaich KM, Liu LS, Piergiovanni L, Yam KL. Development of polyion-complex hydrogels as an alternative approach for the production of bio-based polymers for food packaging applications: a review. Trends Food Sci Technol. 2009;20:316–332.

88. Fayaz AM, Balaji K, Girilal M, Kalaichelvan PT, Venkatesan R. Mycobased synthesis of silver nanoparticles and their incorporation into sodium alginate films for vegetable and fruit preservation. J Agric Food Chem. 2009;57:6246–6252.

89. Fernandes SC, Santos DMPO, Vieira IC. Genipin-cross-linked chitosan as a support for laccase biosensor. Electroanal. 2013;25:557–566.

90. Fishman ML, Coffin DR, Konstance RP, Onwulata CI. Extrusion of pectin/starch blends plasticized with glycerol. Carboh Polym. 2000;41:317–325.

91. Forssell PM, Hulleman SHD, Myllarinen PJ, Moates GK, Parker R. Ageing of rubbery thermoplastic barley and oat starches. Carboh Polym. 1999;39:43–51.

92. García MA, Martino MN, Zaritky NE. Lipid addition to improve barrier properties of edible starch-based films and coatings. J Food Sci. 2000;65:941–947.

93. Gardner DJ, Oporto GS, Mills R, Azizi Samir MAS. Adhesion and surface issues in cellulose and nanocellulose. J Adhes Sci Technol. 2008;22:545–567.

94. Gennadios A, McHugh T, Weller C, Krochta JM. Edible coatings and films based on proteins. In: Krochta JM, Baldwin EA, Nisperos-Carriedo M, eds. Edible Coatings and Films to Improve Food Quality. Lancaster, PA: Technomic; 1994;231–247.

95. Ghanbarzadeh B, Oromiehi AR. Biodegradable biocomposite films based on whey protein and zein: barrier, mechanical properties and AFM analysis. Int J Biol Macromol. 2008;43:209–215.

96. Ghanbarzadeh B, Musavi M, Oromiehie AR, Rezayi K, Razmi E, Milani J. Effect of plasticizing sugars on water vapour permeability, surface energy and microstructure properties of zein films. LWT – Food Sci Technol. 2007;40:1191–1197.

97. Ghosh A, Ali MA, Dias GJ. Effect of cross-linking on microstructure and physical performance of casein protein. Biomacromol. 2009;10:1681–1688.

98. Gialamas H, Zinoviadou KG, Biliaderis CG, Koutsoumanis KP. Development of a novel bioactive packaging based on the incorporation of Lactobacillus sakei into sodium-caseinate films for controlling Listeria monocytogenes in foods. Food Res Int. 2010;43:2402–2408.

99. Ginestra G, Parker ML, Bennett RN, et al. Anatomical, chemical, and biochemical characterization of cladodes from prickly pear [Opuntia ficus-indica (L.) Mill.]. J Agric Food Chem. 2009;57:10323–10330.

100. Goddard JM, Hotchkiss JH. Polymer surface modification for the attachment of bioactive compounds. Prog Polym Sci. 2007;698–725.

101. Goesaert H, Brijs K, Veraverbeke WS, Courtin CM, Gebruers K, Delcour JA. Wheat flour constituents: how they impact bread quality, and how to impact their functionality. Trends Food Sci Technol. 2005;16:12–30.

102. Gómez-Guillén MC, Turnay J, Fernandez-Diaz MD, Ulmo N, Lizarbe MA, Montero P. Structural and physical properties of gelatin extracted from different marine especies: a comparative study. Food Hydrocoll. 2002;16:25–34.

103. Gómez-Guillén MC, Giménez B, López-Caballero ME, Montero MP. Functional and bioactive properties of collagen and gelatin from alternative sources: a review. Food Hydrocoll. 2011;25:1813–1827.

104. Gonçalves FP, Martins MC, Silva Jr GJ, Lourenço SA, Amorim L. Postharvest control of brown rot and Rhizopus rot in plums and nectarines using carnauba wax. Postharv Biol Technol. 2010;58:211–217.

105. Guhados G, Wan WK, Hutter JL. Measurement of the elastic modulus of single bacterial cellulose fibers using atomic force microscopy. Langmuir. 2005;21:6642–6646.

106. Guillard V, Issoupov V, Redl A, Gontard N. Food preservative content reduction by controlling sorbic acid release from a superficial coating. Innov Food Sci Emerg Technol. 2009;10:108–115.

107. Han JH. Antimicrobial food packaging. In: Ahvenainen R, ed. Novel Food Packaging Techniques. Cambridge: Woodhead Publishing Limited; 2003;50–70.

108. Han JH, Gennadios A. Edible films and coatings: a review. In: Han JH, ed. Innovations in Food Packaging. London: Elsevier; 2005;239–262.

109. Hanani ZAN, Beatty E, Roos YH, Morris MA, Kerry JP. Manufacture and characterization of gelatin films derived from beef, pork and fish sources using twin screw extrusion. J Food Eng. 2012a;113:606–614.

110. Hanani ZAN, Roos YH, Kerry JP. Use of beef, pork and fish gelatin sources in the manufacture of films and assessment of their composition and mechanical properties. Food Hydrocoll. 2012b;29:144–151.

111. Hansen NML, Plackett D. Sustainable films and coatings from hemicelluloses: a review. Biomacromol. 2008;9:1493–1505.

112. Haug IJ, Draget KI, Smidsrod O. Physical and rheological properties of fish gelatin compared to mammalian gelatin. Food Hydrocoll. 2004;18:203–213.

113. He L, Mu C, Shi J, Zhang Q, Shi B, Lin W. Modification of collagen with a natural cross-linker, procyanidin. Int J Biol Macromol. 2011;48:354–359.

114. Helander IM, Nurmiaho-Lassila EL, Ahvenainen R, Rhoades J, Roller S. Chitosan disrupts the barrier properties of the outer membrane of Gramnegative bacteria. Int J Food Microbiol. 2001;71:235–244.

115. Helbert W, Cavaillé CY, Dufresne A. Thermoplastic nanocomposites filled with wheat straw cellulose whiskers Part I: processing and mechanical behaviour. Polym Compos. 1996;17:604–611.

116. Kaiser Consultancy Helmut. Bioplastics market worldwide 2010/11–2015-2020–2025. available at http://www.hkc22.com/bioplastics.html; 2012.

117. Hernandez-Izquierdo VM, Krochta JM. Thermoplastic processing of proteins for film formation – a review. J Food Sci. 2008;73:30–39.

118. Hernandez-Izquierdo VM, Reid DS, McHugh TH, Berrios JJ, Krochta JM. Thermal transitions and extrusion of glycerol-plasticized protein mixtures. J Food Sci. 2008;73:E169–E175.

119. Hernández-Muñoz P, Kanavouras A, Ng PKW, Gavara R. Development and characterization of biodegradable films made from wheat gluten protein fractions. J Agric Food Chem. 2003;51:7647–7654.

120. Hernández-Muñoz P, Almenar E, Ocio MJ, Gavara R. Effect of calcium dips and chitosan coatings on postharvest life of strawberries (Fragaria ananassa). Postharvest Biol Technol. 2006;39:247–253.

121. Hochstetter A, Talja RA, Helen HJ, Hyvonen L, Jouppila K. Properties of gluten-based sheet produced by twin-screw extruder. LWT – Food Sci Technol. 2006;39:893–901.

122. Höije A, Grondahl M, Tommeraas K, Gatenholm P. Isolation and characterization of physicochemical and material properties of arabinoxylans from barley husks. Carbohy Polym. 2005;61:266–275.

123. Huang M, Yu J, Ma X. Ethanolamine as a novel plasticiser for thermoplastic starch. Polym Degrad Stab. 2005;90:501–507.

124. Hubbe MA, Rojas OJ, Lucia LA, Sain M. Cellulosic nanocomposites: a review. Bioresources. 2008;3:929–980.

125. Huneault MA, Li H. Morphology and properties of compatibilized polylactide/thermoplastic starch blends. Polym. 2007;48:270–280.

126. Iguchi M, Yamanaka S, Budhiono A. Bacterial cellulose – a masterpiece of nature’s arts. J Mater Sci. 2000;35:261–270.

127. Ikeda A, Takemura A, Ono H. Preparation of low-molecular weight alginic acid by acid hydrolysis. Carbohy Polym. 2000;42:421–425.

128. Isogai A, Saito T, Fukuzumi H. TEMPO-oxidized cellulose nanofibers. Nanoscale. 2011;3:71–85.

129. Iwamoto S, Nakagaito AN, Yano H. Nano-fibrillation of pulp fibers for the processing of transparent nanocomposites. Appl Phys A. 2007;89:461–466.

130. Janjarasskul T, Krochta JM. Edible packaging materials. Annu Rev Food Sci Technol. 2010;1:415–448.

131. Jayasekara R, Harding I, Bowater I, Lornergan G. Biodegradability of selected range of polymers and polymer blends and standard methods for assessment of biodegradation. J Polym Environ. 2005;13:231–251.

132. Jerez A, Partal P, Martinez I, Gallegos C, Guerrero A. Protein-based bioplastics: effect of thermo-mechanical processing. Rheol Acta. 2007;46:711–720.

133. Jiang T, Feng L, Zheng X, Li L. Physicochemical responses and microbial characteristics of shiitake mushroom (Lentinus edodes) to gum arabic coating enriched with natamycin during storage. Food Chem. 2013;138:1992–1997.

134. Jin J, Song M, Hourston DJ. Novel chitosan-based films cross-linked by genipin with improved physical properties. Biomacromol. 2004;5:162–168.

135. Johnson DT, Taconi KA. The glycerin glut: options for the value-added conversion of crude glycerol resulting from biodiesel production. Environ Progr. 2007;26:338–348.

136. Jongjareonrak A, Benjakul S, Visessanguan W, Prodpran T, Tanaka M. Characterization of edible films from skin gelatin of brownstripe red snapper and bigeye snapper. Food Hydrocoll. 2006;20:492–501.

137. Jordan J, Jacob KI, Tannenbaum R, Sharaf MA, Jasiuk I. Experimental trends in polymer nanocomposites: a review. Mat Sci Eng A. 2005;393:1–11.

138. Jouki M, Khazaei N, Ghasemlou M, HadiNezhad M. Effect of glycerol concentration on edible film production from cress seed carbohydrate gum. Carboh Polym. 2013;96:39–46.

139. Kang HJ, Jo C, Lee NY, Kwon JH, Byun MW. A combination of gamma irradiation and CaCl2 immersion for a pectin-based biodegradable film. Carboh Polym. 2005;60:547–551.

140. Karbowiak T, Gougeon RD, Rigolet S, Delmotte L, Debeaufort F, Voilley A. Diffusion of small molecules in edible films: effect of water and interactions between diffusant and biopolymers. Food Chem. 2008;106:1340–1349.

141. Karbowiak T, Debeaufort F, Voilley A, Trystam G. From macroscopic to molecular scale investigations of mass transfer of small molecules through edible packaging applied at interfaces of multiphase food products. Innov Food Sci Emerg Technol. 2009;10:116–127.

142. Karim AA, Bhat R. Fish gelatin: properties, challenges, and prospects as an alternative to mammalian gelatins. Food Hydrocoll. 2009;23:563–576.

143. Kester JJ, Fennema OR. Edible films and coatings: a review. Food Technol. 1986;40:47–59.

144. Kim Y, Jung R, Kim HS, Jin HJ. Transparent nanocomposites prepared by incorporating microbial nanofibrils into poly(L-lactic acid). Curr Appl Phys. 2009;9:S69–S71.

145. Klemm D, Heublein B, Fink HP, Bohn A. Cellulose: fascinating biopolymer and sustainable raw material. Angew Chem Int Ed. 2005;44:3358–3393.

146. Kokini JL, Cocero AM, Madeka H, de Graaf E. The development of state diagrams for cereal proteins. Trends Food Sci Technol. 1994;5:281–288.

147. Konwarh R, Karak N, Misra M. Electrospun cellulose acetate nanofibers: the present status and gamut of biotechnological applications. Biotechnol Adv. 2013;31:421–437.

148. Kramer ME. Structure and function of starch-based edible films and coatings. In: Embuscado ME, Huber KC, eds. Edible Films and Coatings for Food Applications. New York: Springer; 2009;113–134.

149. Krishna M, Nindo CI, Min SC. Development of fish gelatin edible films using extrusion and compression molding. J Food Eng. 2012;108:337–344.

150. Kristo E, Biliaderis CG. Physical properites of starch nanocrystalreinforced pullulan films. Carboh Polym. 2007;68:146–158.

151. Krochta JM. Proteins as raw materials for films and coatings: definitions, current status, and opportunities. In: Gennadios A, ed. Protein-Based Films and Coatings. Boca Raton, FL: CRC; 2002;1–41.

152. Krochta JM, De Mulder-Johnston C. Edible and biodegradable polymer films: challenges and opportunities. Food Technol. 1997;51:61–73.

153. Kugel A, Stafslien S, Chisholm BJ. Antimicrobial coatings produced by “tethering” biocides to the coating matrix: a comprehensive review. Prog Org Coat. 2011;72:222–252.

154. Kumar P, Sandeep KP, Alavi S, Truong VD, Gorga RE. Preparation and characterization of bio-nanocomposite films based on soy protein isolate and montmorillonite using melt extrusion. J Food Eng. 2010;100:480–489.

155. Kvien I, Oksman K. Orientation of cellulose nanowhiskers in polyvinyl alcohol. Appl Phys A. 2007;87:641–643.

156. Lacroix M, Cooksey K. Edible films and coatings from animal-origin proteins coatings. In: Han JH, ed. Innovations in Food Packaging. London: Elsevier; 2005;301–317.

157. Lagrain B, Goderis B, Brijs K, Delcour JA. Molecular basis of processing wheat gluten toward biobased materials. Biomacromol. 2010;11:533–541.

158. Le Corre D, Bras J, Dufresne A. Starch nanoparticles: a review. Biomacromol. 2010;11:1139–1153.

159. Li HL, Wu B, Mu CD, Lin W. Concomitant degradation in periodate oxidation of carboxymethyl cellulose. Carbohy Polym. 2011a;84:881–886.

160. Li M, Liu P, Zou W, et al. Extrusion processing and characterization of edible starch films with different amylose contents. J Food Eng. 2011b;106:95–101.

161. Li R, Hu P, Ren X, Worley SD, Huang TS. Antimicrobial N-halamine modified chitosan films. Carboh Polym. 2013;92:534–539.

162. Lian ZX, Ma ZS, Wei J, Liu H. Preparation and characterization of immobilized lysozyme and evaluation of its application in edible coatings. Proc Biochem. 2012;47:201–208.

163. Limpisophon K, Tanaka M, Osako K. Characterization of gelatin–fatty acid emulsion films based on blue shark (Prionace glauca) skin gelatin. Food Chem. 2010;122:1095–1101.

164. Lin B, Du Y, Liang X, Wang X, Wang X. Effect of chitosan coating on respiratory behavior and quality of stored litchi under ambient temperature. J Food Eng. 2011;102:94–99.

165. López de Lacey AM, López-Caballero ME, Gómez-Estaca J, Gómez-Guillén MC, Montero P. Functionality of Lactobacillus acidophilus and Bifidobacterium bifidum incorporated to edible coatings and films. Innov Food Sci Emerg Technol. 2012;16:277–282.

166. López-León T, Carvalho ELS, Seijo B, Ortega-Vinuesa JL, Bastos-González D. Physicochemical characterization of chitosan nanoparticles: electrokinetic and stability behavior. J Colloid Interf Sci. 2005;283:344–351.

167. Lopez-Rubio A, Gavara R, Lagaron JM. Bioactive packaging: turning foods into healthier foods through biomaterials. Trends Food Sci Technol. 2006;17:567–575.

168. Lu Y, Weng L, Zhang L. Morphology and properties of soy protein isolate thermoplastics reinforced with chitin whiskers. Biomacromol. 2004;5:1046–1051.

169. Ma XF, Yu JG, Wan JJ. Urea and ethanolamine as a mixed plasticizer for thermoplastic starch. Carboh Polym. 2006;64:267–273.

170. Ma X, Jian R, Chang PR, Yu J. Fabrication and characterization of citric acid-modified starch nanoparticles/plasticized-starch composites. Biomacromol. 2008;9:3314–3320.

171. Maftoonazad N, Ramaswamy HS. Postharvest shelf-life extension of avocados using methyl cellulose-based coating. LWT – Food Sci Technol. 2005;38:617–624.

172. Maftoonazad N, Ramaswamy HS, Moalemiyan M, Kushalappa AC. Effect of pectin-based edible emulsion coating on changes in quality of avocado exposed to Lasiodiplodia theobromae infection. Carboh Polym. 2007;68:341–349.

173. Maftoonazad N, Ramaswamy HS, Marcotte M. Shelf-life extension of peaches through sodium alginate and methyl cellulose edible coatings. Int J Food Sci Technol. 2008;43:951–957.

174. Mali S, Grossmann MVE, García MA, Martino MM, Zaritzky NE. Barrier, mechanical and optical properties of plasticized yam starch films. Carboh Polym. 2004;56:129–135.

175. Mali S, Sakanaka LS, Yamashita F, Grossmann MVE. Water sorption and mechanical properties of cassava starch films and their relations to plasticizing effect. Carboh Polym. 2005;60:283–289.

176. Mangavel C, Rossignol N, Perronnet A, Barbot J, Popineau Y, Gueguen J. Properties and microstructure of thermopressed wheat gluten films: a comparison with cast films. Biomacromol. 2004;5:1596–1601.

177. Marcos B, Aymerich T, Monfort JM, Garriga M. High-pressure processing and antimicrobial biodegradable packaging to control Listeria monocytogenes during storage of cooked ham. Food Microbiol. 2008;25:177–182.

178. Martins JT, Cerqueira MA, Souza BWS, Avides MC, Vicente AA. Shelf life extension of ricotta cheese using coatings of galactomannans from nonconventional sources incorporating nisin against Listeria monocytogenes. J Agric Food Chem. 2010;58:1884–1891.

179. Martins JT, Cerqueira MA, Vicente AA. Influence of α-tocopherol on physicochemical properties of chitosan-based films. Food Hydrocoll. 2012;27:220–227.

180. Martucci JF, Ruseckaite RA. Tensile properties, barrier properties, and biodegradation in soil of compression-molded gelatin-dialdehyde starch films. J Appl Polym Sci. 2009;112:2166–2178.

181. Mathew S, Abraham TE. Characterisation of ferulic acid incorporated starch-chitosan blend films. Food Hydrocoll. 2008;22:826–835.

182. McGarvie D, Parolis H. The mucilage of Opuntia ficus-indica. Carboh Res. 1979;69:171–179.

183. McGuire RG, Baldwin EA. Acidic fruit coatings for maintenance of color and decay prevention on lychees postharvest. Pr Fl St Hortic Soc. 1998;111:243–247.

184. McHugh TH, Aujard JF, Krochta JM. Plasticized whey protein edible films: water vapor permeability properties. J Food Sci. 1994;59:416–419.

185. Mehyar GF, Al-Ismail K, Han JH, Chee GW. Characterization of edible coatings consisting of pea starch, whey protein isolate, and carnauba wax and their effects on oil rancidity and sensory properties of walnuts and pine nuts. J Food Sci. 2012;77:E52–E59.

186. Mi FL, Shyu SS, Peng CK. Characterization of ring-opening polymerization of genipin and pH-dependent cross-linking reactions between chitosan and genipin. J Polym Sci A Polym Chem. 2005;43:1985–2000.

187. Mi FL, Huang CT, Liang HF, et al. Physicochemical, antimicrobial, and cytotoxic characteristics of a chitosan film cross-linked by a naturally occurring cross-linking agent, aglycone geniposidic acid. J Agric Food Chem. 2006;54:3290–3296.

188. Mikkonen KS, Tenkanen M. Sustainable food-packaging materials based on future biorefinery products: xylans and mannans. Trends Food Sci Technol. 2012;28:90–102.

189. Mikkonen KS, Rita H, Helén H, Talja RA, Hyvonen L, Tenkanen M. Effect of polysaccharide structure on mechanical and thermal properties of galactomannan-based films. Biomacromol. 2007;8:3198–3205.

190. Mild RM, Joens LA, Friedman M, et al. Antimicrobial edible apple films inactivate antibiotic resistant and susceptible Campylobacter jejuni strains on chicken breast. J Food Sci. 2011;76:M163–M168.

191. Min S, Krochta JM. Ascorbic acid-containing whey protein film coatings for control of oxidation. J Agric Food Chem. 2007;55:2964–2969.

192. Miranda RL. Cashew tree bark secretion – perspectives for its use in protein isolation strategies. Open Glycosci. 2009;2:16–19.

193. Mishra SB, Mishra AK, Kaushik NK, Khan MA. Study of performance properties of lignin-based polyblends with polyvinyl chloride. J Mat Proc Technol. 2007;183:273–276.

194. Moates GK, Noel TR, Parker R, Ring SG. Dynamic mechanical and dielectric characterisation of amylose–glycerol films. Carboh Polym. 2001;44:247–253.

195. Mohareb E, Mittal GS. Formulation and process conditions for biodegradable/edible soy-based packaging trays. Packag Technol Sci. 2007;20:1–15.

196. Moon RJ, Martini A, Nairn J, Simonsen J, Youngblood J. Cellulose nanomaterials review: structure, properties and nanocomposites. Chem Soc Rev. 2011;40:3941–3994.

197. Morillon V, Debeaufort F, Blond G, Capelle M, Voilley A. Factors affecting the moisture permeability of lipid-based edible films: a review. Crit Rev Food Sci Nutr. 2002;42:67–89.

198. Mu C, Guo J, Li X, Lin W, Li D. Preparation and properties of dialdehyde carboxymethyl cellulose crosslinked gelatin edible films. Food Hydrocoll. 2012;27:22–29.

199. Mueller RJ. Biological degradation of synthetic polyesters – enzymes as potential catalysts for polyester recycling. Proc Biochem. 2006;41:2124–2128.

200. Müller-Buschbaum P, Gebhardt R, Maurer E, Bauer E, Gehrke R, Doster W. Thin casein films as prepared by spin-coating: influence of film thickness and of pH. Biomacromol. 2006;7:1773–1780.

201. Muzzarelli RAA. Genipin-crosslinked chitosan hydrogels as biomedical and pharmaceutical aids. Carboh Polym. 2009;77:1–9.

202. Nair KG, Dufresne A. Crab shell chitin whisker reinforced natural rubber nanocomposites 1 Processing and swelling behavior. Biomacromol. 2003;4:657–665.

203. Nakagaito AN, Yano H. Novel high-strength biocomposites based on microfibrillated cellulose having nano-order-unit web-like network structure. Appl Phys A Mater Sci Process. 2005;80:155–159.

204. Navarro-Tarazaga ML, Massa A, Pérez-Gago MB. Effect of beeswax content on hydroxypropyl methylcellulose-based edible film properties and postharvest quality of coated plums (Cv Angeleno). LWT – Food Sci Technol. 2011;44:2328–2334.

205. Ng A, Greenshields RN, Waldron KW. Oxidative cross-linking of corn bran hemicellulose: formation of ferulic acid dehydrodimers. Carboh Res. 1997;303:459–462.

206. Núñez-Flores R, Giménez B, Fernández-Martín F, López-Caballero ME, Montero MP, Gómez-Guillén MC. Physical and functional characterization of active fish gelatin films incorporated with lignin. Food Hydrocoll. 2013;30:163–172.

207. Ojagh SM, Núñez-Flores R, López-Caballero ME, Montero MP, Gómez-Guillén MC. Lessening of high-pressure-induced changes in Atlantic salmon muscle by the combined use of a fish gelatin–lignin film. Food Chem. 2011;125:595–606.

208. Olivas GI, Barbosa-Cánovas G. Alginate-calcium films: water vapor permeability and mechanical properties as affected by plasticizer and relative humidity. LWT – Food Sci Technol. 2008;41:359–366.

209. O’Neill MA, York WS. The composition and structure of plant primary cell walls. In: Rose JKC, ed. Boca Raton, FL: CRC Press; 2003;1–54. The Plant Cell Wall. 8 Annual Plant Reviews.

210. Osés J, Fabregat-Vázquez M, Pedroza-Islas R, Tomás SA, Cruz-Orea A, Maté JI. Development and characterization of composite edible films based on whey protein isolate and mesquite gum. J Food Eng. 2009;92:56–62.

211. Ou S, Wang Y, Tang S, Huang C, Jackson MG. Role of ferulic acid in preparing edible films from soy protein isolate. J Food Eng. 2005;70:205–210.

212. Paes SS, Yakimets I, Wellner N, Hill SE, Wilson RH, Mitchell JR. Fracture mechanisms in biopolymer films using coupling of mechanical analysis and high speed visualization technique. Eur Polymer J. 2010;46:2300–2309.

213. Paillet M, Dufresne A. Chitin whisker reinforced thermoplastic nanocomposites. Macromol. 2001;34:6527–6530.

214. Park HJ, Chinnan MS. Gas and water vapor barrier properties of edible films from protein and cellulosic materials. J Food Eng. 1995;25:497–507.

215. Park SI, Zhao Y. Development and characterization of edible films from cranberry pomace extracts. J Food Sci. 2006;71:E95–E101.

216. Park Y, Doherty WOS, Halley PJ. Developing lignin-based resin coatings and composites. Ind Crops Prod. 2008;27:163–167.

217. Parker ML, Ng A, Waldron KW. The phenolic acid and polysaccharide composition of cell walls of bran layers of mature wheat (Triticum aestivum L cv Avalon) grains. J Sci Food Agric. 2005;85:2539–2547.

218. Paula EL, Mano V, Pereira VJ. Influence of cellulose nanowhiskers on the hydrolytic degradation behavior of poly(d, l-lactide). Polym Degrad Stab. 2011;96:1631–1638.

219. Peressini D, Bravin B, Lapasin R, Rizzotti C, Sensidoni A. Starch–methylcellulose based edible films: rheological properties of film-forming dispersions. J Food Eng. 2003;59:25–32.

220. Pérez-Gago MB, Krochta JM. Formation and properties of whey protein films and coatings. In: Gennadios A, ed. Protein-Based Films and Coatings. Boca Raton, FL: CRC Press; 2002;159–180.

221. Pérez-Gago MB, Serra M, Alonso M, Mateos M, Del Río MA. Effect of solid content and lipid content of whey protein isolate-beeswax edible coatings on color change of fresh-cut apples. J Food Sci. 2003;68:2186–2191.

222. Petersson L, Oksman K. Preparation and properties of biopolymer based nanocomposite films using microcrystalline cellulose. In: Oksman K, Sain M, eds. Cellulose Nanocomposites, Processing, Characterization and Properties. Oxford: Oxford University Press; 2006;132–150. ACS Symposium Series 938.

223. Petersson L, Kvien I, Oksman K. Structure and thermal properties of poly(lactic acid)/cellulose whiskers nanocomposite materials. Compos Sci Technol. 2007;67:2535–2544.

224. Piermaria J, Bosch A, Pinotti A, Yantorno O, Garcia MA, Abraham AG. Kefiran films plasticized with sugars and polyols: water vapor barrier and mechanical properties in relation to their microstructure analyzed by ATR/FT-IR spectroscopy. Food Hydrocoll. 2011;25:1261–1269.

225. Podsiadlo P, Choi SY, Shim B, Lee J, Cuddihy M, Kotov NA. Molecularly engineered nanocomposites: layer-by-layer assembly of cellulose nanocrystals. Biomacromol. 2005;6:2914–2918.

226. Pushpadass HA, Marx DB, Hanna MA. Effects of extrusion temperature and plasticizers on the physical and functional properties of starch films. Starch. 2008;60:527–538.

227. Putra A, Kakugo A, Furukawa H, Gong JP, Osada Y. Tubular bacterial cellulose gel with oriented fibrils on the curved surface. Polym. 2008;49:1885–1891.

228. Qiao R, Brinson LC. Simulation of interphase percolation and gradients in polymer nanocomposites. Compos Sci Technol. 2009;69:491–499.

229. Qiao X, Tang Z, Sun K. Plasticization of corn starch by polyol mixtures. Carboh Polym. 2011;83:659–664.

230. Qiu X, Hu S. “Smart” materials based on cellulose: a review of the preparations, properties, and applications. Materials. 2013;6:738–781.

231. Ramos LP. The chemistry involved in the stream treatment of lignocellulosic materials. Química Nova. 2003;26:863–871.

232. Ramos OL, Reinas I, Silva SI, et al. Effect of whey protein purity and glycerol content upon physical properties of edible films manufactured therefrom. Food Hydrocoll. 2013;30:110–122.

233. Redl A, Morel MH, Bonicel J, Vergnes B, Guilbert S. Extrusion of wheat gluten plasticized with glycerol: influence of process conditions on flow behavior, rheological properties, and molecular size distribution. Cereal Chem. 1999;76:361–370.

234. Reinoso E, Mittal GS, Lim LT. Influence of whey protein composite coatings on plum (Prunus Domestica L.) fruit quality. Food Bioprocess Technol. 2008;1:314–325.

235. Retegi A, Gabilondo N, Peña C, et al. Bacterial cellulose films with controlled microstructure–mechanical property relationships. Cellulose. 2010;17:661–669.

236. Rhim JW, Ng PKW. Natural biopolymer-based nanocomposite films for packaging applications. Crit Rev Food Sci Nutr. 2007;47:411–433.

237. Rhim JW, Shellhammer TH. Lipid-based edible films and coatings. In: Han JH, ed. Innovations in Food Packaging. London: Elsevier; 2005;362–383.

238. Ribeiro C, Vicente AA, Teixeira JA, Miranda C. Optimization of edible coating composition to retard strawberry fruit senescence. Postharv Biol Technol. 2007;44:63–70.

239. Rosa MF, Medeiros ES, Malmonge JA, et al. Cellulose nanowhiskers from coconut husk fibers: effect of preparation conditions on their thermal and morphological behavior. Carboh Polym. 2010;81:83–92.

240. Rose JKC, Bennett AB. Cooperative disassembly of the cellulosexyloglucan network of plant cell walls: parallels between cell expansion and fruit ripening. Trends Plant Sci. 1999;4:176–183.

241. Rossman JM. Commercial manufacture of edible films. In: Embuscado ME, Huber KC, eds. Edible Films and Coatings for Food Applications. Dordrecht: Springer; 2009;367–390.

242. Sanchez-Garcia MD, Gimenez E, Lagaron JM. Morphology and barrier properties of solvent cast composites of thermoplastic biopolymers and purified cellulose fibers. Carboh Polym. 2008;71:235–244.

243. Sanjurjo K, Flores S, Gerschenson L, Jagus R. Study of the performance of nisin supported in edible films. Food Res Int. 2006;39:749–754.

244. Sarabia AI, Gómez-Guillén MC, Montero P. The effect of added salt on the viscoelastic properties of fish skin gelatin. Food Chem. 2000;70:71–76.

245. Saucedo-Pompa S, Rojas-Molina R, Aguilera-Carbó AF, et al. Edible film based on candelilla wax to improve the shelf life and quality of avocado. Food Res Int. 2009;42:511–515.

246. Scheller HV, Ulvskov P. Hemicelluloses. Annu Rev Plant Biol. 2010;61:263–289.

247. Schou M, Longares A, Montesinos-Herrero C, Monahan FJ, O’Riordan D, O’Sullivan M. Properties of edible sodium caseinate films and their application as food wrapping. LWT – Food Sci Technol. 2005;38:605–610.

248. Shen XL, Wu JM, Chen Y, Zhao G. Antimicrobial and physical properties of sweet potato starch films incorporated with potassium sorbate or chitosan. Food Hydrocoll. 2010;24:285–290.

249. Shi K, Kokini JL, Huang Q. Engineering zein films with controlled surface morphology and hydrophilicity. J Agric Food Chem. 2009;57:2186–2192.

250. Shi K, Huang Y, Yu H, Lee TC, Huang Q. Reducing the brittleness of zein films through chemical modification. J Agric Food Chem. 2011;59:56–61.

251. Shon J, Chin KB. Effect of whey protein coating on quality attributes of low-fat, aerobically packaged sausage during refrigerated storage. J Food Sci. 2008;73:C469–C475.

252. Sims IM, Furneaux RH. Structure of the exudates gum from Meryta sinclairii. Carboh Polym. 2003;52:423–431.

253. Siracusa V, Rocculi P, Romani S, Dalla Rosa M. Biodegradable polymers for food packaging: a review. Trends Food Sci Technol. 2008;19:634–643.

254. Siró I, Plackett D. Microfibrillated cellulose and new nanocomposite materials: a review. Cellulose. 2010;17:459–494.

255. Sonkaew P, Sane A, Suppakul P. Antioxidant activities of curcumin and ascorbyl dipalmitate nanoparticles and their activities after incorporation into cellulose-based packaging films. J Agric Food Chem. 2012;60:5388–5399.

256. Sothornvit R, Krochta JM. Plasticizers in edible films and coatings. In: Han JH, ed. Innovations in Food Packaging. London: Elsevier; 2005;403–463.

257. Sothornvit R, Pitak N. Oxygen permeability and mechanical properties of banana films. Food Res Int. 2007;40:365–370.

258. Sothornvit R, Rodsamran P. Effect of a mango film on quality of whole and minimally processed mangoes. Postharv Biol Technol. 2008;47:407–415.

259. Sriamornsak P. Chemistry of pectin and its pharmaceutical uses: a review. Silpakorn Univ Int J. 2003;3:206–228.

260. Srinivasa PC, Ramesh MN, Tharanathan RN. Effect of plasticizers and fatty acids on mechanical and permeability characteristics of chitosan films. Food Hydrocoll. 2007;21:1113–1122.

261. Sriupayo J, Supaphol P, Blackwell J, Rujiravanit R. Preparation and characterization of α-chitin whisker-reinforced chitosan nanocomposite films with or without heat treatment. Carboh Polym. 2005;62:130–136.

262. Srivastava M, Kapoor VP. Seed galactomannans: an overview. Chem Biodiv. 2005;2:295–317.

263. Stepto RFT. Thermoplastic starch. Macromol Symp. 2009;279:163–168.

264. Strauss G, Gibson SM. Plant phenolics as cross-linkers of gelatin gels and gelatin-based coacervates for use as food ingredients. Food Hydrocoll. 2004;18:81–89.

265. Sun RC, Sun XF, Tomkinson J. Hemicelluloses and their derivatives. In: Gatenholm P, Tenkanen M, eds. Hemicellulose: Science and Technology. Washington, DC: American Chemical Society; 2004;2–22.

266. Sung HW, Huang RN, Huang LLH, Tsai CC. In vitro evaluation of cytotoxicity of a naturally occurring cross-linking reagent for biological tissue fixation. J Biomat Sci Polym Ed. 1999;10:63–78.

267. Suppakul P, Jutakorn K, Bangchokedee Y. Efficacy of cellulose-based coating on enhancing the shelf life of fresh eggs. J Food Eng. 2010;98:207–213.

268. Talens P, Krochta JM. Plasticizing effects of beeswax and carnauba wax on tensile and water vapor permeability properties of whey protein films. J Food Sci. 2005;70:E239–E243.

269. Tang C, Liu H. Cellulose nanofiber reinforced poly(vinyl alcohol) composite film with high visible light transmittance. Compos A Appl Sci Manuf. 2008;39:1638–1643.

270. Tapia MS, Rojas-Graü MA, Rodríguez FJ, Ramírez J, Carmona A, MartinBelloso O. Alginate- and gellan-based edible films for probiotic coatings on fresh-cut fruits. J Food Sci. 2007;72:E190–E196.

271. Tapia-Blácido DR, Sobral PJA, Menegalli FC. Effect of drying conditions and plasticizer type on some physical and mechanical properties of amaranth flour films. LWT – Food Sci Technol. 2013;50:392–400.

272. Tharanathan RN. Biodegradable films and composite coatings: past, present and future. Trends Food Sci Technol. 2003;14:71–78.

273. Thunwall M, Kuthanová V, Boldizar A, Rigdahl M. Film blowing of thermoplastic starch. Carboh Polym. 2008;71:583–590.

274. Ugartondo V, Mitjans M, Vinardell MP. Applicability of lignins from different sources as antioxidants based on the protective effects on lipid peroxidation induced by oxygen radicals. Ind Crops Prod. 2009;30:184–187.

275. Vachon C, Yu HL, Yefsah R, Alain R, St-Gelais D, Lacroix M. Mechanical and structural properties of milk protein edible films cross-linked by heating and gamma irradiation. J Agric Food Chem. 2000;48:3202–3209.

276. Van de Velde F, Knutsen SH, Usov AL, Rollema HS, Cerezo AS. 1H and 13C high resolution NMR spectroscopy of carrageenans: application in research and industry. Trends Food Sci Technol. 2002;13:73–92.

277. VanderZee M. Biodegradability of polymers – mechanisms and evaluation methods. In: Bastioli C, ed. Handbook of Biodegradable Polymers. Shropshire: Rapra; 2005;1–31.

278. Vengal JC, Srikumar M. Processing and study of novel lignin-starch and lignin-gelatin biodegradable polymeric films. Trends Biomater Artif Organs. 2005;18:238–241.

279. Verbeek CJR, van den Berg LE. Extrusion processing and properties of protein-based thermoplastics. Macromol Mater Eng. 2010;295:10–21.

280. Vieira MGA, Silva MA, Santos LO, Beppu MM. Natural-based plasticizers and biopolymer films: a review. Eur Polym J. 2011;47:254–263.

281. Wang X, Sun X, Liu H, Li M, Ma Z. Barrier and mechanical properties of carrot puree films. Food Bioprod Process. 2011;89:149–156.

282. Wardy W, Torrico DD, Jirangrat W, No HK, Saalia FK, Prinyawiwatkul W. Chitosan-soybean oil emulsion coating affects physico-functional and sensory quality of eggs during storage. LWT – Food Sci Technol. 2011;44:2349–2355.

283. Wasswa J, Tang J, Gu X. Utilization of fish processing by-products in the gelatin industry. Food Rev Int. 2007;23:159–174.

284. Wu Q, Henriksson M, Liu X, Berglund LA. A high strength nanocomposite based on microcrystalline cellulose and polyurethane. Biomacromol. 2007;8:3687–3692.

285. Wyman CE, Decker SR, Himmel ME, Brady JW, Skopec CE, Viikari L. Hydrolysis of cellulose and hemicellulose. In: Dumitriu S, ed. Polysaccharides: Structural Diversity and Functional Versatility. New York: Marcel Dekker; 2005;995–1033.

286. Yakimets I, Paes SS, Wellner N, Smith AC, Wilson RH, Mitchell JR. Effect of water content on the structural reorganization and elastic properties of biopolymer films: a comparative study. Biomacromol. 2007;8:1710–1722.

287. Yan Z, Chen S, Wang H, Wang B, Wang C, Jiang J. Cellulose synthesized by Acetobacter xylinum in the presence of multi-walled carbon nanotubes. Carboh Res. 2008;343:73–80.

288. Yang F, Hanna MA, Sun R. Value-added uses for crude glycerol – a byproduct of biodiesel production. Biotechn Biofuels. 2012;5:13.

289. Yang Q, Fukuzumi H, Saito T, Isogai A, Zhang L. Transparent cellulose films with high gas barrier properties fabricated from aqueous alkali/urea solutions. Biomacromol. 2011;12:2766–2771.

290. Ying R, Barron C, Saulnier L, Rondeau-Mouro C. Water mobility within arabinoxylan and β-glucan films studied by NMR and dynamic vapour sorption. J Sci Food Agric. 2011;91:2601–2605.

291. Yousefi H, Nishino T, Faezipour M, Ebrahimi G, Shakeri A. Direct fabrication of all-cellulose nanocomposite from cellulose microfibers using ionic liquid-based nanowelding. Biomacromol. 2011;12:4080–4085.

292. Yousefi H, Faezipour M, Hedjazi S, Mousavi MM, Azusa Y, Heidari AH. Comparative study of paper and nanopaper properties prepared from bacterial cellulose nanofibers and fibers/ground cellulose nanofibers of canola straw. Ind Crop Prod. 2013;43:732–737.

293. Yuan Y, Chesnutt BM, Utturkar G, et al. The effect of cross-linking of chitosan microspheres with genipin on protein release. Carboh Polym. 2007;68:561–567.

294. Zainuddin SYZ, Ahmad I, Kargarzadeh H. Cassava starch biocomposites reinforced with cellulose nanocrystals from kenaf fibers. Compos Interface. 2013;20:189–199.

295. Zhang XQ, Burgar I, Do MD, Lourbakos E. Intermolecular interactions and phase structures of plasticized wheat proteins materials. Biomacromol. 2005;6:1661–1671.

296. Zhang X, Do MD, Casey P, et al. Chemical cross-linking gelatin with natural phenolic compounds as studied by high-resolution NMR spectroscopy. Biomacromol. 2010a;11:1125–1132.

297. Zhang X, Do MD, Casey P, et al. Chemical modification of gelatin by a natural phenolic cross-linker, tannic acid. J Agric Food Chem. 2010b;58:6809–6815.

298. Zhang Y, Han JH. Mechanical and thermal characteristics of pea starch films plasticized with monosaccharides and polyols. J Food Sci. 2006;71:E109–E118.

299. Zhang Y, Pitkänen L, Douglade J, Tenkanen M, Remond C, Joly C. Wheat bran arabinoxylans: chemical structure and film properties of three isolated fractions. Carboh Polym. 2011;86:852–859.

300. Zhou JJ, Wang SY, Gunasekaran S. Preparation and characterization of whey protein film incorporated with TiO2 nanoparticles. J Food Sci. 2009;74:N50–N56.

301. Zimmermann T, Bordeanu N, Strub E. Properties of nanofibrillated cellulose from different raw materials and its reinforcement potential. Carboh Polym. 2010;79:1086–1093.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset