Next Article in Journal
Hypoxia Suppresses Spontaneous Mineralization and Osteogenic Differentiation of Mesenchymal Stem Cells via IGFBP3 Up-Regulation
Previous Article in Journal
Guidance of Signaling Activations by Cadherins and Integrins in Epithelial Ovarian Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Physical, Chemical and Biochemical Modifications of Protein-Based Films and Coatings: An Extensive Review

1
Fraunhofer Institute for Process Engineering and Packaging IVV, Giggenhauser Strasse 35, Freising 85354, Germany
2
Chair of Food Packaging Technology, Technische Universität München, Weihenstephaner Steig 22, Freising 85354, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(9), 1376; https://doi.org/10.3390/ijms17091376
Submission received: 17 June 2016 / Revised: 17 July 2016 / Accepted: 15 August 2016 / Published: 23 August 2016
(This article belongs to the Section Materials Science)

Abstract

:
Protein-based films and coatings are an interesting alternative to traditional petroleum-based materials. However, their mechanical and barrier properties need to be enhanced in order to match those of the latter. Physical, chemical, and biochemical methods can be used for this purpose. The aim of this article is to provide an overview of the effects of various treatments on whey, soy, and wheat gluten protein-based films and coatings. These three protein sources have been chosen since they are among the most abundantly used and are well described in the literature. Similar behavior might be expected for other protein sources. Most of the modifications are still not fully understood at a fundamental level, but all the methods discussed change the properties of the proteins and resulting products. Mastering these modifications is an important step towards the industrial implementation of protein-based films.

1. Introduction

The importance of food packaging materials has increased enormously over the last 70 years during the change from a society based on rural self-sufficiency to the highly industrialized food industry of today. The focus has constantly been the protection of foods from mechanical damage, water vapor, and oxygen and the improvement of the shelf life of the packaged goods [1,2]. Accordingly, interest in protein-based films and coatings has increased considerably over recent years due to their advantages over conventional petroleum-based materials and over other biological materials such as polysaccharides and lipids [3]. Protein-based materials are usually biodegradable and are extracted from renewable sources. Furthermore, proteins are generally superior to polysaccharides in their ability to form films with high mechanical and barrier properties and they provide higher nutritional value [4,5]. Moreover, they can be used for controlled release of additives and bioactive compounds [6,7,8,9]. Thus, many studies have been conducted in order to produce films and coatings from various protein sources such as whey [10,11,12,13], soy [14,15], wheat gluten (WG) [16,17], pea [18], and amaranth [19]. Whey, along with soy and wheat gluten proteins (WG), is one of the most interesting materials as it is one of the world’s largest and essentially unexploited protein sources [5,20]. Soy and wheat gluten proteins are also abundantly available [21,22,23,24].
Protein-based films and coatings have good mechanical, optical, and oxygen barrier properties but also have high sensitivity to moisture and poor water vapor barrier properties due to their hydrophilic nature [25,26]. The required characteristics of a packaging material vary with its intended use and thus have to be customized. Many studies have therefore been conducted to optimize the functional properties of protein-based films and coatings using a variety of methods [27,28,29].
The aim of this review is to provide an overview of the effects and mechanisms of the physical, chemical, and biochemical treatment of protein-based films and coatings. Special attention is placed on the mechanical and barrier properties and the emphasis is put on whey, soy, and wheat gluten proteins which are among the most widely studied.

2. Protein Films and Coatings

2.1. Definition and Characteristics of Proteins

Proteins are organic macromolecules composed of α-amino acids which are linked by peptide bonds to form the primary structure. They differ from peptides by their larger size of approximately 100 amino acids. The polypeptide backbone is essential for the folding of the protein which is designated as the secondary structure. Stabilized by intramolecular interactions, α-helices and β-sheets are established as well as loops and bends. The tertiary structure is the global configuration of proteins, resulting from intermolecular interactions of the protein side chains. Moreover, certain proteins develop a quaternary structure. This means they are able to form relatively loose and reversible molecular aggregates with a specific geometry. The protein structure is especially important for film formation since it determines the ability of proteins to interact with themselves and other components [30,31].
The reactivity of the individual amino acid side chains plays a major role for protein modification. Amongst others, glutamic and aspartic acids as well as lysine are very important. Glutamic and aspartic acids are important because of their carboxyl groups and lysine because of its amino group. The reactivity of these side chains depends on the different reagents and also on their positions within the protein structure and on the pH [32].

2.2. Characteristics of Protein-Based Films and Coatings

A film is a preformed thin layer which can be placed around or between foods, while a coating is directly formed on food products as a coat [33,34]. In order to compare the effects of various treatments on protein films, their mechanical and barrier properties have to be determined. The mechanical characteristics described in the relevant literature are as follows: (a) the tensile strength (TS) which is the pulling force per film cross-sectional area required to break the film; (b) the elongation at break for the degree to which the film can stretch before breaking, and (c) the elastic modulus or Young’s modulus which provides information about a film’s resistance to deformation. As for the barrier properties, the water vapor permeability (WVP) along with relative humidity (RH) and the oxygen permeability (OP) are considered [35,36,37,38].

2.3. Processing of Protein-Based Films and Coatings

There are two technological processes used to make these protein-based materials, namely the “solvent process” and the “thermoplastic process” [39]. During the solvent process, the proteins are dispersed and solubilized in solvents such as water or ethanol, followed by casting, spraying, or dipping (shaping) and then drying. With the aid of plasticizers and rising temperature, proteins pass the glass transition during the thermoplastic process. Consequently, a rubbery mass is created which can be shaped and is stabilized by cooling or by eliminating the volatile plasticizers (e.g., water) [40,41,42]. Thermoplastic processes such as compression molding and extrusion are vital for industrial applications because they allow the profitable production of films and coatings on a larger scale [43,44].

2.4. Whey Proteins

There has been a large amount of interest in recent years in the production of whey protein based films and coatings. Whey proteins are a by-product from the precipitation of proteins in milk and during cheese manufacturing [45,46]. Whey is a dilute nutrient stream. It can be dried to provide whey powder. Depending on the protein content, the powder is called either whey protein concentrate (25%–80%, WPC) or whey protein isolate (WPI) which contains >90% protein on a dry weight basis [47].
The major proteins in whey products are α-lactalbumins (α-La), β-lactoglobulins (β-Lg), bovine serum albumin, immunoglobulins, and proteose-peptones. β-Lg is the most prevalent protein in the whey protein fraction. It makes up about 57% of the protein in whey [48]. Monomeric β-lactoglobulins contain one free sulfhydryl group and two disulfide groups [49]. Studies have shown that β-Lg exists in a globular form, with a hydrophobic center which can be involved in the binding of hydrophobic molecules [50]. The second most abundant protein in whey, α-lactalbumin, accounts for about 19% of the total whey protein [48]. It is a globular protein with great flexibility due to approximately 61% unordered secondary structure [51]. Nevertheless, bound calcium is responsible for intermolecular ionic bonding and S–S bridges maintain sufficient stabilization. In addition, disulfide interchange reactions occur with the β-lactoglobulin component [26]. Bovine serum albumin (BSA, 7%) is also a globular protein. It contains 17 disulfide bonds and one free thiol group and is therefore highly relevant for the formation of the films. Due to effective binding of free fatty acids and other lipids, BSA is stabilized against denaturation [52]. Immunoglobulins and proteose-peptones respectively make up to 6% and 11% approximately of total whey proteins [30].

2.5. Soy Proteins

Soy protein is extracted from soybeans during the production of soy oil. Soy flour is a secondary product and it can be purified to obtain soy protein isolate (SPI) [53]. Soy protein isolate is a mixture of proteins having different molecular properties. About 90% of soy proteins are globulins [54]. Globulins are protein fractions in which the subunits are associated via hydrophobic and hydrogen bonding [55]. The soy protein globulins can be fractionated into 2S, 7S, 11S, and 15S according to their sedimentation coefficients. The 7S (β-conglycinin) and 11S (glycinin) fractions make up about 37% and 31% of the total extractable proteins and have the ability to polymerize [56,57]. Structural differences mean there are variations in the functional properties of the 7S and 11S fractions [58]. The extensively glycosylated 7S protein consists of three peptide subunits (α, α′, and β) which results in different film formation depending on their combinations [59]. On the other hand, the sulfhydryl groups of the 11S protein were reported to be responsible for the formation of disulfide links which result in the formation of a three dimensional network [54,60].

2.6. Wheat Gluten Proteins

Wheat gluten proteins are the storage proteins of wheat. The water insoluble fractions of the wheat proteins are prolamin and glutelin. In wheat, the prolamin fraction is called gliadin and the glutelin proteins are referred to as glutenin. Both represent respectively about 33% and 46% of the total protein content in wheat. Prolamin is considered as the solvent for glutelin and therefore the main factor responsible for the viscosity of gluten. On the other hand, the fibrous glutelin fraction in gluten defines its elasticity and firmness. The glutelin fraction can be divided into two different subfractions, high molecular wheat glutelin (HMW) and low molecular wheat glutelin (LMW). Moreover, the HMW subunit can be separated into two types, x and y. Likewise, the prolamin fraction consists of ω5-, ω1,2-gliadine as well as α- and γ-gliadin. The content of subunits varies with the wheat variety. According to the molecular mass of the gliadin and glutenin subunits, three distinct groups of proteins can be defined: high molecular mass, medium molecular mass, and low molecular mass. These groups are shown in Table 1 with their wheat gluten and cysteine amino acid contents [61,62,63].

3. Physical Modifications of Protein-Based Films and Coatings

3.1. Heating

It is well known that proteins usually tend to irreversibly aggregate and eventually form gel networks when exposed to high temperatures. Increasing the temperature results in greater mobility of the peptide chains, so changing the protein structure and aggregation state. This leads to different inter and intramolecular hydrophobic and electrostatic interactions as well as hydrogen and disulfide bonds due to the presence of cysteine residues. The properties of the resulting protein network strongly depend on the temperature and also on the ionic strength and the presence of other molecules [21,64,65]. In the literature this treatment is often described as heat curing which is, according to Soroka [66], a process where a substrate is exposed to one or more heating cycles aimed at changing the molecular structure and rearranging the polymers.
Thermal denaturation of whey proteins begins at approximately 70 °C [67], whilst for gluten proteins it starts at 90 °C [68] and for soy proteins at 90 °C [69]. However, the denaturation temperature strongly depends on the protein concentration and the solvent properties [70]. For example, in an extrusion process the soy protein denaturation temperature was found to be 120 °C [71].
The thermal pretreatment of proteins, and also heating a formed protein film, leads to augmented crosslinking of the disulphid, hydrogen, and hydrophobic bonding types. Thus, whey, soy, and gluten based films processed under increased temperature show a significantly higher tensile strength [17,23,27,72,73,74,75,76,77,78,79,80,81,82].
Vachon et al. [77] ascertained that pretreated WPC-based films are weaker than WPIs. Perez-Gago [79] reported a significant lower oxygen permeability after heat treatment of whey protein films. Heat treatments also involving a reduction of the water content increase the film’s strength and rigidity [80,81]. Barreto et al. [83] investigated the thermal denaturation of WPC films and reported two denaturation stages. The first was due to water loss up to 200 °C and the second involved degradation starting at 295 °C. The thermal stability was highly dependent on other molecules in the film.
Heat treatment of soy protein prior to film formation produces smoother and more transparent films with reduced water vapor permeability [84]. Furthermore, the heat curing of soy and gluten protein films increases their elongation properties [23,78,84].

3.2. Shearing

Shearing can have three distinct effects on proteins: aggregation, de-aggregation, and protein denaturation. However, extremely high shear rates are necessary to denature proteins [85].
The breakup of aggregates is due to erosion of the particle surface and pressure changes in the fluid leading to deformation and fragmentation of the particles [86,87]. On the other hand, an increased collision rate can favor the formation of aggregates. The collision of small particles (below 1 µm) is driven by Brownian motion while larger particles are influenced by the fluid flow induced by shear. The presence of larger and more numerous protein aggregates therefore increase the number of collisions during shear treatment [88,89].
Steventon et al. [86] reported a break up of whey protein aggregates during the shear treatment and short period of heating in an extrusion process. The opposite was observed by Cheftel et al. [90] for whey and soy protein while processing with longer heating time. More recently, Wolz et al. [91] ascertained that increasing the whey protein concentration up to 30% results in smaller, more compact, and stable aggregates. This is due to the higher viscosity and shear stress. The particle size increases at first at low protein concentrations due to higher collision rates. The higher shear stress occurring with increased protein concentration leads to the formation of smaller particles. Fang et al. [92] reported that mechanical shearing during the extrusion of soy proteins also reduced the size of its particles.
Pommet and Redl et al. [93] stated that high shear rates significantly lower the activation energy for the crosslinking of gluten proteins.

3.3. Hydrostatic Pressure

3.3.1. Mechanism and Effects of Hydrostatic Pressure

Hydrostatic pressure (HP) processing uses water as a medium to transmit pressure from 100 to 1000 MPa to foods under isothermal conditions [94,95,96,97]. The effect of pressure on proteins is often described by the principle of Le Chatelier, whereby a system reduces its free energy by minimizing the effect of the external factor. Consequently, a change in pressure is compensated by modification of the system volume [98,99]. In solution, the volume of a protein is determined by the volume of its atoms and the internal cavities due to imperfect packaging of the amino acid residues. On the other hand, a volume decrease results from peptide bonds and polar amino acid hydration [100]. Therefore, almost all globular proteins have a positive compressibility due to their cavities [101].
HP thus disrupts intermolecular hydrophobic and electrostatic interactions whose formation result in a volume increase. It also appears that HP increases the reactivity of sulfhydryl groups. However, high pressure stabilizes hydrogen bonds since their formation slightly reduces the volume. Covalent bonds are not affected. Consequently, high hydrostatic pressure changes the quaternary, tertiary, and secondary conformations of proteins as well as their ionic and hydrophobic stabilized aggregates but has no influence on the primary structure. Proteins subsequently unfold and, if their concentrations are high enough, form gel networks and precipitate [102,103,104,105,106,107].
Hydrophobic interactions are strengthened or weakened depending on the protein itself, its state in solution, and other unsettled influences [108,109]. Hence, the behavior of proteins under pressure cannot by determined from known structural models and thus require individual testing [110].

3.3.2. Physical Influence of HP on Whey, Soy, and Gluten Protein Films and Gels

Up until now only a few studies on the influence of HP processing on protein-based films and coatings have been published [19,111]. A patent has been granted for using this process to produce films [112]. This section therefore mainly summarizes the effects of HP on whey, soy, and WG protein gels used to produce edible films and coatings.
High pressure affects the conformation of whey proteins and dissociates large aggregates, exposing hydrophobic groups by unfolding and allowing the formation of intermolecular sulfhydryl bonds [104,113,114,115,116,117,118,119]. Compared to thermal processing, pressure induced β-Lg gels appear to form a more porous and thicker stranded structure with weaker intermolecular interactions. The resulting gels have stronger water exudation and water solubility, lower rigidity, and the proteins tend to aggregate during storage [116]. Increased pressure leads to harder gels with higher breaking stress and lower solubility [120,121]. However, van Camp et al. [122] produced WPC gels with comparable strength to those induced by heat and even stronger with high protein concentration. Gentle thermal pretreatment (55 °C) of whey protein improves film forming ability but higher temperatures have a negative effect due to protein aggregation [123].
The formation of high pressure induced soy protein gels has also been reported. The gels had lower hardness than the ones produced by thermal treatment and had high water holding capacity [124,125]. Speroni et al. [126] showed that β-conglycinin and glycinin formed hydrophobic interactions and disulfide bonds during HP processing. However, HP induced weak gels with low elasticity and diminished the ability to form hydrophobic interactions on heating. Increasing pressure and holding time leads to a stronger soy protein gel network with stronger hydrophobicity possibly crosslinked by disulfide or hydrogen bonds [127,128,129]. The formation of random coiled structures from β-sheets and α-helices by HP was observed [130]. Subirade et al. [131] indicated that β-structures might be essential for the formation of soy protein film networks.
Low pressure processing (200 MPa) of WG results in an increase in the ethanol soluble fraction and thiol content. It seems that the α- and γ-gliadins in the cysteine are sensitive to pressure while the ω-gliadins only change their conformation and transfer to the ethanol soluble fraction. The glutenin fractions, which are relatively rich in thiol groups, are also strongly affected by high pressure. Higher pressure and temperature cause significant gluten strengthening allowing the production of stronger gels than solely by heat treatment [116,122,132]. Apichartsrangkoon [133] stated that significant S–S crosslinking only occurs under extreme conditions (800 MPa for 50 min).
In a gluten-soy blend gel, the large gluten proteins showed much greater gel strengthening influenced by higher concentration, pressure, and temperature than the smaller soy proteins [134].

3.4. Ultrasound

3.4.1. Mechanism and Effects of Ultrasound

Ultrasound is defined as an acoustic wave with a frequency higher than 20 kHz, which is the upper threshold of human auditory detection [135]. It can by differentiated into two categories which are described differently depending on the author and application: Low and high energy, low and high power, low and high intensity, and so on. Low energy ultrasound ranges from 100 kHz to 1 MHz at intensities lower than 1 W·cm−2 while high energy ultrasound ranges from 16 to 100 kHz at intensities over 1 W·cm−2 [136,137,138,139].
High intensity ultrasound in a liquid generates cavitation due to the compression and decompression cycles of the sonic waves. The violent collapse of the gas bubbles results in high shearing effects. These shear forces and the energy inputs are strong enough to break covalent bonds of proteins dissolved in aqueous solutions [140,141]. Furthermore, regions of high local pressure and temperature up to 50,000 kPa and 5000 K are induced [142,143].
Due to the thermal decomposition of water, free hydroxyl and hydrogen radicals are formed [144,145]. This sonolysis of water, along with the shear forces, is mainly responsible for the denaturation of proteins by high intensity ultrasound [146,147]. The high temperatures are thought to have a negligible effect on biotechnological processes due to the narrow ranges. However, the gas bubbles generate microstreaming favoring the convection of reactive components. Consequently, chemical reactions taking place in proteins are accelerated. Thus, the turbulence created by ultra-sonication can also be used for homogenization. The strength of these effects is strongly dependent on the wave characteristics and distance from the emitting electrodes [148,149].

3.4.2. Influence of Ultrasonic Processing on Protein-Based Films and Coatings

Kadam et al. [150] investigated the effect of ultrasound on whey protein isolate films containing nanoparticles. Before casting, the WPI solutions were sonicated at increasing amplitudes to obtain a homogenous protein-based film. The results showed there was a significant increase in film strength, elasticity, and hydrophobicity with higher sonication amplitude. This was stated as being due to improvement of the homogenous distribution of the film by sonication. However, the water vapor permeability remained unchanged. The same results had previously been obtained by Barnerjee et al. [151]. This could be explained by the smaller particles in film forming solutions obtained by the sonication process. The increased molecular interactions due to the energy input could lead to higher molecular order and therefore increased film strength [15]. Chen et al. [152] reported that intense ultrasonic treatment of whey drives off more moisture from the protein films. This could not be confirmed by Barnerjee and Kadam who observed no significant changes in the moisture content.
Rodriguez et al. [153] investigated the effect of ultrasound treated whey protein coating on the quality of frozen fish. The sonicated coatings exhibited considerably lower lipid oxidation than the untreated coatings and no sensory changes could be detected. Although protein films generally display a good O2 barrier [154], chemical mechanisms could not been excluded.
Furthermore, Guzey et al. [155] determined that high-intensity ultrasonic processing of BSA increases its intramolecular mobility and surface activity.
Jambrak et al. [156] observed an increased solubility and specific surface area in soy protein treated by low-frequency ultrasound. These changes were partially explained by free hydroxyl radical formation. The same explanation was given by Wang et al. [157] as the vapor and oxygen barrier of a soy protein containing film was ameliorated by this process. The same main author also observed increased hydrophobicity in another soy protein containing film treated by ultrasound and attributed this to the cavitation improving the film density [158]. Furthermore, Hu et al. [159] ascertained that the free sulfhydryl content, surface hydrophobicity, and protein solubility of soy protein solutions increased with low-frequency ultrasonic treatment. The pretreatment also bestowed better water holding capacity and gel strength but no changes in the particle size distribution.
Ultrasonic treatment dissolved gluten protein aggregates prior to gluten protein film formation and therefore changed its appearance. Nonetheless, the expected solubility change from the dissolution could not been measured [160,161].

3.5. Ultraviolet and γ Irradiation

Radiation can be absorbed by atoms and molecules in a system. The captured energy is converted into chemical energy, inducing photoisomerization. Consequently, the exposure of double bonds and aromatic rings of proteins to UV radiation (180–400 nm) lead to free radical formation in amino acids such as tyrosine and phenylalanine. These changes induce the formation of intermolecular covalent bonds [162,163]. Ionizing radiation such as γ and high-frequency UV radiation additionally causes oxidation of amino acids, rupture of covalent bonds, formation of protein free radicals, and water radiolysis producing free oxygen radicals. Brault et al. [164] demonstrated that γ-irradiation induces the formation of bityrosine bridges in caseinate-based films. Irradiation therefore can have a direct influence on proteins as well as an indirect influence by impairing the surroundings [162,165,166,167,168,169,170,171,172].

3.5.1. Effect of Ultraviolet Irradiation on Protein-Based Films and Coatings

Ustunol et al. [173] observed a significant increase in the tensile strength of whey-based films by UV irradiation of the protein solution prior to film formation. However, the radiation treatment had no influence on the barrier properties and elongation at break of the films. The same effects were reported by Schmid et al. [13] who investigated the impact of UV irradiation on cast whey protein films. The German research team also concluded that increasing the radiation dose leads to increased molecular interactions within the protein network.
Gennadios et al. [162] reported a linear increase in the tensile strength and a linear decrease in the elongation at break with increasing UV treatment intensity (λ = 253.7 nm up to 103.7 J·m−2) of cast soy protein films. These results could not be confirmed by Vaz et al. [174] who treated the soy protein solution and the films at a wavelength of λ = 366 nm and 4.5 J·m−2. This research team only observed a small increase in the mechanical properties of the films with no change in the protein network.
An increase in tensile strength was ascertained by Rhim et al. [163] after UV irradiation (253.7 nm, 51.8 J·m−2) of cast wheat gluten protein sheets. However, Micard et al. [27] reported no significant change in the mechanical properties after UV treatment (λ = 254 nm and 1 J·m−2) of films based on the same protein.
Furthermore, increased yellowness of whey, soy, and wheat gluten films has been observed after UV irradiation [13,162,163].

3.5.2. Effect of γ Irradiation on Protein-Based Films and Coatings

γ irradiation (32 kGy) of WPI solutions increases the content of β-strands and β-sheets. This could explain the finer stranded network and increased fracture strength of the whey protein gels and films cast with the irradiated solutions [175]. Cho et al. [169] reported that γ-irradiation of BSA and β-Lg in solutions caused the disruption and degradation of the protein structure as well as crosslinking and aggregation of the polypeptide chains. Furthermore, Ouattara et al. [176] detected a significant decrease in the water vapor permeability and an increase in the molecular weight of the protein particles in solution after irradiation using the same dose.
Lacroix et al. [177] investigated the effect of γ irradiation prior to casting SPI and SPI/WPI blend films. There was an increase in the puncture strength of both films after radiation treatment at 128 kGy. However, there was only a reduction in the water vapor permeability for the SPI film, while the blend film remained unchanged. Sabato et al. [60] also reported an increase in puncture strength and puncture deformation after γ-irradiation (32 kGy) of SPI and SPI/WPI films. Here, the effect was also greater on the SPI film.
Micard et al. [27] reported an increase in TS and WVP as well as a decrease in elongation for γ-irradiated gluten films (10 kGy). Higher radiation doses up to 40 kGy had the inverse effect. This could be explained by the decrease in glutenin solubility by breaking down covalent linkages and de-polymerization [164,178].
The effect of γ-irradiation on SPI and gluten solutions prior to film formation was investigated by Lee et al. [179]. The viscosity of the gluten-based films decreased at a radiation dose below 16 kGy due to cleavage but increased at higher dose up to 50 kGy due to protein aggregation. As for the SPI-based films, only a decrease in viscosity was reported. Furthermore, the WVP decreased by up to 13% for the SPI-based film and by 29% for the gluten-based film while the TS increased by respectively 2- and 1.5-fold at 50 kGy.

3.6. Thermoplastic Processing

Increasing the chain mobility by denaturation of the film forming proteins is essential for thermoplastic processing. Changes in chain mobility are related to thermal transitions such as the glass transition temperature (Tg) and flow temperature (Tf), also referred to as the softening temperature, which occurs at higher temperature. Above the Tf the protein-based mixture exhibits the low viscosity necessary to process the material. Unfortunately, proteins with less than 5% water have a Tg above or equal to their decomposition temperature [180]. The temperature difference between the Tg and Tf of soy and gluten based films is about 40 °C [181]. Moreover, films based solely on proteins are fragile and brittle because of the bonds and interactions between the protein chains [182]. As shown in Figure 1, the addition of plasticizers resolves this issue by impairing macromolecular associations and therefore reducing the Tg and Tf. The chain mobility increases on addition of plasticizers, due to replacement of protein interactions with protein-plasticizer interactions [183].
Water is the most effective plasticizer but it increases the melt viscosity at higher content [180]. This leads to reduced protein transformation resulting from the small temperature increase due to the small motor torque and specific mechanical input [184]. Therefore, other plasticizers varying in size, shape, composition, and hydrophilicity are used depending on the protein source and the use of the thermoplastic product [185]. As an example, glycerol inserts itself within the polymer network [186] and is therefore an excellent plasticizer widely used for the thermoplastic processing of proteins [82,93,184,187,188,189,190,191].

3.6.1. Compression Molding

In compression molding, protein-plasticizer mixtures are placed in an open mold. Subsequently, a plunger applies pressure forcing the material to assume the desired form [192]. This usually occurs relatively quickly at low moisture contents, high temperature, and pressure in order to transform the blends into viscoelastic melts. The combination of heat and pressure leads to the denaturation of the proteins [193]. The subsequent cooling of the product determines its form by covalent, ionic, hydrogen bounding as well as hydrophobic and hydrophilic interactions [108]. Knowledge about optimal ingredient ratio and processing parameters are also vital for film formation by extrusion [82,191].
Sothornovit et al. [191] formed transparent WPI-based films by compression molding with 30% to 50% water or glycerol. The films processed with water were brittle, while those containing glycerol as plasticizer were flexible. The same research team also investigated the effects of the plasticizer content and processing conditions compared to casting. The processing pressure and temperature had a negligible effect on the mechanical properties of the protein sheets produced at up to 140 °C and 2.25 MPa for 2 min. However, the tensile strength and elastic modulus were smaller at higher glycerol content [82].
Soy protein films molded under compression at high temperatures up to 150 °C with glycerol usually show higher elongation and tensile strength than the ones obtained by conventional casting [193,194,195,196]. The soy protein-based films produced by thermoplastic processing are also more transparent [53,196]. Ciannamea et al. [196] reported lower WVP but higher OP for the soy protein films. The same research group also ascertained that disulfide bonds were mainly responsible for the characteristics of the compression molded films whereas hydrogen bridges and hydrophobic interactions dominate in the cast soy protein films.
Increasing the processing temperature during compression molding leads to higher crosslinking of the wheat gluten protein network. This results in improved tensile strength and lower WVP and OP [197,198]. The activation energy for the crosslinking was found to be 170 kJ/mol by Pommet et al. [93]. However, longer holding times seem to have a smaller effect than the temperature [43].

3.6.2. Extrusion

Up until now, extrusion is the most commonly used process for polymer production [199]. It is a continuous process whereby the raw materials are constantly introduced into a hopper feeding a horizontal barrel. The extrudate is subsequently conveyed by one or two rotating augers and finally pushed through a die. This process can involve various operations such as heating, shearing, mixing, compressing, melting, and shaping. Generally, the extruder barrel is composed of three sections: the feeding, transition, and metering section. The ingredients are introduced into the feeding section where also mixing, degassing, and slight compression occur. The continuous rotation of the screw(s) also moves the mixture to the transition zone. As indicated by the name of this section, this is where the raw material changes into extrudate whilst the pressure and temperature increase. This effect is induced by reduction of the flow channels, so compressing the product and dissipating mechanical energy. In the subsequent metering section, also referred to as the heating or cooking zone, the highest temperatures, pressures, and shear rates are experienced. This is also where the product acquires its final aspect before it is pressed through the die [184,200,201].
The extrusion process can proceed under various conditions using different screw configurations, lengths, diameters, speeds, temperature profiles, feeding rates, and the addition of ingredients at the beginning and during the process [184].
Onwulata et al. [202] extruded a WPC mixture containing 38% moisture with a twin screw extruder at temperatures ranging from 35 to 100 °C. Under these conditions the degree of denaturation of the proteins increased from 30% to 95%. The processed proteins where subsequently dried, mixed with water, and then heated to form gel samples. The gels obtained from proteins extruded at the lowest temperatures up to 45 °C were stronger than the ones produced with unprocessed WPI. However, the gels made with proteins processed at higher temperatures were weaker. Hernandez et al. [184] investigated the effect of glycerol and moisture content on the thermal transitions of WPI during extrusion. There was a decrease in TS and elasticity with increasing glycerol content. The same main author also investigated the heat seal strength of extruded whey protein films [203]. In this study, extruded sheets showed lower seal strength then cast sheets since they were thicker and therefore harder to seal. In a study conducted by Qi and Onwulata [204] higher moisture content slightly increased the protein solubility and reduced the content of β-Lg while the α-La remained unchanged. Schmid et al. [41,42] extruded WPI with ethylene vinyl acetate (EVA). The resulting sheets exhibited higher WVP with increasing WPI content compared to pure EVA films.
Zhang et al. [205] investigated the mechanical and thermal properties of soy protein films extruded with various plasticizers. The film sheets were relatively strong and elastic except the ones with high moisture and glycerol content. It was ascertained by Chen et al. [206] that non-covalent bonds were more important for the soy’s extrudate structure than covalent bonds.

4. Chemical Modifications of Protein-Based Films and Coatings

This section deals with chemical modifications. This includes reactions with chemical agents and modification by pH alteration. For these modifications the protein side chains play a major role. Table 2 shows the reactive groups of the side chains and their occurrence in selected proteins. In addition to the side chain composition, the location within the protein and the external conditions such as the pH also influence the reaction rates [32,207].

4.1. Reactions with Chemical Agents

4.1.1. Alkylation

The substitution or addition of alkyl groups in organic compounds is called alkylation [207]. The alkylation of proteins mainly takes place at the amino groups of protein side chains. Therefore lysine is essential for this reaction. For example, the addition of formaldehyde in combination with sodium borohydride to the film building solution leads to reductive methylation (Figure 2). The reaction starts with the condensation of the amino group by a carboxyl group, resulting in the formation of an imine. The imine is subsequently reduced by a mild reducing agent, such as sodium borohydride, which is oxidized itself at the same time. Thereby methylamino groups form. These are immediately transformed into dimethylamino groups by additional formaldehyde and reducing agent, whereby formaldehyde acts as the oxidant. The formation of these dimethylamino groups, which replace the initially amino groups, is responsible for the change in the functional properties of the proteins [32]. The reaction is illustrated below.
Reductive alkylation is not a commonly used method for improving the functionality of proteins. However, some studies have demonstrated slight modifications. For example, Kester et al. confirmed changes to the functional properties of proteins [207]. However, the effects of this chemical modification are so slight that reductive alkylation was used to perform studies relating to the protein structure, such as radiolabeling. In order to significantly enhance the functional properties of proteins, formaldehyde has to be replaced by carboxyl compounds with a more complex structure. For example, propylenglycol alginate, with sodium cyanoborohydrid as a reducing agent, is used in soy protein films to improve the mechanical properties [209].

4.1.2. Acylation

For an acylation reaction a protein must have a nucleophile amino acid residue, for example an amino or phenol group [207,210]. Such residues react with acylating agents which possess a carbonyl group, such as activated acid anhydrides [30,210]. The acylation of these groups leads to enhanced functional properties, especially in food proteins [211,212,213].

4.1.3. Acetylation

Acetylation, a specific form of acylation, is the integration of an acetyl group into compounds which contain amino, hydroxyl, or thiol groups. In order to accelerate the reaction, acid or basic catalysts, such as zinc chloride or sulfuric acid, can be added [214].
Acetic anhydrides are used for the acetylation of proteins (Figure 3). This leads to covalent bonding of acetyl groups with the amino groups of the protein. Due to the fact that oppositely charged amino acid side chains attract each other, a reduction in the number of amino groups leads to partial unfolding of the protein backbone [207] Thereby, for example, the gel strength and water binding capacity of soy protein isolate (SPI) decrease. The solubility in the pH range from 4.5 to 7 becomes greater [216]. However, Franzen et al. [217] found that acetylation only has a minor effect on the functional properties of soy protein. Compared to the effect of succinylation, the effect of acetylation on the protein structure is significantly smaller [207]. Additionally, Ghorpade et al. [218] showed that acylation with acetic anhydride could not improve the tensile strength nor the water vapor permeation (WVP) or oxygen permeation (OP) of soy protein films.

4.1.4. Succinylation

Succinylation is also a specific form of the acylation reaction. The acylation of amino acid residues in the succinylation reaction is performed with dicarboxylic acid anhydrides, for example succinic acid anhydride [217]. Thereby, H2O, which results from combination of a hydroxyl group of the carbonyl group of the acid anhydride with a hydrogen atom of the functional groups of the amino acids, is split off [30]. This reaction is shown in the Figure 4 below.
In general, succinylation leads to improved solubility of the proteins [30]. The acylation mainly occurs at the ε-amino group of lysine. Thereby, the cationic amino group is transformed into an anionic residue. This leads to a change in the foaming properties as well as in the emulsifying capacity. Increased aqueous solubility is a further consequence of the raised net negative charge induced by succinate anions [217]. Barber et al. [219] confirmed this for wheat gluten. Their research showed that succinylation of wheat gluten is able to increase the solubility, emulsification capacity, water adsorption capacity, and water holding capacity. Also, glycinin with 25% succinylation was found to have an increased surface pressure as well as increased film yield stress, film elasticity, and foam stability [220]. In contrast, soy protein isolate films which were succinylated showed neither improved mechanical properties nor improved barrier properties compared to the non-succinylated films. However, they showed enhanced water solubility [109].

4.1.5. Incorporation of Fatty Acid Chlorides (Grafting)

The incorporation of fatty acid chlorides represents an acylation reaction. Palmitic acid chloride is commonly used [221]. The process is mainly based on Schotten-Baumann’s reaction [222,223]. For incorporation into protein structures an alkaline medium is recommended. The reaction equation is shown below (Figure 5) [224].
This leads to the integration of long alkyl chains which act as internal plasticizers. These are able to reduce or even interrupt intermolecular interactions between the protein side chains. This causes changes to the thermal properties and protein folding. An important factor is the number of available functional groups, because they determine the number of bonded fatty acid chlorides. For example, Bräuer et al. [221] was able to integrate 4 mmol·g−1 palmitoyl groups into wheat gluten as well as into soy protein. Incorporation into corn zein and pea protein was significantly lower. Another factor which influences the substitution of ε-amino groups with fatty acids is the solvent. The dilution of, for example, β-lactoglobulin in an organic solvent showed higher incorporation rates than under aqueous conditions [226]. The acylation has various effects on the functional properties of the protein. First of all, the modified proteins show higher hydrophobicity and therefore higher water resistance. This leads to lower water absorption and the proteins are less hygroscopic, which contributes to poorer welling. Another positive effect concerns the thermal properties. The modified proteins, for example the acylated palmitoyl derivatives, melt between 150 °C and 200 °C and therefore show better thermal characteristics than the native proteins, which cannot be melted. This can be due to the reduction in the intermolecular interactions, especially hydrogen bonding, induced by the modification with fatty acid chlorides [221]. The incorporation of fatty acids is able to improve the moisture barrier of protein films and coatings [29].
Considering all the reactions with chemical agents, it is clear that neither alkylation nor acylation methods are able to improve the technofunctional properties significantly. However, the incorporation of fatty acids is a promising acylation method, although very little research has been carried out to date on this. A big problem with alkylated films is the mostly toxic reagents which lead to non-edible films [209,218].

4.2. Modification by pH Alteration

4.2.1. Hydrolysis

Little research has hitherto been devoted to the non-enzymatic hydrolysis of food proteins. Acidic and basic hydrolysis of proteins is mainly used to determine the amino acid structure [227,228]. The mechanisms involved here are shown in the Figure 6 below.
Akkermans et al. [230] found that β-lactoglobulin which was heated for 20 h at 85 °C at pH 2 leads to hydrolyzed peptides with molecular masses between 2000 and 8000 Da. Similar studies showed that the hydrolysis influences the viscosity. Mudgal et al. (2011) found that β-lactoglobulin dispersions heated at pH 2 were very viscous. The power law index was about 0.33 at 7% (w/w) and about 0.17 at 8% (w/w). This means that the solutions demonstrate pseudoplastic behavior (7% (w/w)) or even form solid like gels (8% (w/w)) [231]. All in all, the effects of acidic and basic hydrolysis on the functional properties of proteins should be similar to the effects of enzymatic hydrolysis due to the fact that both reactions shorten the protein backbone by splitting peptide bonds.

4.2.2. Change in Protein Structure

An important factor for the functional properties of proteins is the pH value. This is due to the fact that proteins change their net charge at a pH value that is different to their isoelectric point. There is a positive net charge at a lower pH value and a negative net charge at a higher pH value [29].
In SPI and wheat gluten films, intermolecular interactions such as disulfide bridges, hydrogen bonds, and hydrophobic interactions are especially influenced by a pH shift. SPI films in the pH ranges 1–3 and 6–12 showed the best film formation properties while homogenous free standing WG films were achieved in the pH ranges 2–4 and 9–13. Between those ranges, more precisely pH 4–5 for SPI and pH 5–8 for WG, there was rather poor film formation or even no film formation at all. These ranges more or less correspond to the isoelectric point, which is pH 4.5 for SPI and 7.5 for wheat gluten. The main components of wheat gluten, gliadin and glutenin, have isoelectric points at 8.1 and 7.1 respectively [232,233,234]. The SPI films which were produced at an alkaline pH (6 to 11) had higher tensile strength, higher elongation at break, and lower water vapor permeability than the films produced under acidic conditions. The WG films formed in the pH range from 9 to 13 had significantly higher tensile strength compared to the films produced in a pH range from 2 to 4 [232].
In whey protein films, pH alteration also changes the quaternary structure of β-Lg. In particular, at low temperatures and at low concentration of β-Lg the pH value is responsible for changes to the quaternary structure [234,235].
As can be seen in Figure 7, a pH value smaller than 3.5 or higher than 8.0 leads to a monomer, while a pH between 3.5 and 5.2 results in an octamer. In the pH range between 5.2 and 8.0, which is particularly important because it includes the pH of milk, β-lactoglobulin forms dimers [234,235].
The effect of pH alteration at pH values between 7 and 9 on the mechanical properties of whey protein films was investigated by Anker et al. (1999). Amongst other things he found that the Elastic Modulus and the stress at break developed a maximum at the critical gel concentration at pH 7 and 8, while the elongation at break increased with rising pH value [236]. Regarding the barrier properties, it can be stated that the hydrophilic whey protein films have a poor water vapor barrier compared to other non-hydrophilic materials. As such the pH value has only a minor impact on the water vapor permeability. In the pH range from 4 to 9, the WVP at pH 5 was slightly increased whilst all the other parameters showed no significant differences. Additionally, it was attempted to form films at pH 3 but the lack of film forming ability prevented this [237].
Li and Lee [238] determined that glutenins and gliadins were mainly responsible for wheat gluten film formation during extrusion. It was also found that during this process the WG proteins primarily aggregate through intermolecular disulfide bonds and hydrophobic interactions. The authors also stated that high temperatures of about 185 °C might disrupt the aggregation forces. As for whey and soy protein films, Hochstetter et al. [239] found that the mechanical properties of extruded WG films decrease with higher moisture content. However, the OP was as low as for cast WG sheets.
All in all, a pH change in the film forming solution influences both the mechanical and barrier properties. As such the isoelectric point of the protein that is used plays an important role. The water vapor permeation and the oxygen permeation of the films have their maxima at the isoelectric point. Furthermore, both WPI and SPI films showed increasing elongation at break with increasing pH. Comparing the elongation of SPI films with WPI films during pH alteration, it is apparent that the WPI films are able to double their values, while the levels of the SPI films are between about 70% and 90% [232,236,237].

5. Biochemical Modifications of Protein-based Films and Coatings

5.1. Biochemical Transformation by Enzymes

Enzymes are proteins having catalytic activity which are synthesized by organic cells. They are able to influence a high number of reactions in proteins, for example hydrolytic reactions such as the splitting of peptide bonds, transfer reactions, redox reactions, and crosslinking. Enzymes are sensitive to pH alteration and high temperatures. They reduce the activation energy of otherwise non-spontaneous reactions and enhance the reaction speed. Enzymes are generally specific to the substrates of the catalyzed reaction or to the reaction itself [30].

5.1.1. Enzymatic Hydrolysis

Proteins are formed by acid-amino-bonds, also called peptide bonds, which link the amino acids together. Hydrolysis leads to the cleavage of these bonds, whereby the proteins end up again as amino acids (Figure 8) [30].
Partial enzymatic hydrolysis is able to improve the solubility of many proteins. Enzymes which are specific for the hydrolytic splitting of peptide bonds are called peptide hydrolases or peptidases. There are two types of peptidases. Exopeptidases catalyze the cleavage of terminal peptide bonds and thus gradually separate single amino acids or dipeptides. Endopeptidases cleave the peptide bonds of nonterminal amino acids [30].
The shortening of the protein chains and consequent reduction of the molecular weight can lead to reduced intermolecular interactions between the individual protein chains. This contributes to higher flexibility of the molecule chains and increased free volume [185,240].
The degree of hydrolysis (DH) represents the amount of hydrolyzed peptide bonds. In particular it indicates the percentage of cleaved peptide bonds relative to the total number of peptide bonds. The higher the degree of hydrolysis the smaller the protein fragments [241]. The influence of the degree of hydrolysis on the mechanical properties and oxygen permeability in whey protein coatings was previously analyzed by Sothornvit and Krochta [185]. They found that the application of proteases to reduce the molecular weight of the protein chains led to an increase in the flexibility at constant oxygen permeability. Whey protein films with the required mechanical properties and adequate oxygen permeability could be cast using a decreased amount of plasticizer. Later Schmid et al. [242] performed a study with a constant amount of plasticizer and increasing concentrations of hydrolyzed whey protein. It was demonstrated that the barrier properties remained almost at the same level, while the tensile strength and Young’s modulus decreased. The enhancement of the functional properties of soy protein isolate by enzymatic hydrolysis has been cited in many publications [243,244,245,246,247]. Unfortunately, an increasing degree of hydrolysation leads to decreased mechanical properties, while the barrier properties almost remain the same [185,227,228].

5.1.2. Crosslinking by Transglutaminase

Transglutaminases are enzymes which catalyze the modification of proteins. They catalyze the acyl transfer reaction between the γ-carboxyl group of a glutamine side chain and the ε-amino group of a lysine side chain. Thereby a ε-(γ-glutamyl) lysine bond is formed between the protein chains [248]. This bond, which crosslinks the protein chains, is stable to proteolysis and improves the enzymatic, chemical, and mechanical stability [249]. However, transglutaminase catalyzes three reactions in total (see Figure 9 below).
Transglutaminase is also used to form biopolymers containing different proteins such as 11S soy protein and whey protein [250,251]. Biopolymers were also formed by the transglutaminase catalyzed crosslinking of soy protein and casein. This was demonstrated by HPLC and SDS-polyacrylamide gel electrophoresis [252,253]. Due to its high number of glutamine side chains, wheat gluten is particularly suitable for protein crosslinking using transglutaminase. Wang et al. [254] also discovered that heating increases the number of ε-(γ-glutamyl) lysine bonds in wheat gluten, due to the increasing number of accessible glutamine and lysine residues [254,255]. In whey proteins, denaturation which results in unfolding of the chains is an important factor for the crosslinking. It is the unfolding which leads to accessible lysine and glutamin residues [256]. Furthermore, in whey protein films the transglutaminase crosslinking reaction increased the average tensile strength [251]. For soy protein isolate films, the functional properties could be improved. Tang et al. [257] reported a decrease in the moisture content and the total soluble matter due to the transglutaminase treatment. The elongation at break was lowered. In addition, the tensile strength and the surface hydrophobicity were increased. However, the water vapor transmission rate and water vapor permeability remained at almost the same level.
A number of research groups have employed enzymatic crosslinking to try to increase the cohesion forces of the proteins within the matrix. Aboumahmoud and Savello [258] reported successful crosslinking of whey proteins with guinea pig liver transglutaminase. In a subsequent study they effectually used TG to covalently crosslink α-lactalbumin and β-lactoglobulin for film formation [259]. Yildirim et al. [260] manufactured a transglutaminase-crosslinked WPI and soybean film which showed good stability. Two years later, Yildirim and Hettiarachchy [251] published another attempt to improve WPI films by crosslinking them with 11S globulin. The research showed that all TG-crosslinked films were more than two times stronger than non-crosslinked films. However, the water vapor permeability (WVP) also significantly increased, making the films unusable for products where moisture levels have to be controlled. On the other hand, there may be food products requiring higher levels of humidity for which these films could be useful. Furthermore, Oh et al. [261] successfully developed TG-crosslinked WPI and casein films with lower WVP, albeit not significantly lower. A very successful crosslinking of chitosan-whey protein edible films by microbial transglutaminase was reported by Di Pierro et al. [262]. The films were found to have enhanced mechanical resistance, reduced deformability, markedly improved oxygen and carbon dioxide barrier properties, and also a lower permeability to water vapor. In another study by Hernàndez-Balada et al. [263], biopolymers produced by treating WPI and blends of gelatin with microbial transglutaminase (MTGase) were reported to have improved strength and stability. A small amount of gelatin was sufficient to realize a dramatic rise in viscosity and higher gel strength. Numerous other reports describe improved properties of transglutaminase-crosslinked whey proteins. The result is an improved gas barrier or lower WVP for WPI-based films [264,265,266,267,268].
As for the treatment of whey proteins, transglutaminase is nowadays the most frequently used enzyme for the crosslinking of soy proteins. Motoki et al. [269] crosslinked milk casein and soybean globulin using transglutaminase. Mariniello et al. [270] prepared pectin and soy flour based films in the presence of TG, resulting in smoother surfaces and higher homogeneity. The films had increased strength and reduced flexibility due to a higher degree of association between the different molecular components. Tang et al. [257] were first to report and evaluate the effects of TG treatment on the properties of basic SPI films. They investigated the effect of microbial transglutaminase (MTGase) treatment on the microstructure of SPI films. Low concentrations of MTGase slowed down the moisture loss rate, which was consistent with the increased surface hydrophobicity of SPI films. Moreover, it was shown that MTGase-treated films had a rougher surface, more homogeneous cross-section, higher tensile strength, and lower elongation at break compared to the controls. However, the WVP was not significantly affected by the treatment and the transparency decreased. Jiang et al. [271] were able to confirm these results, indicating that the improvement of the properties of SPI films by MTGase is largely dependent on many processing parameters, for example the enzyme concentration, the pH of the film forming solution, and the temperature. Treatment at low concentration significantly increased the tensile strength whereas high concentrations of MTGase resulted in lower tensile strength. Gan et al. [272] optimized the SPI gels formed using MTGase by subsequent heat treatment with ribose which induced Maillard crosslinking. The resulting gels were free-standing, with an improved mechanical microstructure due to the higher crosslinking density achieved by the combined crosslinking techniques. MTGase was also reported to improve the functional performance of biodegradable nanocomposite materials such as montmorillonite (MMT) nanoclay intercalated with soy protein [273]. Lastly, Weng and Zheng [274] recently reported the formation of compact film network structures and increased film strength, improved water resistance properties, and better thermal stability when treating gelatin films with transglutaminase in the presence of soy protein isolate. On the whole, however, research work to produce enzymatically improved protein films for packaging materials has not yet managed to develop films having physical properties superior to those of synthetic films [29]. This is why research on SPI composite films has gained in popularity. SPI has been blended with many different polymers to achieve the desired properties [275].
According to Schmid and Hammann [208], wheat gluten film-forming solutions are well suited for enzymatic crosslinking by transglutaminase due to the high number of glutamine residues. There are also other studies on the favorable effects of wheat gluten proteins treated with TG [254,255]. However, there is only one study which discusses the effects of enzymatic treatment on wheat gluten based films. Larré et al. [276] investigated the properties of enzymatically crosslinked, deamidated gluten films. TG was able to introduce covalent bonds into the films, resulting in greater insolubility and increased ability to stretch. However, there was simultaneously reduced surface hydrophobicity.
All in all, crosslinking with the formation of ε-(γ-glutamyl)lysine bonds leads to increased tensile strength in both whey protein isolate and soy protein isolate films. The WPI films are able to increase the tensile strength from ~6 to 13 MPa, while the SPI films only show a small increase [251,257]. These differences may result from the different available side chains but they can also be caused by different enzymatic activities. Regarding the barrier properties, it is apparent that the water vapor permeation of these films almost remains at the same level [251,257].

5.1.3. Peroxydase

An early attempt to enzymatically modify the functional properties of soy protein films was undertaken by Stuchell and Krochta in 1994. No improvement in the water vapor permeability and increased brittleness were reported after treatment with horseradish peroxidase (EC 1.11.1.7). However, the enzyme increased the tensile strength and protein solubility while decreasing the elongation. As horseradish peroxidase catalyzes the oxidation of amino acid side chains, so promoting crosslinks, they concluded that oxidative crosslinking is not sufficiently specific to enhance soy protein films [84].

5.2. Composite Films and Addition of Bioactive Compounds

An important trend in recent years has been investigation of novel approaches for further enhancement of the barrier and mechanical properties of protein films. Active and intelligent packaging, or bio-packaging, are aspects which have received much attention. An active packaging is a type of material that extends the shelf life, enhances security, and maintains the product quality by changing its packaging conditions and interacting with the food [275]. In summary, these novel approaches can be of a physical, chemical, or biochemical nature. Generally, these approaches involve modification of the protein structure and/or interactions among protein molecules. Another much investigated approach has involved the creation of blended or composite films [35,277].
Proteins show good matrix as co-polymer in many blended films such as mixtures with polysaccharides. The interaction of proteins and polysaccharides is described in various recent review papers [278,279,280]. The sections below give among other compounds a small description of some of those blends, the reader is invited to look up the cited literature for more detailed information.

5.2.1. Addition of Nanocomposites

A novel upcoming trend is the preparation of WPI films with nanocomposites [29]. For instance, Li et al. [281] highlighted that more than 70% of visible light and more than 90% of UV light can be blocked by composite WPI/biodegradable titanium dioxide (TiO2) films. A year earlier, Zhou et al. [282] significantly increased the tensile properties and elastic modulus of WPI films by adding less than 1% (by wt.) of TiO2 nanoparticles. However, the moisture barrier properties were decreased and both research groups state that improvements are necessary for new composite film combinations. Zinc (Zn) is an essential micronutrient. Shi et al. [283] fabricated nanocrystalline zinc oxide (ZnO) coated with WPI, thus a nutraceutical agent within a WPI coating. Sothornvit et al. [284] fabricated WPI/nanoclay composite films which had decreased water vapor permeation. Also, the incorporation of the nanoclay Colisite 30B into WPI films exhibited gave a significant bacteriostatic effect against L. monocytogenes. Lastly, Oymaci and Altinkaya [285] reported excellent results by blending WPI-based films with zein nanoparticles (ZNP). The water vapor barrier and mechanical properties were improved, suggesting that ZNP/WPI nanocomposites films have great potential for use as biodegradable food packaging materials.
As the case for WPI-based films and coatings, nano-additives and nano-biocomposites are very promising materials for enhancing the functional properties of SPI-based films and coatings and this area has excited a lot of interest [286,287,288]. Improvements due to nanocomposites arise from the strong interactions between the matrices and the nano-reinforcements [289], involving a synergistic effect from combination of the matrix and the reinforcement [290]. In a study by Wang et al. [291], nanocrystalline titanium dioxide (TiO2) particles coated with SPI were manufactured. TiO2 has been studied extensively as the particles are cheap, photostable, and nontoxic at the recommended safe dosage, [291,292]. SPI/TiO2 combinations were found to be bactericidal to E. coli and S. aureus after the films had been irradiated with UV light at 365 nm for two hours. Another extensively studied bio-nanocomposite is montmorillonite nanoclay (MMT) [29]. Due to its unique structure and properties, this nanoclay has proven to be very effective in improving the mechanical properties of biopolymers by bearing a significant portion of the applied stress [293,294]. This has been confirmed by Kumar et al. [295,296], who reported significant enhancements in the tensile strength, elongation at break, and water vapor permeability compared to regular SPI films and by Jin and Zhong [273] who further treated the film with MTGase. Chen and Zhang (2006) and Echeverria et al. (2014) also reported similar results [297,298]. However, Kumar et al. indicated that substantial improvements in the moisture barrier properties are still required in order to match synthetic materials. Gonzalez et al. [290] developed SPI films reinforced with starch nanocrystals (SNC). The films were transparent, homogeneous, and as the amount of SNC increased the films exhibited lower affinity for water and became more resistant and less extendable. Lastly, Li et al. [299] recently succeeded in enhancing the TS, WVP, and thermal stability of soy protein films with peanut protein nanoparticles.
In addition to whey protein and soy protein films, wheat gluten films have been improved by the incorporation of nanoclays such as montmorillonite (MMT). Tunc et al. [300] prepared WG/MMT nanocomposite films. The presence of MMT led to a different protein network structure, resulting in significant reduction of the water sensitivity. The oxygen and CO2 permeabilities remained unchanged, whereas TS slightly increased. Guilherme et al. [301] reported that protein-based nanocomposites consisting of wheat gluten matrix (WG) and MMT exhibited lower WVP, confirming the results of Tunc. With the antimicrobial properties of edible films and coatings pioneering the concept of active packaging, Mascheroni et al. [302] developed an antimicrobial delivery system from film-forming solutions containing wheat gluten as matrix, MMT as structuring agent, and carvacrol as active agent. The results demonstrated there was effective retention and protection of the antimicrobial agent (carvacrol) during the processing stage. Last but not least, MMT was sandwiched between two layers of WG, forming a coating for paperboard. The oxygen barrier was around 25 times better than that of a single layer of WG. Moreover, the water vapor transmission was 6- to 8-fold lower than the uncoated paperboard [303].

5.2.2. Addition of Antimicrobial Materials

WPI films are frequently selected as model edible coating materials for incorporating various additives such as antimicrobial agents, hydrophobic materials, and antioxidants [304].
For example, the design and manufacture of edible films with encapsulated antimicrobial materials are a highly promising strategy for advancing active packaging technology, since prior studies have demonstrated the effectiveness of antimicrobial films in reducing the growth of inoculated bacteria [305]. Joerger et al. [305] and Rocha et al. [306] published detailed reviews about the potential use of antimicrobial films. Pathogen specificity may improve the antimicrobial efficacy while protecting microbes necessary for human health such as probiotic microbes and also bacteria controlling the growth of pathogenic bacteria [307]. Developing novel antimicrobial packaging materials with high specificity for targeting only pathogenic organisms and not affecting symbiotic bacteria was accomplished by Vonasek et al. [304]. Other antimicrobial agents such as essential oils, nisin, and the bioactive proteins lactoperoxidase, lactoferrin, and lysozyme have also been investigated [308,309,310,311]. Concerns about the incorporation of antimicrobial additives were pointed out by Chen [312] and Hotchkiss [313] who indicated potential negative effects on the film’s mechanical and optical properties. Ozdemir and Floros [314,315] have been working on optimizing the mechanical and optical properties of films containing preservatives with favorable results. Also, micro-encapsulated food additives in whey protein-based films were reported by Young, Sarda, and Rosenberg [316,317].
For ensuring high quality products the fabrication of films and coatings with antimicrobial properties is a suitable approach [306]. Gonzalez and Igarzabal [318] incorporated an antifungal agent and an antibacterial agent (natamycin and thymol respectively) into bilayer films produced from SPI and poly lactic acid (PLA). The SPI/PLA films had high transparency, strong adhesion between layers, suitable mechanical properties with respect to those of pure SPI films, and decreased WVP. The antibacterial additives were able to markedly inhibit the growth of mold, yeast, and two strains of bacteria. In another example, oregano or thyme essential oils reduced the counts of coliform and Pseudomonas spp. on beef patties [319]. Furthermore, soy protein films with incorporated grape seed extract exhibited inhibitory activity against L. monocytogenes [320]. With reference to other potential additives, the anti-oxidative effect of ferulic acid in soy protein films and coatings was reported by Ou et al. [321]. Further, Friesen et al. [322] recently reported that rutin and epicatechin can be used as crosslinking agents as a natural means for improving specific properties of SPI films. The addition of rutin increased the puncture strength, whereas epicatechin was found to increase the water vapor permeability.

5.2.3. Addition of Lipid Materials

Due to the high water vapor permeability of whey protein films, much research has focused on improving the barrier properties of the films by the addition of hydrophobic materials such as waxes and lipids [310,323]. Shellhammer et al. [324] investigated carnauba wax, candelia wax, milkfat fraction, and beeswax (BW, a solid and highly hydrophobic lipid). When incorporated into whey protein films, the viscoelastic milk fat and beeswax improved the water vapor permeability more than candelia or carnauba wax. Perez-Gago and Krochta [325] found that the tensile strength and elongation significantly increased when the particle size of the beeswax in whey protein films decreased. Anker et al. [326] reported a 70-fold reduction of the water vapor permeability when acetylated monoglyceride was added to whey protein films. However, lipids were shown to also negatively affect the mechanical properties of other proteins. The tensile strength and elongation can decrease at high concentrations. Furthermore, the films can become brittle and hard to handle without breaking [324] and other features such as the transparency can be affected [327]. Edible whey protein films prepared with almond and walnut oils have been reported to have increased film opacity, even though they had improved water vapor barrier properties [328]. As numerous studies on the modification of the whey protein film structure by addition of waxes have indicated, lower lipid and wax contents and smaller particle sizes result in significantly enhanced WPI films [325,329,330,331]. Nevertheless, compared to synthetic films and coatings, WPI films still only have moderate moisture barriers even with the inclusion of lipids. With regard to potential applications, whey protein films may be best for foods that need a low to moderate moisture barrier [310]. Furthermore, WPI-based films containing oregano and pimento essential oils were reported to have anti-oxidative activity for meat [332]. Other anti-oxidative compounds were reported to be vitamin E for peanuts [333,334] and ascorbic acid for apples [335].
Early studies of Gontard et al. [336] have shown that the WVP of WG films can be optimized by incorporation of lipid materials (emulsion or bilayer film). The WVP was reduced by 200-fold compared to an uncoated control film as a result of beeswax being laminated onto the WG films. Gennadios, Weller, and Testin [337] added nonpolar hydrophobic substances to the film forming solution as well. Mineral oil was able to reduce the WVP by about 25%. However, the TS also decreased. In addition, wheat gluten coatings were improved by the incorporation of lipids (beeswax, stearic and palmitic acids) by Tanada-Palmu and Grosso [338], resulting in good decay-control properties for fruit. Lastly, increased hydrophobicity was also achieved through the addition of epoxidized soybean oil (ESO) under alkaline conditions [339].

5.2.4. Other Bioactive Compounds

In 1998, Rhim et al. investigated the effect of dialdehyde starch (DAS) incorporated in SPI films. Due to water absorption by hydrophilic groups along the DAS polymer chains, small increases in the WVP were observed. However, with increased tensile strength and substantially reduced solubility in water, DAS showed the potential for increasing the resistance of SPI films, thus improving their functionality [340]. In order to improve the water barrier, SPI-based films have been combined with hydrophobic lipid materials. One approach is to add lipids to protein film forming solutions, which are then cast to prepare emulsified, bi-component or multi-component films [341]. Alternatively, molten lipids can be laminated onto protein films, resulting in bi-layer or multi-layer films [342]. Gennadios et al. [343] prepared soy protein-fatty acid bi-component films. The SPI films had notably lower WVP values than the control films. Unfortunately, the addition of fatty acids substantially reduced the film TS. Lipid oxidation may be another obstacle. This might be the reason why there is little literature on SPI-lipid composite films.
There have been many more attempts to fabricate composite SPI films and coatings, as elaborately reviewed by Koshy et al. [275]. A distinction was made between organic filler biocomposites (chitin, starch, and other polymers), organic filler bio-nanocomposites, and inorganic bio-nanocomposites. Readers are encouraged to refer to the relevant publication for further information about composite SPI films.
In another approach, Cao et al. [344] prepared composite films from SPI and gelatin and reported improved mechanical properties. The tensile strength, elongation at break, and elastic modulus were increased. Moreover, the films became more transparent and easier to handle. Guerrero et al. [53] incorporated gelatin into SPI-based films and reported similar results. The films showed higher tensile strength and similar elongation at break compared to the control. Moreover, the hydrophilicity decreased, while the UV barrier properties were maintained, suggesting a potential use of SPI films for the retardation of product oxidation induced by UV light.
An interesting approach was the addition of cysteine and gluten to soy protein films. Due to an increase in disulfide bond formation, the cysteine increased the tensile strength of soy:gluten (4:1) films. Furthermore, as gluten has large numbers of non-polar amino acids which contribute to hydrophobic interactions, its addition lowered the WVP [345].

6. Summary Tables

The properties of protein-based films are summarized in this section.

6.1. Mechanical Properties of Protein-Based Films

Table 3 lists the mechanical characteristics of different protein-based films. Readers are encouraged to refer to the respective publications for further information. See the bottom of the tables for descriptions of the footnotes.

6.2. Barrier Properties of Protein-Based Films

Table 4 gives an overview of the WVP and OP properties of different protein-based films. Readers are encouraged to refer to the respective publications for further information. See the bottom of the tables for descriptions of the footnotes.

7. Conclusions and Future Trends

All the physical, chemical, and biochemical methods that have been discussed have an effect on the films and coatings produced from the three protein sources. The resulting mechanical and barrier properties can therefore be adversely affected. However, the efficiency of the various techniques not only depends on their intensities but is also strongly influenced by the concentration and presence of other components and on the preparation methods.
Unfortunately, the respective methods often yield ambivalent results, meaning no single film or coating is appropriate for all applications for food protection. Different foods place different requirements on packaging materials. For example, the shelf life of fresh fruit or vegetables depends on the exchange of water vapor and other gas transport. On the other hand, products containing lipids, such as milk, have to be protected against oxidation by light and oxygen.
Even though there are some safety concerns about nanomaterials, nanotechnological research has recently been a most exciting development. Developments here are likely to elevate food packaging to new heights and capture the market within the next few years. Furthermore, in order to fully protect and extend the shelf life of a food product, the use of multiple components is, and will continue to be, imperative. The future challenge will therefore be to take advantage of the various technologies to form optimal protein-based films and coatings for each individual application.
With regards to the information collated for this paper, it can be concluded that the combination of nanotechnological advancements with enzymatic and physical treatments has the greatest promise.
However, it is possible to enhance the characteristics of protein-based films to such a level that they represent an interesting alternative to conventional petroleum-based films and coatings [280,347,348,349]. The general aim must remain biopolymer production on a large technical scale for industrial applications. Nevertheless, the use of protein based films and coatings is not limited to packaging material. Recent studies and reviews on protein-based adhesives [350,351,352,353,354,355,356,357,358], protectants for building materials [358], thermoplastic composites [359], bioelectronics [360], heat sealable [361] and microorganism carrier films [362] show some future prospects of these upcoming biopolymer materials.

Acknowledgments

This work was supported by the German Research Foundation (DFG) and the Technical University of Munich (TUM) in the framework of the Open Access Publishing Program. The Authors thank the DFG and TUM for their support.

Author Contributions

Joël Zink wrote the physical modification section, was in charge of the manuscript and involved in the revision and completion of the work. Tom Wyrobnik and Tobias Prinz wrote and edited the Biochemical and Chemical modification sections respectively. Markus Schmid was responsible for the overall outline and contributed to all sections as well as to the revision and completion of the manuscript. All authors approved the final version.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Multon, J. The role of packaging in preserving foodstuffs. Food Package Technol. 1996, 1, 3–23. [Google Scholar]
  2. Buchner, N. Verpackung von lebensmitteln. Lebensmitteltechnologische, Verpackungstechnische und Mikrobiologische Grundlagen; Springer: Heidelberg, Germany, 1999. [Google Scholar]
  3. Gontard, N.; Duchez, C.; Cuq, J.-L.; Guilbert, S. Edible composite films of wheat gluten and lipids: Water vapour permeability and other physical properties. Int. J. Food Sci. Technol. 1994, 29, 39–50. [Google Scholar] [CrossRef]
  4. Cuq, B.; Gontard, N.; Guilbert, S. Proteins as agricultural polymers for packaging production. Cereal Chem. 1998, 75, 1–9. [Google Scholar] [CrossRef]
  5. Khwaldia, K.; Perez, C.; Banon, S.; Desobry, S.; Hardy, J. Milk proteins for edible films and coatings. Crit. Rev. Food Sci. Nutr. 2004, 44, 239–251. [Google Scholar] [CrossRef] [PubMed]
  6. Tharanathan, R.N. Biodegradable films and composite coatings: Past, present and future. Trend. Food Sci. Technol. 2003, 14, 71–78. [Google Scholar] [CrossRef]
  7. Bucci, D.Z.; Tavares, L.B.B.; Sell, I. PHB packaging for the storage of food products. Polym. Test. 2005, 24, 564–571. [Google Scholar] [CrossRef]
  8. Verbeek, C.J.R.; van den Berg, L.E. Extrusion processing and properties of protein-based thermoplastics. Macromol. Mater. Eng. 2010, 295, 10–21. [Google Scholar] [CrossRef]
  9. Baldwin, E.A.; Hagenmaier, R.; Bai, J. Edible Coatings and Films to Improve Food Quality, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2011; pp. 291–318. [Google Scholar]
  10. Galietta, G.; di Gioia, L.; Guilbert, S.; Cuq, B. Mechanical and thermomechanical properties of films based on whey proteins as affected by plasticizer and crosslinking agents. J. Dairy Sci. 1998, 81, 3123–3130. [Google Scholar] [CrossRef]
  11. Mahmoud, R.; Savello, P.A. Mechanical properties of and water vapor transferability through whey protein films. J. Dairy Sci. 1992, 75, 942–946. [Google Scholar] [CrossRef]
  12. Maté, Juan I.; Frankel, E.N.; Krochta, J.M. Whey protein isolate edible coatings: Effect on the rancidity process of dry roasted peanuts. J. Agric. Food Chem. 1996, 44, 1736–1740. [Google Scholar]
  13. Schmid, M.; Held, J.; Hammann, F.; Schlemmer, D.; Noller, K. Effect of UV-radiation on the packaging-related properties of whey protein isolate based films and coatings. Packag. Technol. Sci. 2015, 28, 883–899. [Google Scholar] [CrossRef]
  14. Gennadios, A.; Weller, C.L.; Testin, R.F. Temperature effect on oxygen permeability of edible protein-based films. J. Food Sci. 1993, 58, 212–214. [Google Scholar] [CrossRef]
  15. Guilbert, S.; Gontard, N.; Cuq, B. Technology and applications of edible protective films. Packag. Technol. Sci. 1995, 8, 339–346. [Google Scholar] [CrossRef]
  16. vo Hong, N.; Pyka, G.; Wevers, M.; Goderis, B.; van Puyvelde, P.; Verpoest, I.; van Vuure, A.W. Processing rigid wheat gluten biocomposites for high mechanical performance. Compos. Part A: Appl. Sci. Manuf. 2015, 79, 74–81. [Google Scholar] [CrossRef]
  17. Zubeldía, F.; Ansorena, M.R.; Marcovich, N.E. Wheat gluten films obtained by compression molding. Polym. Test. 2015, 43, 68–77. [Google Scholar] [CrossRef]
  18. Sun, Q.; Sun, C.; Xiong, L. Mechanical, barrier and morphological properties of pea starch and peanut protein isolate blend films. Carbohydr. Polym. 2013, 98, 630–637. [Google Scholar] [CrossRef] [PubMed]
  19. Condés, M.C.; Añón, M.C.; Mauri, A.N. Amaranth protein films prepared with high-pressure treated proteins. J. Food Eng. 2015, 166, 38–44. [Google Scholar] [CrossRef]
  20. Bylund, G. Dairy Processing Handbook; Tetra Pak Processing Systems AB: Lund, Sweden, 2003; pp. 241–262. [Google Scholar]
  21. Fachin, L.; Viotto, W.H. Effect of pH and heat treatment of cheese whey on solubility and emulsifying properties of whey protein concentrate produced by ultrafiltration. Int. Dairy J. 2005, 15, 325–332. [Google Scholar] [CrossRef]
  22. Amin, S.; Ustunol, Z. Solubility and mechanical properties of heat-cured whey protein-based edible films compared with that of collagen and natural casings. Int. J. Dairy Technol. 2007, 60, 149–153. [Google Scholar] [CrossRef]
  23. Kim, K.M.; Weller, C.L.; Hanna, M.A.; Gennadios, A. Heat curing of soy protein films at atmospheric and sub-atmospheric conditions. J. Food Sci. 2002, 67, 708–713. [Google Scholar] [CrossRef]
  24. Reddy, N.; Yang, Y. Novel protein fibers from wheat gluten. Biomacromolecules 2007, 8, 638–643. [Google Scholar] [CrossRef] [PubMed]
  25. Krochta, J. Control of mass transfer in foods with edible coatings and films. Advance. Food Eng. 1992, 517–538. [Google Scholar]
  26. McHugh, T.; Krochta, J. Permeability properties of edible films. Edible Coat. Films Improv. Food Qual. 1994, 42, 139–187. [Google Scholar]
  27. Micard, V.; Belamri, R.; Morel, M.-H.; Guilbert, S. Properties of chemically and physically treated wheat gluten films. J. Agric. Food Chem. 2000, 48, 2948–2953. [Google Scholar] [CrossRef] [PubMed]
  28. González, A.; Strumia, M.C.; Alvarez Igarzabal, C.I. Cross-linked soy protein as material for biodegradable films: Synthesis, characterization and biodegradation. J. Food Eng. 2011, 106, 331–338. [Google Scholar] [CrossRef]
  29. Wihodo, M.; Moraru, C.I. Physical and chemical methods used to enhance the structure and mechanical properties of protein films: A review. J. Food Eng. 2013, 114, 292–302. [Google Scholar] [CrossRef]
  30. Belitz, H.-D.; Grosch, W.; Schieberle, P. Lehrbuch der Lebensmittelchemie; Springer: Berlin, Germany, 2008. [Google Scholar]
  31. Buddrus, J. Grundlagen der Organischen Chemie; Walter de Gruyter: Berlin, Germany, 2011; pp. 797–844. [Google Scholar]
  32. Means, G.E.; Feeney, R.E. Chemical modifications of proteins: A review. J. Food Biochem. 1998, 22, 399–426. [Google Scholar] [CrossRef]
  33. McHugh, T.H. Protein-lipid interactions in edible films and coatings. Nahrung. Food 2000, 44, 148–151. [Google Scholar] [CrossRef]
  34. Krochta, J.M.; Mulder-Johnston, C.D. Edible and biodegradable polymer films: Challenges and opportunities. In Food Technology USA; Institute of Food Technologists: Chicago, IL, USA, 1997. [Google Scholar]
  35. Krochta, J.M. Proteins as raw materials for films and coatings: Definitions, current status, and opportunities. Protein-Based Films Coat. 2002, 1–41. [Google Scholar]
  36. Knight, R.D. Physics for Scientists and Engineers: A Strategic Approach: With Modern Physics, 3rd ed.; Pearson: Boston, London, 2013; pp. 1213–1279. [Google Scholar]
  37. McHugh, T.H.; Avena-Bustillos, R.; Krochta, J.M. Hydrophilic edible films: Modified procedure for water vapor permeability and explanation of thickness effects. J. Food Sci. 1993, 58, 899–903. [Google Scholar] [CrossRef]
  38. Committee, F. Test Method for Oxygen Gas Transmission Rate through Plastic Film and Sheeting Using Various Sensors; ASTM International: West Conshohocken, PA, USA, 2013. [Google Scholar]
  39. Guilbert, S.; Gontard, N.; Morel, M.; Chalier, P.; Micard, V.; Redl, A. Formation and properties of wheat gluten films and coatings. In Protein-Based Films and Coatings; CRC Press: New York, USA, 2002; pp. 69–122. [Google Scholar]
  40. Cuq, B. Formation and properties of fish myofibrillar protein films and coatings. In Protein-Based Films and Coatings; CRC Press: New York, USA, 2002; pp. 213–232. [Google Scholar]
  41. Schmid, M.; Müller, K.; Sängerlaub, S.; Stäbler, A.; Starck, V.; Ecker, F.; Noller, K. Mechanical and barrier properties of thermoplastic whey protein isolate/ethylene vinyl acetate blends. J. Appl. Polym. Sci. 2014, 131. [Google Scholar] [CrossRef]
  42. Schmid, M.; Hammann, F.; Winkler, H. Technofunctional properties of films made from ethylene vinyl acetate/whey protein isolate compounds. Packag. Technol. Sci. 2014, 27, 521–533. [Google Scholar] [CrossRef]
  43. Gällstedt, M.; Mattozzi, A.; Johansson, E.; Hedenqvist, M.S. Transport and tensile properties of compression-molded wheat gluten films. Biomacromolecules 2004, 5, 2020–2028. [Google Scholar] [CrossRef] [PubMed]
  44. Belyamani, I.; Prochazka, F.; Assezat, G. Production and characterization of sodium caseinate edible films made by blown-film extrusion. J. Food Eng. 2014, 121, 39–47. [Google Scholar] [CrossRef]
  45. De Wit, J.; Fox, P. Functional properties of whey proteins. In Developments in Dairy Chemistry—4. Functional milk proteins; Elsevier Applied Science: New York, USA, 1989; pp. 285–321. [Google Scholar]
  46. Morr, C.; Ha, E. Whey protein concentrates and isolates: Processing and functional properties. Crit. Rev. Food Sci. Nutr. 1993, 33, 431–476. [Google Scholar] [CrossRef] [PubMed]
  47. Kilara, A.; Vaghela, M.; Yada, R. Whey proteins. In Proteins in Food Processing; Woodhead Publishing Limited: Cambridge, England, 2004; pp. 72–99. [Google Scholar]
  48. Dybing, S.; Smith, D. Relation of chemistry and processing precedures to whey protein functionality: A review. Cult. Dairy Prod. J. 1991. [Google Scholar] [CrossRef]
  49. Brunner, J. Milk proteins. In Food Proteins; Avi Publishers Inc: Westport, USA, 1977; pp. 175–208. [Google Scholar]
  50. Pérez-Gago, M.B.; Krochta, J.M. Formation and properties of whey protein films and coatings. In Protein-Based Films and Coatings; CRC Press: New York, USA, 2002; Volume 6, pp. 159–180. [Google Scholar]
  51. Alexandrescu, A.T.; Evans, P.A.; Pitkeathly, M.; Baum, J.; Dobson, C.M. Structure and dynamics of the acid-denatured molten globule state of α-lactalbumin: A two-dimensional NMR study. Biochemistry 1993, 32, 1707–1718. [Google Scholar] [CrossRef] [PubMed]
  52. Brown, J.R. Serum albumin: Amino acid sequence. In Albumin Structure, Function and Uses; Pergamon Press Inc.: Oxford, England, 1977; pp. 27–51. [Google Scholar]
  53. Guerrero, P.; Stefani, P.; Ruseckaite, R.; de la Caba, K. Functional properties of films based on soy protein isolate and gelatin processed by compression molding. J. Food Eng. 2011, 105, 65–72. [Google Scholar] [CrossRef]
  54. Cho, S.Y.; Rhee, C. Mechanical properties and water vapor permeability of edible films made from fractionated soy proteins with ultrafiltration. LWT-Food Sci. Technol. 2004, 37, 833–839. [Google Scholar] [CrossRef]
  55. Thanh, V.H.; Shibasaki, K. Major proteins of soybean seeds. A straightforward fractionation and their characterization. J. Agric. Food Chem. 1976, 24, 1117–1121. [Google Scholar] [CrossRef] [PubMed]
  56. Kinsella, J. Functional properties of soy proteins. J. Am. Oil Chem.’ Soc. 1979, 56, 242–258. [Google Scholar] [CrossRef]
  57. Nisperos-Carriedo, M.O.; Krochta, J.M.; Baldwin, E.A. Edible Coatings and Films to Improve Food Quality; Technomic Publishing Company: Lancaster, PA, USA, 1994. [Google Scholar]
  58. Kunte, L.; Gennadios, A.; Cuppett, S.; Hanna, M.; Weller, C.L. Cast films from soy protein isolates and fractions 1. Cereal Chem. 1997, 74, 115–118. [Google Scholar] [CrossRef]
  59. Iwabuchi, S.; Yamauchi, F. Electrophoretic analysis of whey proteins present in soybean globulin fractions. J. Agric. Food Chem. 1987, 35, 205–209. [Google Scholar] [CrossRef]
  60. Sabato, S.F.; Ouattara, B.; Yu, H.; D’Aprano, G.; Le Tien, C.; Mateescu, M.A.; Lacroix, M. Mechanical and barrier properties of cross-linked soy and whey protein based films. J. Agric. Food Chem. 2001, 49, 1397–1403. [Google Scholar] [CrossRef] [PubMed]
  61. Hoseney, R.C. Principles of Cereal Science and Technology; American Association of Cereal Chemists: St.-Paul, MN, USA, 1994; pp. 23–52. [Google Scholar]
  62. Belitz, H.-D.; Grosch, W.; Schieberle, P. Lehrbuch der Lebensmittelchemie—5; Springer: Berlin, Germeny, 2001; p. 1059. [Google Scholar]
  63. Dendy, D.A.V.; Dobraszczyk, B.J. Cereals and Cereal Products: Chemistry and Technology; Aspen Publishing: Gaithersburg, Maryland, 2001; p. 429. [Google Scholar]
  64. Nicorescu, I.; Loisel, C.; Vial, C.; Riaublanc, A.; Djelveh, G.; Cuvelier, G.; Legrand, J. Combined effect of dynamic heat treatment and ionic strength on the properties of whey protein foams–part II. Food Res. Int. 2008, 41, 980–988. [Google Scholar] [CrossRef]
  65. Nicolai, T.; Britten, M.; Schmitt, C. β-lactoglobulin and WPI aggregates: Formation, structure and applications. Food Hydrocoll. 2011, 25, 1945–1962. [Google Scholar] [CrossRef]
  66. Soroka, W. Fundamentals of Packaging Technology, 4th ed.; Institute of Packaging Professionals: Naperville III, Lancaster, CA, USA, 2009. [Google Scholar]
  67. Parris, N.; Purcell, J.M.; Ptashkin, S.M. Thermal denaturation of whey proteins in skim milk. J. Agric. Food Chem. 1991, 39, 2167–2170. [Google Scholar] [CrossRef]
  68. Singh, H.; MacRitchie, F. Changes in proteins induced by heating gluten dispersions at high temperature. J. Cereal Sci. 2004, 39, 297–301. [Google Scholar] [CrossRef]
  69. Lakemond, C.M.; Jongh, H.H.d.; Hessing, M.; Gruppen, H.; Voragen, A.G. Heat denaturation of soy glycinin: Influence of ph and ionic strength on molecular structure. J. Agric. Food Chem. 2000, 48, 1991–1995. [Google Scholar] [CrossRef] [PubMed]
  70. Renkema, J.M.; Lakemond, C.M.; Jongh, H.H.d.; Gruppen, H.; van Vliet, T. The effect of ph on heat denaturation and gel forming properties of soy proteins. J. Biotechnol. 2000, 79, 223–230. [Google Scholar] [CrossRef]
  71. Guo, G.; Zhang, C.; Du, Z.; Zou, W.; Xiang, A.; Li, H. Processing and properties of phthalic anhydride modified soy protein/glycerol plasticized soy protein composite films. J. Appl. Polym. Sci. 2015, 132. [Google Scholar] [CrossRef]
  72. Furukawa, T.; Ohta, S.; Yamamoto, A. Texture-structure relationships in heat-induced soy protein gels. J. Texture Stud. 1980, 10, 333–346. [Google Scholar] [CrossRef]
  73. Miller, K.S.; Chiang, M.T.; Krochta, J.M. Heat curing of whey protein films. J. Food Chem. 1997, 62, 1189–1193. [Google Scholar] [CrossRef]
  74. Rangavajhyala, N.; Ghorpade, V.; Hanna, M. Solubility and molecular properties of heat-cured soy protein films. J. Agric. Food Chem. 1997, 45, 4204–4208. [Google Scholar] [CrossRef]
  75. Perez-Gago, M.B.; Nadaud, P.; Krochta, J.M. Water vapor permeability, solubility, and tensile properties of heat-denatured versus native whey protein films. J. Food Sci. 1999, 64, 1034–1037. [Google Scholar] [CrossRef]
  76. Roy, S.; Weller, C.L.; Gennadios, A.; Zeece, M.G.; Testin, R.F. Physical and molecular properties of wheat gluten films cast from heated film-forming solutions. J. Food Sci. 1999, 64, 57–60. [Google Scholar] [CrossRef]
  77. Vachon, C.; Yu, H.-L.; Yefsah, R.; Alain, R.; St-Gelais, D.; Lacroix, M. Mechanical and structural properties of milk protein edible films cross-linked by heating and γ-irradiation. J. Agric. Food Chem. 2000, 48, 3202–3209. [Google Scholar] [CrossRef] [PubMed]
  78. Rhim, J.-W.; Gennadios, A.; Handa, A.; Weller, C.L.; Hanna, M.A. Solubility, tensile, and color properties of modified soy protein isolate films. J. Agric. Food Chem. 2000, 48, 4937–4941. [Google Scholar] [CrossRef] [PubMed]
  79. Perez-Gago, M.B.; Krochta, J.M. Denaturation time and temperature effects on solubility, tensile properties, and oxygen permeability of whey protein edible films. J. Food Chem. 2001, 66, 705–710. [Google Scholar] [CrossRef]
  80. Simelane, S.; Ustunol, Z. Mechanical properties of heat-cured whey protein-based edible films compared with collagen casings under sausage manufacturing conditions. J. Food Chem. 2005, 70, E131–E134. [Google Scholar] [CrossRef]
  81. Hong, S.-I.; Krochta, J.M. Oxygen barrier performance of whey-protein-coated plastic films as affected by temperature, relative humidity, base film and protein type. J. Food Eng. 2006, 77, 739–745. [Google Scholar] [CrossRef]
  82. Sothornvit, R.; Olsen, C.W.; McHugh, T.H.; Krochta, J.M. Tensile properties of compression-molded whey protein sheets: Determination of molding condition and glycerol-content effects and comparison with solution-cast films. J. Food Eng. 2007, 78, 855–860. [Google Scholar] [CrossRef]
  83. Barreto, P.L.M.; Pires, A.T.N.; Soldi, V. Thermal degradation of edible films based on milk proteins and gelatin in inert atmosphere. Polym. Degrad. Stab. 2003, 79, 147–152. [Google Scholar] [CrossRef]
  84. Stuchell, Y.M.; Krochta, J.M. Enzymatic treatments and thermal effects on edible soy protein films. J. Food Chem. 1994, 59, 1332–1337. [Google Scholar] [CrossRef]
  85. Thomas, C.R.; Geer, D. Effects of shear on proteins in solution. Biotechnol. Lett. 2011, 33, 443–456. [Google Scholar] [CrossRef] [PubMed]
  86. Steventon, A.J. Thermal Aggregation of Whey Proteins. Doctoral Thesis, University of Cambridge, Cambridge, UK, 1993. [Google Scholar]
  87. Taylor, S.M.; Fryer, P.J. The effect of temperature/shear history on the thermal gelation of whey protein concentrates. Food Hydrocoll. 1994, 8, 45–61. [Google Scholar] [CrossRef]
  88. Ker, Y.C.; Toledo, R.T. Influence of shear treatments on consistency and gelling properties of whey protein isolate suspensions. J. Food Chem. 1992, 57, 82–85. [Google Scholar] [CrossRef]
  89. Simmons, M.J.H.; Jayaraman, P.; Fryer, P.J. The effect of temperature and shear rate upon the aggregation of whey protein and its implications for milk fouling. J. Food Eng. 2007, 79, 517–528. [Google Scholar] [CrossRef]
  90. Cheftel, J.C.; Kitagawa, M.; Quéguiner, C. New protein texturization processes by extrusion cooking at high moisture levels. Food Rev. Int. 1992, 8, 235–275. [Google Scholar] [CrossRef]
  91. Wolz, M.; Mersch, E.; Kulozik, U. Thermal aggregation of whey proteins under shear stress. Food Hydrocoll. 2016. [Google Scholar] [CrossRef]
  92. Fang, Y.; Zhang, B.; Wei, Y.; Li, S. Effects of specific mechanical energy on soy protein aggregation during extrusion process studied by size exclusion chromatography coupled with multi-angle laser light scattering. J. Food Eng. 2013, 115, 220–225. [Google Scholar] [CrossRef]
  93. Pommet, M.; Redl, A.; Morel, M.-H.; Domenek, S.; Guilbert, S. Thermoplastic processing of protein-based bioplastics: Chemical engineering aspects of mixing, extrusion and hot molding. Macromol. Symp. 2003, 197, 207–218. [Google Scholar] [CrossRef]
  94. Galazka, V.B.; Dickinson, E.; Ledward, D.A. Emulsifying behaviour of 11S globulin vicia faba in mixtures with sulphated polysaccharides: Comparison of thermal and high-pressure treatments. Food Hydrocoll. 1999, 13, 425–435. [Google Scholar] [CrossRef]
  95. Galazka, V. Influence of high pressure on interactions of 11S globulin vicia faba with I-carrageenan in bulk solution and at interfaces. Food Hydrocoll. 2000, 14, 551–560. [Google Scholar] [CrossRef]
  96. Patras, A.; Brunton, N.; Da Pieve, S.; Butler, F.; Downey, G. Effect of thermal and high pressure processing on antioxidant activity and instrumental colour of tomato and carrot purées. Innov. Food Sci. Emerg. Technol. 2009, 10, 16–22. [Google Scholar] [CrossRef]
  97. Lorido, L.; Estévez, M.; Ventanas, J.; Ventanas, S. Comparative study between serrano and iberian dry-cured hams in relation to the application of high hydrostatic pressure and temporal sensory perceptions. LWT Food Sci. Technol. 2015, 64, 1234–1242. [Google Scholar] [CrossRef]
  98. Heremans, K. High pressure effects on proteins and other biomolecules. Annu. Rev. Biophys. Bioeng. 1982, 11, 1–21. [Google Scholar] [CrossRef] [PubMed]
  99. Mozhaev, V.V.; Heremans, K.; Frank, J.; Masson, P.; Balny, C. High pressure effects on protein structure and function. Proteins: Struct. Funct. Genet. 1996, 24, 81–91. [Google Scholar] [CrossRef]
  100. Kauzmann, W. Some Factors in the Interpretation of Protein Denaturation. In Advances in Protein Chemistry 14; Anfinsen, C.B., Ed.; Academic: New York, USA, 1959; Volume 14, pp. 1–63. [Google Scholar]
  101. Gekko, K.; Hasegawa, Y. Compressibility-structure relationship of globular proteins. Biochemistry 1986, 25, 6563–6571. [Google Scholar] [CrossRef] [PubMed]
  102. Wong, P.T.T.; Heremans, K. Pressure effects on protein secondary structure and hydrogen deuterium exchange in chymotrypsinogen: A Fourier transform infrared spectroscopic study. BBA Protein Struct. Mol. Enzymol. 1988, 956, 1–9. [Google Scholar] [CrossRef]
  103. Tauscher, B. Pasteurization of food by hydrostatic high pressure: Chemical aspects. Zeitschrift für Lebensmittel-Untersuchung und -Forschung 1995, 200, 3–13. [Google Scholar] [CrossRef] [PubMed]
  104. Grinberg, V.Y.; Haertlé, T. Reducer driven baric denaturation and oligomerisation of whey proteins. J. Biotechnol. 2000, 79, 205–209. [Google Scholar] [CrossRef]
  105. Ludikhuyze, L.; van Loey, A.; Indrawati; Smout, C.; Hendrickx, M. Effects of combined pressure and temperature on enzymes related to quality of fruits and vegetables: From kinetic information to process engineering aspects. Crit. Rev. Food Sci. Nutr. 2003, 43, 527–586. [Google Scholar] [CrossRef] [PubMed]
  106. Balny, C. High pressure and biotechnology. In Proceedings of the First European Seminar on High Pressure and Biotechnology, a Joint Meeting with the Fifth Symposium on High Pressure and Food Science, la grande motte, France, 13–17 September 1992; John Libbey Eurotext: Montrouge, France, 1992; p. 565. [Google Scholar]
  107. Balny, C.; Masson, P. Effects of high pressure on proteins. Food Rev. Int. 2009, 9, 611–628. [Google Scholar] [CrossRef]
  108. Balny, C.; Masson, P.; Heremans, K. High pressure effects on biological macromolecules: From structural changes to alteration of cellular processes. Biochim. Biophys. Acta 2002, 1595, 3–10. [Google Scholar] [CrossRef]
  109. Boonyaratanakornkit, B.B.; Park, C.B.; Clark, D.S. Pressure effects on intra and intermolecular interactions within proteins. BBA Protein Struct. Mol. Enzymol. 2002, 1595, 235–249. [Google Scholar] [CrossRef]
  110. Baier, A.K.; Knorr, D. Influence of high sostatic pressure on structural and functional characteristics of potato protein. Food Res. Int. 2015. [Google Scholar] [CrossRef]
  111. Molinaro, S.; Cruz-Romero, M.; Sensidoni, A.; Morris, M.; Lagazio, C.; Kerry, J.P. Combination of high-pressure treatment, mild heating and holding time effects as a means of improving the barrier properties of gelatin-based packaging films using response surface modeling. Innov. Food Sci. Emerg. Technol. 2015, 30, 15–23. [Google Scholar] [CrossRef]
  112. Fetzer, R.W.; Ramachandran, K.S. Protein film process. Patent No. US4133901 A, 9 January 1979. [Google Scholar]
  113. Dumay, E.M.; Kalichevsky, M.T.; Cheftel, J.C. High-pressure unfolding and aggregation of β-lactoglobulin and the baroprotective effects of sucrose. J. Agric. Food Chem. 1994, 42, 1861–1868. [Google Scholar] [CrossRef]
  114. Dumay, E.M.; Kalichevsky, M.T.; Cheftel, J.C. Characteristics of pressure-induced gels of β-lactoglobulin at various times after pressure release. LWT Food Sci. Technol. 1998, 31, 10–19. [Google Scholar] [CrossRef]
  115. Olsen, K.; Ipsen, R.; Otte, J.; Skibsted, L.H. Effect of high pressure on aggregation and thermal gelation of β-lactoglobulin. Milchwissenschaft 1999, 54, 543–546. [Google Scholar]
  116. Tedford, L.-A.; Schaschke, C.J. Induced structural change to β-lactoglobulin by combined pressure and temperature. Biochem. Eng. J. 2000, 5, 73–76. [Google Scholar] [CrossRef]
  117. Liu, L.; Kerry, J.F.; Kerry, J.P. Selection of optimum extrusion technology parameters in the manufacture of edible/biodegradable packaging films derived from food-based polymers. J. Food Agric. Environ. 2005, 3. [Google Scholar]
  118. Bouaouina, H.; Desrumaux, A.; Loisel, C.; Legrand, J. Functional properties of whey proteins as affected by dynamic high-pressure treatment. Int. Dairy J. 2006, 16, 275–284. [Google Scholar] [CrossRef]
  119. Lee, S.-H.; Lefèvre, T.; Subirade, M.; Paquin, P. Changes and roles of secondary structures of whey protein for the formation of protein membrane at soy oil/water interface under high-pressure homogenization. J. Agric. Food Chem. 2007, 55, 10924–10931. [Google Scholar] [CrossRef] [PubMed]
  120. Kanno, C.; Mu, T.-H.; Hagiwara, T.; Ametani, M.; Azuma, N. Gel formation from industrial milk whey proteins under hydrostatic pressure: Effect of hydrostatic pressure and protein concentration. J. Agric. Food Chem. 1998, 46, 417–424. [Google Scholar] [CrossRef] [PubMed]
  121. Famelart, M.-H.; Chapron, L.; Piot, M.; Brulé, G.; Durier, C. High pressure-induced gel formation of milk and whey concentrates. J. Food Eng. 1998, 36, 149–164. [Google Scholar]
  122. Van Camp, J.; Feys, G.; Huyghebaert, A. High pressure induced gel formation of haemoglobin and whey proteins at elevated temperatures. LWT Food Sci. Technol. 1996, 29, 49–57. [Google Scholar] [CrossRef]
  123. Phillips, L.G.; German, J.B.; O’Neill, T.E.; Foegeding, E.A.; Harwalkar, V.R.; Kilara, A.; Lewis, B.A.; Mangino, M.E.; Morr, C.V.; Regenstein, J.M.; et al. Standardized procedure for measuring foaming properties of three proteins, a collaborative study. J. Food Chem. 1990, 55, 1441–1444. [Google Scholar] [CrossRef]
  124. Dumoulin, M.; Ozawa, S.; Hayashi, R. Textural properties of pressure-induced gels of food proteins obtained under different temperatures including subzero. J. Food Chem. 1998, 63, 92–95. [Google Scholar] [CrossRef]
  125. Molina, E.; Defaye, A.B.; Ledward, D.A. Soy protein pressure-induced gels. Food Hydrocoll. 2002, 16, 625–632. [Google Scholar] [CrossRef]
  126. Speroni, F.; Beaumal, V.; Lamballerie, M.D.; Anton, M.; Añón, M.C.; Puppo, M.C. Gelation of soybean proteins induced by sequential high-pressure and thermal treatments. Food Hydrocoll. 2009, 23, 1433–1442. [Google Scholar] [CrossRef]
  127. Apichartsrangkoon, A.; Ledward, D.A.; Bell, A.E.; Brennan, J.G. Physicochemical properties of high pressure treated wheat gluten. Food Chem. 1998, 63, 215–220. [Google Scholar] [CrossRef]
  128. Molina, E.; Papadopoulou, A.; Ledward, D.A. Emulsifying properties of high pressure treated soy protein isolate and 7S and 11S globulins. Food Hydrocoll. 2001, 15, 263–269. [Google Scholar] [CrossRef]
  129. Zhang, H.; Li, L.; Tatsumi, E.; Kotwal, S. Influence of high pressure on conformational changes of soybean glycinin. Innov. Food Sci. Emerg. Technol. 2003, 4, 269–275. [Google Scholar] [CrossRef]
  130. Alvarez, P.A.; Ramaswamy, H.S.; Ismail, A.A. High pressure gelation of soy proteins: Effect of concentration, ph and additives. J. Food Eng. 2008, 88, 331–340. [Google Scholar] [CrossRef]
  131. Subirade, M.; Kelly, I.; Guéguen, J.; Pézolet, M. Molecular basis of film formation from a soybean protein: Comparison between the conformation of glycinin in aqueous solution and in films. Int. J. Biol. Macromol. 1998, 23, 241–249. [Google Scholar] [CrossRef]
  132. Kieffer, R.; Schurer, F.; Köhler, P.; Wieser, H. Effect of hydrostatic pressure and temperature on the chemical and functional properties of wheat gluten: Studies on gluten, gliadin and glutenin. J. Cereal Sci. 2007, 45, 285–292. [Google Scholar] [CrossRef]
  133. Apichartsrangkoon, A.; Bell, A.E.; Ledward, D.A.; Schofield, J.D. Dynamic viscoelastic behavior of high-pressure-treated wheat gluten. Cereal Chem. 1999, 76, 777–782. [Google Scholar] [CrossRef]
  134. Apichartsrangkoon, A.; Ledward, D.A. Dynamic viscoelastic behaviour of high pressure treated gluten–soy mixtures. Food Chem. 2002, 77, 317–323. [Google Scholar] [CrossRef]
  135. Corso, J.F. Bone-conduction thresholds for sonic and ultrasonic frequencies. J. Acoust. Soc. Am. 1963, 35, 1738. [Google Scholar] [CrossRef]
  136. Chemat, F.; Zill-e-Huma; Khan, M.K. Applications of ultrasound in food technology: Processing, preservation and extraction. Ultrason. Sonochem. 2011, 18, 813–835. [Google Scholar] [CrossRef] [PubMed]
  137. Van Eldik, R.; Hubbard, C.D. Chemistry under Extreme and Non-Classical Conditions; John Wiley and Sons: New York, NY, USA, 1996; pp. 317–428. [Google Scholar]
  138. McClements, D.J. Advances in the application of ultrasound in food analysis and processing. Trend. Food Sci. Technol. 1995, 6, 293–299. [Google Scholar] [CrossRef]
  139. Povey, M.J.W.; Mason, T.J. Ultrasound in Food Processing; Springer: New York, NY, USA, 1998; pp. 30–65. [Google Scholar]
  140. Gülseren, I.; Güzey, D.; Bruce, B.D.; Weiss, J. Structural and functional changes in ultrasonicated bovine serum albumin solutions. Ultrason. Sonochem. 2007, 14, 173–183. [Google Scholar] [CrossRef] [PubMed]
  141. Brennan, J.G. Food Processing Handbook; Wiley-VCH: Weinheim, Germany, 2006; pp. 513–558. [Google Scholar]
  142. Mason, T.J. Practical sonochemistry: User/s Guide to Applications in Chemistry and Chemical Engineering; Ellis Horwood: New York, NY, USA, 1991; p. 186. [Google Scholar]
  143. Piyasena, P.; Mohareb, E.; McKellar, R.C. Inactivation of microbes using ultrasound: A review. Int. J. Food Microbiol. 2003, 87, 207–216. [Google Scholar] [CrossRef]
  144. Riezs, P.; Kondo, T. Free radical formation induced by ultrasound and its biological implications. Free Radic. Biol. Med. 1992, 13, 247–270. [Google Scholar]
  145. Petrier, C.; Jeunet, A.; Luche, J.L.; Reverdy, G. Unexpected frequency effects on the rate of oxidative processes induced by ultrasound. J. Am. Chem. Soc. 1992, 114, 3148–3150. [Google Scholar] [CrossRef]
  146. Mead, E.L.; Sutherland, R.G.; Verrall, R.E. The effect of ultrasound on water in the presence of dissolved gases. Can. J. Chem. 1976, 54, 1114–1120. [Google Scholar] [CrossRef]
  147. Suslick, K.S.; Casadonte, D.J.; Green, M.L.H.; Thompson, M.E. Effects of high intensity ultrasound on inorganic solids. Ultrasonics 1987, 25, 56–59. [Google Scholar] [CrossRef]
  148. Sinisterra, J.V. Application of ultrasound to biotechnology: An overview. Ultrasonics 1992, 30, 180–185. [Google Scholar] [CrossRef]
  149. Coleman, S.; Roy, S. Effect of ultrasound on mass transfer during electrodeposition for electrodes separated by a narrow gap. Chem. Eng. Sci. 2014, 113, 35–44. [Google Scholar] [CrossRef] [Green Version]
  150. Kadam, D.M.; Thunga, M.; Wang, S.; Kessler, M.R.; Grewell, D.; Lamsal, B.; Yu, C. Preparation and characterization of whey protein isolate films reinforced with porous silica coated titania nanoparticles. J. Food Eng. 2013, 117, 133–140. [Google Scholar] [CrossRef]
  151. Banerjee, R.; Chen, H.; Wu, J. Milk protein-based edible film mechanical strength changes due to ultrasound process. J. Food Chem. 1996, 61, 824–828. [Google Scholar] [CrossRef]
  152. Chen, H.; Banerjee, R.; Wu, J.R. Strengths of thin films derived from whey proteins. Am. Soc. Agric. Eng. 1993, 93, 6528. [Google Scholar]
  153. Rodriguez-Turienzo, L.; Cobos, A.; Diaz, O. Effects of edible coatings based on ultrasound-treated whey proteins in quality attributes of frozen atlantic salmon (salmo salar). Innov. Food Sci. Emerg. Technol. 2012, 14, 92–98. [Google Scholar] [CrossRef]
  154. Debeaufort, F.; Quezada-Gallo, J.A.; Voilley, A. Edible films and coatings: Tomorrow’s packagings: A review. Crit. Rev. Food Sci. Nutr. 1998, 38, 299–313. [Google Scholar] [CrossRef] [PubMed]
  155. Guzey, D.; Gulseren, I.; Bruce, B.; Weiss, J. Interfacial properties and structural conformation of thermosonicated bovine serum albumin. Food Hydrocoll. 2006, 20, 669–677. [Google Scholar] [CrossRef]
  156. Jambrak, A.R.; Lelas, V.; Mason, T.J.; Krešić, G.; Badanjak, M. Physical properties of ultrasound treated soy proteins. J. Food Eng. 2009, 93, 386–393. [Google Scholar] [CrossRef]
  157. Wang, Z.; Sun, X.-X.; Lian, Z.-X.; Wang, X.-X.; Zhou, J.; Ma, Z.-S. The effects of ultrasonic/microwave assisted treatment on the properties of soy protein isolate/microcrystalline wheat-bran cellulose film. J. Food Eng. 2013, 114, 183–191. [Google Scholar] [CrossRef]
  158. Wang, Z.; Zhou, J.; Wang, X.-X.; Zhang, N.; Sun, X.-X.; Ma, Z.-S. The effects of ultrasonic/microwave assisted treatment on the water vapor barrier properties of soybean protein isolate-based oleic acid/stearic acid blend edible films. Food Hydrocoll. 2014, 35, 51–58. [Google Scholar] [CrossRef]
  159. Hu, H.; Li-Chan, E.C.Y.; Wan, L.; Tian, M.; Pan, S. The effect of high intensity ultrasonic pre-treatment on the properties of soybean protein isolate gel induced by calcium sulfate. Food Hydrocoll. 2013, 32, 303–311. [Google Scholar] [CrossRef]
  160. Singh, N.K.; Donovan, G.R.; Batey, I.L.; MacRitchie, F. Use of sonication and size-exclusion high-performance liquid use of sonication and size-exclusion high-performance liquid chromatography in the study of wheat flour proteins. II. Relative quantity of glutenin as a measure of breadmaking quality. Cereal Chem. 1990, 67, 161–170. [Google Scholar]
  161. Marcuzzo, E.; Peressini, D.; Debeaufort, F.; Sensidoni, A. Effect of ultrasound treatment on properties of gluten-based film. Innov. Food Sci. Emerg. Technol. 2010, 11, 451–457. [Google Scholar] [CrossRef]
  162. Gennadios, A.; Rhim, J.-W.; Handa, A.; Weller, C.L.; Hanna, M.A. Ultraviolet radiation affects physical and molecular properties of soy protein films. J. Food Chem. 1998, 63, 225–228. [Google Scholar] [CrossRef]
  163. Rhim, J.-W.; Gennadios, A.; Fu, D.; Weller, C.L.; Hanna, M.A. Properties of ultraviolet irradiated protein films. LWT Food Sci. Technol. 1999, 32, 129–133. [Google Scholar] [CrossRef]
  164. Brault, D.; D’Aprano, G.; Lacroix, M. Formation of free-standing sterilized edible films from irradiated caseinates. J. Agric. Food Chem. 1997, 45, 2964–2969. [Google Scholar] [CrossRef]
  165. Garrison, W.M. Reaction mechanisms in the radiolysis of peptides, polypeptides, and proteins. Chem. Rev. 1987, 87, 381–398. [Google Scholar] [CrossRef]
  166. Puchala, M.; Schuessler, H. Oxygen effect in the radiolysis of proteins. IV. Myoglobin. Int. J. Pept. Protein Res. 1995, 46, 326–332. [Google Scholar] [CrossRef] [PubMed]
  167. Filali-Mouhim, A.; Audette, M.; St-Louis, M.; Thauvette, L.; Denoroy, L.; Penin, F.; Chen, X.; Rouleau, N.; Lecaer, J.; Rossier, J.; et al. Lysozyme fragmentation induced by γ-radiolysis. Int. J. Radiat. Biol. 1997, 72, 63–70. [Google Scholar] [CrossRef] [PubMed]
  168. Ressouany, M.; Vachon, C.; Lacroix, M. Irradiation dose and calcium effect on the mechanical properties of cross-linked caseinate films. J. Agric. Food Chem. 1998, 46, 1618–1623. [Google Scholar] [CrossRef]
  169. Cho, Y.; Bin Song, K. Effect of γ-irradiation on the molecular properties of bovine serum albumin and β-lactoglobulin. J. Biochem. Mol. Biol. 2000, 33, 133–137. [Google Scholar]
  170. Molins, R.A. Food Irradiation: Principles and Applications. Wiley: New York, NY, USA, 2001; pp. 37–68. [Google Scholar]
  171. Davies, M.J. Singlet oxygen-mediated damage to proteins and its consequences. Biochem. Biophys. Res. Commun. 2003, 305, 761–770. [Google Scholar] [CrossRef]
  172. Wondraczek, H.; Kotiaho, A.; Fardim, P.; Heinze, T. Photoactive polysaccharides. Carbohydr. Polym. 2011, 83, 1048–1061. [Google Scholar] [CrossRef]
  173. Ustunol, Z.; Mert, B. Water solubility, mechanical, barrier, and thermal properties of cross-linked whey protein isolate-based films. J. Food Chem. 2004, 69, FEP129–FEP133. [Google Scholar] [CrossRef]
  174. Vaz, C.M.; Graaf, L.A.d.; Reis, R.L.; Cunha, A.M. Effect of crosslinking, thermal treatment and uv irradiation on the mechanical properties and in vitro degradation behavior of several natural proteins aimed to be used in the biomedical field. J. Mater. Sci. Mater. Med. 2003, 14, 789–796. [Google Scholar] [CrossRef] [PubMed]
  175. Cieśla, K.; Salmieri, S.; Lacroix, M. γ-irradiation influence on the structure and properties of calcium caseinate-whey protein isolate based films. Part 1. Radiation effect on the structure of proteins gels and films. J. Agric. Food Chem. 2006, 54, 6374–6384. [Google Scholar] [CrossRef] [PubMed]
  176. Ouattara, B.; Canh, L.T.; Vachon, C.; Mateescu, M.A.; Lacroix, M. Use of γ-irradiation cross-linking to improve the water vapor permeability and the chemical stability of milk protein films. Radiat. Phys. Chem. 2002, 63, 821–825. [Google Scholar] [CrossRef]
  177. Lacroix, M.; Le, T.C.; Ouattara, B.; Yu, H.; Letendre, M.; Sabato, S.F.; Mateescu, M.A.; Patterson, G. Use of γ-irradiation to produce films from whey, casein and soya proteins: Structure and functionals characteristics. Radiat. Phys. Chem. 2002, 63, 827–832. [Google Scholar] [CrossRef]
  178. Köksel, H.; Sapirstein, H.D.; Çelik, S.; Bushuk, W. Effects of γ-irradiation of wheat on gluten proteins. J. Cereal Sci. 1998, 28, 243–250. [Google Scholar] [CrossRef]
  179. Lee, M.; Lee, S.; Song, K.B. Effect of γ-irradiation on the physicochemical properties of soy protein isolate films. Radiat. Phys. Chem. 2005, 72, 35–40. [Google Scholar] [CrossRef]
  180. Tolstoguzov, V.B. Some physico-chemical aspects of protein processing in foods. Multicomponent gels. Food Hydrocoll. 1995, 9, 317–332. [Google Scholar] [CrossRef]
  181. Bengoechea, C.; Arrachid, A.; Guerrero, A.; Hill, S.E.; Mitchell, J.R. Relationship between the glass transition temperature and the melt flow behavior for gluten, casein and soya. J. Cereal Sci. 2007, 45, 275–284. [Google Scholar] [CrossRef]
  182. Sothornvit, R.; Krochta, J.M. Plasticizer effect on mechanical properties of β-lactoglobulin films. J. Food Eng. 2001, 50, 149–155. [Google Scholar] [CrossRef]
  183. McHugh, T.H.; Krochta, J.M. Sorbitol- vs. glycerol-plasticized whey protein edible films: Integrated oxygen permeability and tensile property evaluation. J. Agric. Food Chem. 1994, 42, 841–845. [Google Scholar] [CrossRef]
  184. Hernandez-Izquierdo, V.M.; Krochta, J.M. Thermoplastic processing of proteins for film formation—A review. J. Food Chem. 2008, 73, R30–R39. [Google Scholar] [CrossRef] [PubMed]
  185. Sothornvit, R.; Krochta, J.M. Water Vapor Permeability and Solubility of Films from Hydrolyzed Whey Protein. J. Food Sci. 2000, 65, 700–703. [Google Scholar] [CrossRef]
  186. Di Gioia, L.; Guilbert, S. Corn protein-based thermoplastic resins: Effect of some polar and amphiphilic plasticizers. J. Agric. Food Chem. 1999, 47, 1254–1261. [Google Scholar] [CrossRef] [PubMed]
  187. Redl, A.; Morel, M.H.; 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. [Google Scholar] [CrossRef]
  188. Cunningham, P.; Ogale, A.A.; Dawson, P.L.; Acton, J.C. Tensile properties of soy protein isolate films produced by a thermal compaction technique. J. Food Chem. 2000, 65, 668–671. [Google Scholar] [CrossRef]
  189. Hu, H.; Wu, J.; Li-Chan, E.C.Y.; Zhu, L.; Zhang, F.; Xu, X.; Fan, G.; Wang, L.; Huang, X.; Pan, S. Effects of ultrasound on structural and physical properties of soy protein isolate (SPI) dispersions. Food Hydrocoll. 2013, 30, 647–655. [Google Scholar] [CrossRef]
  190. Pommet, M.; Redl, A.; Guilbert, S.; Morel, M.-H. Intrinsic influence of various plasticizers on functional properties and reactivity of wheat gluten thermoplastic materials. J. Cereal Sci. 2005, 42, 81–91. [Google Scholar] [CrossRef]
  191. Sothornvit, R.; Olsen, C.W.; McHugh, T.H.; Krochta, J.M. Formation conditions, water-vapor permeability, and solubility of compression-molded whey protein films. J. Food Chem. 2003, 68, 1985–1999. [Google Scholar] [CrossRef]
  192. Swift, K.G.; Booker, J.D. Manufacturing Process Selection Handbook; Elsevier/BH Butterworth-Heinemann: Amsterdam, The Netherlands, 2013; p. 433. [Google Scholar]
  193. Guerrero, P.; La Caba, K.D. Thermal and mechanical properties of soy protein films processed at different pH by compression. J. Food Eng. 2010, 100, 261–269. [Google Scholar] [CrossRef]
  194. Jane, J.; Lim, S.; Paetau, I.; Spence, K.; Wang, S. Biodegradable Plastics Made from Agricultural Biopolymers. In Polymers from Agricultural Coproducts; Fishman, M.L., Ed.; American Chemical Society: Washington, DC, USA, 1994; pp. 92–100. [Google Scholar]
  195. Paulk, J.M.; Ogale, A.A. Thermal Processing of Food Grade Proteins. US: Soc. Plast. Eng. 1995, 2, 3139–3142. [Google Scholar]
  196. Ciannamea, E.M.; Stefani, P.M.; Ruseckaite, R.A. Physical and mechanical properties of compression molded and solution casting soybean protein concentrate based films. Food Hydrocoll. 2014, 38, 193–204. [Google Scholar] [CrossRef]
  197. Sun, S.; Song, Y.; Zheng, Q. Thermo-molded wheat gluten plastics plasticized with glycerol: Effect of molding temperature. Food Hydrocoll. 2008, 22, 1006–1013. [Google Scholar] [CrossRef]
  198. Balaguer, M.P.; Gomez-Estaca, J.; Cerisuelo, J.P.; Gavara, R.; Hernandez-Munoz, P. Effect of thermo-pressing temperature on the functional properties of bioplastics made from a renewable wheat gliadin resin. LWT Food Sci. Technol. 2014, 56, 161–167. [Google Scholar] [CrossRef]
  199. Robertson, G.L. Food Packaging: Principles and Practice, 3rd ed.; CRC Press: Boca Raton, USA, 2012; pp. 131–164. [Google Scholar]
  200. Hauck, B.W.; Huber, G.R. Single Screw vs. Twin Screw Extrusion. Cereal Foods World USA 1989, 34, 930–939. [Google Scholar]
  201. Singh, R.P.; Heldman, D.R. Introduction to Food Engineering, 5th ed.; Elsevier/Acad. Press: Amsterdam, The Netherlands, 2014; p. 8. [Google Scholar]
  202. Onwulata, C.I.; Konstance, R.P.; Cooke, P.H.; Farrell, H.M. Functionality of extrusion-texturized whey proteins. J. Dairy Sci. 2003, 86, 3775–3782. [Google Scholar] [CrossRef]
  203. Hernandez-Izquierdo, V.M.; Krochta, J.M. Thermal transitions and heat-sealing of glycerol-plasticized whey protein films. Packag. Technol. Sci. 2009, 22, 255–260. [Google Scholar] [CrossRef]
  204. Qi, P.X.; Onwulata, C.I. Physical properties, molecular structures, and protein quality of texturized whey protein isolate: Effect of extrusion moisture content. J. Dairy Sci. 2011, 94, 2231–2244. [Google Scholar] [CrossRef] [PubMed]
  205. Zhang, J.; Mungara, P.; Jane, J. Mechanical and thermal properties of extruded soy protein sheets. Polymer 2001, 42, 2569–2578. [Google Scholar] [CrossRef]
  206. Chen, F.L.; Wei, Y.M.; Zhang, B. Chemical cross-linking and molecular aggregation of soybean protein during extrusion cooking at low and high moisture content. LWT Food Sci. Technol. 2011, 44, 957–962. [Google Scholar] [CrossRef]
  207. Kester, J.J.; Richardson, T. Modification of whey proteins to improve functionality. J. Dairy Sci. 1984, 67, 2757–2774. [Google Scholar] [CrossRef]
  208. Hammann, F.; Schmid, M. Determination and quantification of molecular interactions in protein films: A review. Materials 2014, 7, 7975–7996. [Google Scholar] [CrossRef]
  209. Shih, F.F. Interaction of soy isolate with polysaccharide and its effect on film properties. J. Am. Oil Chem. Soc. 1994, 71, 1281–1285. [Google Scholar] [CrossRef]
  210. Santos, C.V.; Tomasula, P.M. Acylation and solubility of casein precipitated by carbon dioxide. J. Food Chem. 2000, 65, 227–230. [Google Scholar] [CrossRef]
  211. Childs, E.A.; Park, K.K. Functional properties of acylated glandless cottonseed flour. J. Food Chem. 1976, 41, 713–714. [Google Scholar] [CrossRef]
  212. Gould, R.F. Chemical modification for improving functional properties of plant and yeast proteins. In Functionality and Protein Structure; American Chemical Society: Ithaca, NY, USA, 1979; pp. 37–63. [Google Scholar]
  213. McElwain, M.D.; Richardson, T.; Amundson, C.H. Some functional properties of succinylated single cell protein concentrate. J. Milk Food Technol. 1975, 38, 521–526. [Google Scholar]
  214. Falbe, J.; Römpp, H.; Regitz, M. Römpp chemie lexikon; Thieme: Stuttgard, Germany, 1990; Volume 3. [Google Scholar]
  215. Embuscado, M.E.; Huber, K.C. Edible Films and Coatings for Food Applications; Springer: New York, NY, USA, 2009; pp. 25–56. [Google Scholar]
  216. Barman, B.G.; Hansen, J.R.; Mossey, A.R. Modification of the physical properties of soy protein isolate by acetylation. J. Agric. Food Chem. 1977, 25, 638–641. [Google Scholar] [CrossRef] [PubMed]
  217. Franzen, K.L.; Kinsella, J.E. Functional properties of succinylated and acetylated soy protein. J. Agric. Food Chem. 1976, 24, 788–795. [Google Scholar] [CrossRef]
  218. Ghorpade, V.M.; Li, H.; Gennadios, A.; Hanna, M.A. Chemically modified soy protein films. Trans. Asae 1995, 38, 1805–1808. [Google Scholar] [CrossRef]
  219. Barber, K.J.; Warthesen, J.J. Some functional properties of acylated wheat gluten. J. Agric. Food Chem. 1982, 30, 930–934. [Google Scholar] [CrossRef]
  220. Kim, S.H.; Kinsella, J.E. Surface active properties of proteins: Effects of progressive succinylation on film properties and foam stability of glycinin. J. Food Chem. 1987, 52, 1341–1343. [Google Scholar] [CrossRef]
  221. Bräuer, S.; Meister, F.; Gottlöber, R.P.; Nechwatal, A. Preparation and thermoplastic processing of modified plant proteins. Macromol. Mater. Eng. 2007, 292, 176–183. [Google Scholar] [CrossRef]
  222. Schotten, C. Ueber die oxydation des piperidins. Berichte der deutschen chemischen Gesellschaft 1884, 17, 2544–2547. [Google Scholar] [CrossRef]
  223. Baumann, E. Ueber eine einfache methode der darstellung von benzoësäureäthern. Berichte der deutschen chemischen Gesellschaft 1886, 19, 3218–3222. [Google Scholar] [CrossRef]
  224. Roussel-Philippe, C.; Pina, M.; Graille, J. Chemical lipophilization of soy protein isolates and wheat gluten. Eur. J. Lipid Sci. Technol. 2000, 102, 97–101. [Google Scholar] [CrossRef]
  225. Kurti, L.; Czakó, B. Strategic Applications of Named Reactions in Organic Synthesis; Elsevier: San Diego, CA, USA, 2005. [Google Scholar]
  226. Creuzenet, C.; Touati, A.; Dufour, E.; Chobert, J.M.; Haertle, T.; Choiset, Y. Acylation and alkylation of bovine .β-lactoglobulin in organic solvents. J. Agric. Food Chem. 1992, 40, 184–190. [Google Scholar] [CrossRef]
  227. Liu, T.-Y.; Chang, Y.H. Hydrolysis of proteins with p-toluenesulfonic acid: Determination of tryptophan. J. Biol. Chem. 1971, 246, 2842–2848. [Google Scholar] [PubMed]
  228. Fountoulakis, M.; Lahm, H.-W. Hydrolysis and amino acid composition analysis of proteins. J. Chromatogr. A 1998, 826, 109–134. [Google Scholar] [CrossRef]
  229. Gu, L.; Zhu, S.; Hrymak, A.N. Acidic and basic hydrolysis of poly(n-vinylformamide). J. Appl. Polym. Sci. 2002, 86, 3412–3419. [Google Scholar] [CrossRef]
  230. Akkermans, C.; Venema, P.; van der Goot, A.J.; Gruppen, H.; Bakx, E.J.; Boom, R.M.; van der Linden, E. Peptides are building blocks of heat-induced fibrillar protein aggregates of β-lactoglobulin formed at pH 2. Biomacromolecules 2008, 9, 1474–1479. [Google Scholar] [CrossRef] [PubMed]
  231. Mudgal, P.; Daubert, C.R.; Clare, D.A.; Foegeding, E.A. Effect of disulfide interactions and hydrolysis on the thermal aggregation of β-lactoglobulin. J. Agric. Food Chem. 2011, 59, 1491–1497. [Google Scholar] [CrossRef] [PubMed]
  232. Gennadios, A.; Brandenburg, A.H.; Weller, C.L.; Testin, R.F. Effect of ph on properties of wheat gluten and soy protein isolate films. J. Agric. Food Chem. 1993, 41, 1835–1839. [Google Scholar] [CrossRef]
  233. Cheftel, J.; Cuq, J.; Lorient, D. Amino acids, peptides, and proteins. Food Chem. 1985, 2, 245–369. [Google Scholar]
  234. Onwulata, C.; Huth, P. Whey processing, Functionality and Health Benefits. In IFT Press Series; Variation: IFT Press Series; Wiley-Blackwell: Ames, IA, USA, 2008. [Google Scholar]
  235. Verheul, M.; Roefs, S.P.F.M.; de Kruif, K.G. Kinetics of heat-induced aggregation of β-lactoglobulin. J. Agric. Food Chem. 1998, 46, 896–903. [Google Scholar] [CrossRef]
  236. Anker, M.; Stading, M.; Hermansson, A.-M. Effects of ph and the gel state on the mechanical properties, moisture contents, and glass transition temperatures of whey protein films. J. Agric. Food Chem. 1999, 47, 1878–1886. [Google Scholar] [CrossRef] [PubMed]
  237. Pérez-Gago, M.B.; Krochta, J.M. Water vapor permeability of whey protein emulsion films as affected by ph. J. Food Chem. 1999, 64, 695–698. [Google Scholar] [CrossRef]
  238. Li, M.; Lee, T.-C. Effect of extrusion temperature on solubility and molecular weight distribution of wheat flour proteins. J. Agric. Food Chem. 1996, 44, 763–768. [Google Scholar] [CrossRef]
  239. Hochstetter, A.; Talja, R.A.; Helén, H.J.; Hyvönen, L.; Jouppila, K. Properties of gluten-based sheet produced by twin-screw extruder. LWT Food Sci. Technol. 2006, 39, 893–901. [Google Scholar] [CrossRef]
  240. Verbeek, C.J.; Berg, L.E. Recent developments in thermo-mechanical processing of proteinous bioplastics. Recent Patents Mater. Sci. 2009, 2, 171–189. [Google Scholar] [CrossRef]
  241. Adler-Nissen, J. Enzymic hydrolysis of proteins for increased solubility. J. Agric. Food Chem. 1976, 24, 1090–1093. [Google Scholar] [CrossRef] [PubMed]
  242. Schmid, M.; Hinz, L.-V.; Wild, F.; Noller, K. Effects of hydrolysed whey proteins on the techno-functional characteristics of whey protein-based films. Materials 2013, 6, 927. [Google Scholar] [CrossRef]
  243. Kim, S.Y.; Park, P.S.W.; Rhee, K.C. Functional properties of proteolytic enzyme modified soy protein isolate. J. Agric. Food Chem. 1990, 38, 651–656. [Google Scholar] [CrossRef]
  244. Puski, G. Modification of functional properties of soy proteins by proteolytic enzyme treatment (protease preparation from aspergillus oryzae). Cereal Chem. 1975, 52, 655–664. [Google Scholar]
  245. Qi, M.; Hettiarachchy, N.S.; Kalapathy, U. Solubility and emulsifying properties of soy protein isolates modified by pancreatin. J. Food Chem. 1997, 62, 1110–1115. [Google Scholar] [CrossRef]
  246. Were, L.; Hettiarachchy, N.S.; Kalapathy, U. Modified soy proteins with improved foaming and water hydration properties. J. Food Chem. 1997, 62, 821–824. [Google Scholar] [CrossRef]
  247. Zakaria, F.; McFeeters, R.F. Improvement of emulsification properties of soy protein by limited pepsin hydrolysis. Lebensm. Wiss. Technol. 1978, 11, 42–44. [Google Scholar]
  248. Ichinose, A.; Bottenus, R.E.; Davie, E.W. Structure of transglutaminases. J. Biol. Chem. 1990, 265, 13411–13414. [Google Scholar] [PubMed]
  249. Greenberg, C.S.; Birckbichler, P.J.; Rice, R.H. Transglutaminases: Multifunctional cross-linking enzymes that stabilize tissues. FASEB J. 1991, 5, 3071–3077. [Google Scholar] [PubMed]
  250. Yildirim, M.; Hettiarachchy, N.S. Biopolymers produced by cross-linking soybean 11S globulin with whey proteins using transglutaminase. J. Food Chem. 1997, 62, 270–275. [Google Scholar] [CrossRef]
  251. Yildirim, M.; Hettiarachchy, N.S. Properties of films produced by cross-linking whey proteins and 11S globulin using transglutaminase. J. Food Chem. 1998, 63, 248–252. [Google Scholar] [CrossRef]
  252. Motoki, M.; Nio, N.; Takinami, K. Functional- properties of food proteins polymerized by transglutaminase. Agric. Biol. Chem. 1984, 48, 1257–1261. [Google Scholar]
  253. Motoki, M.; Seguro, K. Transglutaminase and its use for food processing. Trends Food Sci. Technol. 1998, 9, 204–210. [Google Scholar] [CrossRef]
  254. Wang, J.-S.; Zhao, M.-M.; Yang, X.-Q.; Jiang, Y.-M.; Chun, C. Gelation behavior of wheat gluten by heat treatment followed by transglutaminase cross-linking reaction. Food Hydrocoll. 2007, 21, 174–179. [Google Scholar] [CrossRef]
  255. Tseng, C.S.; Lai, H.M. Physicochemical properties of wheat flour dough modified by microbial transglutaminase. J. Food Chem. 2002, 67, 750–755. [Google Scholar] [CrossRef]
  256. Lorenzen, P.C. Effects of varying time/temperature-conditions of pre-heating and enzymatic cross-linking on techno-functional properties of reconstituted dairy ingredients. Food Res. Int. 2007, 40, 700–708. [Google Scholar] [CrossRef]
  257. Tang, C.-H.; Jiang, Y.; Wen, Q.-B.; Yang, X.-Q. Effect of transglutaminase treatment on the properties of cast films of soy protein isolates. J. Biotechnol. 2005, 120, 296–307. [Google Scholar] [CrossRef] [PubMed]
  258. Aboumahmoud, R.; Savello, P. Crosslinking of whey protein by transglutaminase. J. Dairy Sci. 1990, 73, 256–263. [Google Scholar] [CrossRef]
  259. Mahmoud, R.; Savello, P.A. Solubility and hydrolyzability of films produced by transglutaminase catalytic crosslinking of whey protein. J. Dairy Sci. 1993, 76, 29–35. [Google Scholar] [CrossRef]
  260. Yildirim, M.; Hettiarachchy, N.; Kalapathy, U. Properties of biopolymers from cross-linking whey protein isolate and soybean 11S globulin. J. Food Sci. 1996, 61, 1129–1132. [Google Scholar] [CrossRef]
  261. Oh, J.H.; Wang, B.; Field, P.D.; Aglan, H.A. Characteristics of edible films made from dairy proteins and zein hydrolysate cross-linked with transglutaminase. Int. J. Food Sci. Technol. 2004, 39, 287–294. [Google Scholar] [CrossRef]
  262. Di Pierro, P.; Chico, B.; Villalonga, R.; Mariniello, L.; Damiao, A.E.; Masi, P.; Porta, R. Chitosan-whey protein edible films produced in the absence or presence of transglutaminase: Analysis of their mechanical and barrier properties. Biomacromolecules 2006, 7, 744–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Hernàndez-Balada, E.; Taylor, M.M.; Phillips, J.G.; Marmer, W.N.; Brown, E.M. Properties of biopolymers produced by transglutaminase treatment of whey protein isolate and gelatin. Bioresour. Technol. 2009, 100, 3638–3643. [Google Scholar] [CrossRef] [PubMed]
  264. Zhu, Y.; Rinzema, A.; Tramper, J.; Bol, J. Microbial transglutaminase—A review of its production and application in food processing. Appl. Microbial. Biotechnol. 1995, 44, 277–282. [Google Scholar] [CrossRef]
  265. DeJong, G.; Koppelman, S. Transglutaminase catalyzed reactions: Impact on food applications. J. Food Chem. 2002, 67, 2798–2806. [Google Scholar] [CrossRef]
  266. Truong, V.-D.; Clare, D.A.; Catignani, G.L.; Swaisgood, H.E. Cross-linking and rheological changes of whey proteins treated with microbial transglutaminase. J. Agric. Food Chem. 2004, 52, 1170–1176. [Google Scholar] [CrossRef] [PubMed]
  267. Eissa, A.S.; Puhl, C.; Kadla, J.F.; Khan, S.A. Enzymatic cross-linking of β-lactoglobulin: Conformational properties using ftir spectroscopy. Biomacromolecules 2006, 7, 1707–1713. [Google Scholar] [CrossRef] [PubMed]
  268. Schmid, M.; Sängerlaub, S.; Wege, L.; Stäbler, A. Properties of transglutaminase crosslinked whey protein isolate coatings and cast films. Packag. Technol. Sci. 2014, 27, 799–817. [Google Scholar] [CrossRef]
  269. Motoki, M.; Aso, H.; Seguro, K.; Nio, N. Immobilization of enzymes in protein films prepared using transglutaminase. Agric. Biol. Chem. 1987, 51, 997–1002. [Google Scholar]
  270. Mariniello, L.; Di Pierro, P.; Esposito, C.; Sorrentino, A.; Masi, P.; Porta, R. Preparation and mechanical properties of edible pectin-soy flour films obtained in the absence or presence of transglutaminase. J. Biotechnol. 2003, 102, 191–198. [Google Scholar] [CrossRef]
  271. Jiang, Y.; Tang, C.-H.; Wen, Q.-B.; Li, L.; Yang, X.-Q. Effect of processing parameters on the properties of transglutaminase-treated soy protein isolate films. Innov. Food Sci. Emerg. Technol. 2007, 8, 218–225. [Google Scholar] [CrossRef]
  272. Gan, C.-Y.; Cheng, L.-H.; Easa, A.M. Physicochemical properties and microstructures of soy protein isolate gels produced using combined cross-linking treatments of microbial transglutaminase and maillard cross-linking. Food Res. Int. 2008, 41, 600–605. [Google Scholar] [CrossRef]
  273. Jin, M.; Zhong, Q. Transglutaminase cross-linking to enhance elastic properties of soy protein hydrogels with intercalated montmorillonite nanoclay. J. Food Eng. 2013, 115, 33–40. [Google Scholar] [CrossRef]
  274. Weng, W.; Zheng, H. Effect of transglutaminase on properties of tilapia scale gelatin films incorporated with soy protein isolate. Food Chem. 2015, 169, 255–260. [Google Scholar] [CrossRef] [PubMed]
  275. Koshy, R.R.; Mary, S.K.; Thomas, S.; Pothan, L.A. Environment friendly green composites based on soy protein isolate—A review. Food Hydrocoll. 2015, 50, 174–192. [Google Scholar] [CrossRef]
  276. Larré, C.; Desserme, C.; Barbot, J.; Gueguen, J. Properties of deamidated gluten films enzymatically cross-linked. J. Agric. Food Chem. 2000, 48, 5444–5449. [Google Scholar] [CrossRef] [PubMed]
  277. Falguera, V.; Quintero, J.P.; Jiménez, A.; Muñoz, J.A.; Ibarz, A. Edible films and coatings: Structures, active functions and trends in their use. Trends Food Sci. Technol. 2011, 22, 292–303. [Google Scholar] [CrossRef]
  278. Azeredo, H.M.C.; Waldron, K.W. Crosslinking in polysaccharide and protein films and coatings for food contact—A review. Trends Food Sci. Technol. 2016, 52, 109–122. [Google Scholar] [CrossRef]
  279. Elsabee, M.Z.; Abdou, E.S. Chitosan based edible films and coatings: A review. Mater. Sci. Eng.: C 2013, 33, 1819–1841. [Google Scholar] [CrossRef] [PubMed]
  280. Coltelli, M.-B.; Wild, F.; Bugnicourt, E.; Cinelli, P.; Lindner, M.; Schmid, M.; Weckel, V.; Müller, K.; Rodriguez, P.; Staebler, A.; et al. State of the art in the development and properties of protein-based films and coatings and their applicability to cellulose based products: An extensive review. Coatings 2016, 6, 1. [Google Scholar] [CrossRef]
  281. Li, Y.; Jiang, Y.; Liu, F.; Ren, F.; Zhao, G.; Leng, X. Fabrication and characterization of TiO2/whey protein isolate nanocomposite film. Food Hydrocoll. 2011, 25, 1098–1104. [Google Scholar] [CrossRef]
  282. Zhou, J.; Wang, S.; Gunasekaran, S. Preparation and characterization of whey protein film incorporated with TiO2 nanoparticles. J. Food Sci. 2009, 74, N50–N56. [Google Scholar] [CrossRef] [PubMed]
  283. Shi, L.; Zhou, J.; Gunasekaran, S. Low temperature fabrication of ZnO–whey protein isolate nanocomposite. Mater. Lett. 2008, 62, 4383–4385. [Google Scholar] [CrossRef]
  284. Sothornvit, R.; Rhim, J.-W.; Hong, S.-I. Effect of nano-clay type on the physical and antimicrobial properties of whey protein isolate/clay composite films. J. Food Eng. 2009, 91, 468–473. [Google Scholar] [CrossRef]
  285. Oymaci, P.; Altinkaya, S.A. Improvement of barrier and mechanical properties of whey protein isolate based food packaging films by incorporation of zein nanoparticles as a novel bionanocomposite. Food Hydrocoll. 2016, 54, 1–9. [Google Scholar] [CrossRef]
  286. De Azeredo, H.M. Nanocomposites for food packaging applications. Food Res. Int. 2009, 42, 1240–1253. [Google Scholar] [CrossRef]
  287. Arora, A.; Padua, G. Review: Nanocomposites in food packaging. J. Food Chem. 2010, 75, R43–R49. [Google Scholar] [CrossRef] [PubMed]
  288. Dang, Q.; Lu, S.; Yu, S.; Sun, P.; Yuan, Z. Silk fibroin/montmorillonite nanocomposites: Effect of pH on the conformational transition and clay dispersion. Biomacromolecules 2010, 11, 1796–1801. [Google Scholar] [CrossRef] [PubMed]
  289. Kristo, E.; Biliaderis, C.G. Physical properties of starch nanocrystal-reinforced pullulan films. Carbohydr. Polym. 2007, 68, 146–158. [Google Scholar] [CrossRef]
  290. González, A.; Igarzabal, C.I.A. Nanocrystal-reinforced soy protein films and their application as active packaging. Food Hydrocoll. 2015, 43, 777–784. [Google Scholar] [CrossRef]
  291. Wang, S.-Y.; Zhu, B.-B.; Li, D.-Z.; Fu, X.-Z.; Shi, L. Preparation and characterization of TiO2/SPI composite film. Mater. Lett. 2012, 83, 42–45. [Google Scholar] [CrossRef]
  292. Zolfi, M.; Khodaiyan, F.; Mousavi, M.; Hashemi, M. Development and characterization of the kefiran-whey protein isolate-TiO2 nanocomposite films. Int. J. Biol. Macromol. 2014, 65, 340–345. [Google Scholar] [CrossRef] [PubMed]
  293. Fornes, T.; Paul, D. Modeling properties of nylon 6/clay nanocomposites using composite theories. Polymer 2003, 44, 4993–5013. [Google Scholar] [CrossRef]
  294. Zeng, Q.; Yu, A.; Lu, G.; Paul, D. Clay-based polymer nanocomposites: Research and commercial development. J. Nanosci. Nanotechnol. 2005, 5, 1574–1592. [Google Scholar] [CrossRef] [PubMed]
  295. Kumar, P.; Sandeep, K.; Alavi, S.; Truong, V.-D.; Gorga, R. Preparation and characterization of bio-nanocomposite films based on soy protein isolate and montmorillonite using melt extrusion. J. Food Eng. 2010, 100, 480–489. [Google Scholar] [CrossRef]
  296. Kumar, P.; Sandeep, K.; Alavi, S.; Truong, V.-D.; Gorga, R. Effect of type and content of modified montmorillonite on the structure and properties of bio-nanocomposite films based on soy protein isolate and montmorillonite. J. Food Sci. 2010, 75, N46–N56. [Google Scholar] [CrossRef] [PubMed]
  297. Chen, P.; Zhang, L. Interaction and properties of highly exfoliated soy protein/montmorillonite nanocomposites. Biomacromolecules 2006, 7, 1700–1706. [Google Scholar] [CrossRef] [PubMed]
  298. Echeverría, I.; Eisenberg, P.; Mauri, A.N. Nanocomposites films based on soy proteins and montmorillonite processed by casting. J. Membr. Sci. 2014, 449, 15–26. [Google Scholar] [CrossRef]
  299. Li, X.; Ji, N.; Qiu, C.; Xia, M.; Xiong, L.; Sun, Q. The effect of peanut protein nanoparticles on characteristics of protein- and starch-based nanocomposite films: A comparative study. Ind. Crops Prod. 2015, 77, 565–574. [Google Scholar] [CrossRef]
  300. Tunc, S.; Angellier, H.; Cahyana, Y.; Chalier, P.; Gontard, N.; Gastaldi, E. Functional properties of wheat gluten/montmorillonite nanocomposite films processed by casting. J. Membr. Sci. 2007, 289, 159–168. [Google Scholar] [CrossRef]
  301. Guilherme, M.R.; Mattoso, L.H.C.; Gontard, N.; Guilbert, S.; Gastaldi, E. Synthesis of nanocomposite films from wheat gluten matrix and mmt intercalated with different quaternary ammonium salts by way of hydroalcoholic solvent casting. Compos. Part A: Appl. Sci. Manuf. 2010, 41, 375–382. [Google Scholar] [CrossRef]
  302. Mascheroni, E.; Chalier, P.; Gontard, N.; Gastaldi, E. Designing of a wheat gluten/montmorillonite based system as carvacrol carrier: Rheological and structural properties. Food Hydrocoll. 2010, 24, 406–413. [Google Scholar] [CrossRef]
  303. Türe, H.; Gällstedt, M.; Johansson, E.; Hedenqvist, M.S. Wheat-gluten/montmorillonite clay multilayer-coated paperboards with high barrier properties. Ind. Crops Prod. 2013, 51, 1–6. [Google Scholar] [CrossRef]
  304. Vonasek, E.; Le, P.; Nitin, N. Encapsulation of bacteriophages in whey protein films for extended storage and release. Food Hydrocoll. 2014, 37, 7–13. [Google Scholar] [CrossRef]
  305. Joerger, R.D. Antimicrobial films for food applications: A quantitative analysis of their effectiveness. Packag. Technol. Sci. 2007, 20, 231–273. [Google Scholar] [CrossRef]
  306. Rocha, M.; Ferreira, F.; Souza, M.; Prentice, C. Antimicrobial Films: A Review. In Microbial Pathogens and Strategies for Combating Them: Science, Technology and Education; Formatex Research Center Badajoz, Spain, 2013; pp. 23–31. [Google Scholar]
  307. Garcia, P.; Martinez, B.; Obeso, J.; Rodriguez, A. Bacteriophages and their application in food safety. Lett. Appl. Microbial. 2008, 47, 479–485. [Google Scholar] [CrossRef] [PubMed]
  308. Min, S.; Harris, L.J.; Krochta, J.M. Listeria monocytogenes inhibition by whey protein films and coatings incorporating the lactoperoxidase system. J. Food Chem. 2005, 70, M317–M324. [Google Scholar] [CrossRef]
  309. Seydim, A.C.; Sarikus, G. Antimicrobial activity of whey protein based edible films incorporated with oregano, rosemary and garlic essential oils. Food Res. Int. 2006, 39, 639–644. [Google Scholar] [CrossRef]
  310. Dangaran, K.; Krochta, J.M. Whey .protein films and coatings. Whey Process Func. Health Benefits 2009, 82, 133. [Google Scholar] [CrossRef]
  311. Pintado, C.M.B.S.; Ferreira, M.A.S.S.; Sousa, I. Control of pathogenic and spoilage microorganisms from cheese surface by whey protein films containing malic acid, nisin and natamycin. Food Control 2010, 21, 240–246. [Google Scholar] [CrossRef]
  312. Chen, H. Functional properties and applications of edible films made of milk proteins. J. Dairy Sci. 1995, 78, 2563–2583. [Google Scholar] [CrossRef]
  313. Hotchkiss, J. Safety Considerations in Active packaging; Springer: Dordrecht, Netherlands, 1995; p. 236. [Google Scholar]
  314. Ozdemir, M.; Floros, J.D. Optimization of edible whey protein films containing preservatives for mechanical and optical properties. J. Food Eng. 2008, 84, 116–123. [Google Scholar] [CrossRef]
  315. Ozdemir, M.; Floros, J.D. Optimization of edible whey protein films containing preservatives for water vapor permeability, water solubility and sensory characteristics. J. Food Eng. 2008, 86, 215–224. [Google Scholar] [CrossRef]
  316. Young, S.; Sarda, X.; Rosenberg, M. Microencapsulating Properties of Whey Proteins. 1. Microencapsulation of anhydrous milk fat. J. Dairy Sci. 1993, 76, 2868–2877. [Google Scholar] [CrossRef]
  317. Young, S.; Sarda, X.; Rosenberg, M. Microencapsulating properties of whey proteins. 2. Combination of whey proteins with carbohydrates. J. Dairy Sci. 1993, 76, 2878–2885. [Google Scholar] [CrossRef]
  318. González, A.; Igarzabal, C.I.A. Soy protein-poly (lactic acid) bilayer films as biodegradable material for active food packaging. Food Hydrocoll. 2013, 33, 289–296. [Google Scholar] [CrossRef]
  319. Emiroğlu, Z.K.; Yemiş, G.P.; Coşkun, B.K.; Candoğan, K. Antimicrobial activity of soy edible films incorporated with thyme and oregano essential oils on fresh ground beef patties. Meat Sci. 2010, 86, 283–288. [Google Scholar] [CrossRef] [PubMed]
  320. Sivarooban, T.; Hettiarachchy, N.; Johnson, M. Physical and antimicrobial properties of grape seed extract, nisin, and EDTA incorporated soy protein edible films. Food Res. Int. 2008, 41, 781–785. [Google Scholar] [CrossRef]
  321. Ou, S.; Wang, Y.; Tang, S.; Huang, C.; Jackson, M.G. Role of ferulic acid in preparing edible films from soy protein isolate. J. Food Eng. 2005, 70, 205–210. [Google Scholar] [CrossRef]
  322. Friesen, K.; Chang, C.; Nickerson, M. Incorporation of phenolic compounds, rutin and epicatechin, into soy protein isolate films: Mechanical, barrier and cross-linking properties. Food Chem. 2015, 172, 18–23. [Google Scholar] [CrossRef] [PubMed]
  323. Banerjee, R.; Chen, H. Functional properties of edible films using whey protein concentrate. J. Dairy Sci. 1995, 78, 1673–1683. [Google Scholar] [CrossRef]
  324. Shellhammer, T.; Krochta, J. Whey protein emulsion film performance as affected by lipid type and amount. J. Food Chem. 1997, 62, 390–394. [Google Scholar] [CrossRef]
  325. Pérez-Gago, M.B.; Krochta, J.M. Lipid particle size effect on water vapor permeability and mechanical properties of whey protein/beeswax emulsion films. J. Agric. Food Chem. 2001, 49, 996–1002. [Google Scholar] [CrossRef] [PubMed]
  326. Anker, M.; Berntsen, J.; Hermansson, A.-M.; Stading, M. Improved water vapor barrier of whey protein films by addition of an acetylated monoglyceride. Innov. Food Sci. Emerg. Technol. 2002, 3, 81–92. [Google Scholar] [CrossRef]
  327. Jiménez, A.; Fabra, M.; Talens, P.; Chiralt, A. Effect of lipid self-association on the microstructure and physical properties of hydroxypropyl-methylcellulose edible films containing fatty acids. Carbohydr. Polym. 2010, 82, 585–593. [Google Scholar] [CrossRef]
  328. Galus, S.; Kadzińska, J. Whey protein edible films modified with almond and walnut oils. Food Hydrocoll. 2016, 52, 78–86. [Google Scholar] [CrossRef]
  329. Talens, P.; Krochta, J.M. Plasticizing effects of beeswax and carnauba wax on tensile and water vapor permeability properties of whey protein films. J. Food Chem. 2005, 70, E239–E243. [Google Scholar] [CrossRef]
  330. Soazo, M.; Pérez, L.; Rubiolo, A.; Verdini, R. Effect of freezing on physical properties of whey protein emulsion films. Food Hydrocoll. 2013, 31, 256–263. [Google Scholar] [CrossRef]
  331. Janjarasskul, T.; Rauch, D.J.; McCarthy, K.L.; Krochta, J.M. Barrier and tensile properties of whey protein-candelilla wax film/sheet. LWT–Food Sci. Technol. 2014, 56, 377–382. [Google Scholar] [CrossRef]
  332. Oussalah, M.; Caillet, S.; Salmiéri, S.; Saucier, L.; Lacroix, M. Antimicrobial and antioxidant effects of milk protein-based film containing essential oils for the preservation of whole beef muscle. J. Agric. Food Chem. 2004, 52, 5598–5605. [Google Scholar] [CrossRef] [PubMed]
  333. Lee, S.Y.; Trezza, T.; Guinard, J.X.; Krochta, J. Whey-protein-coated peanuts assessed by sensory evaluation and static headspace gas chromatography. J. Food Sci. 2002, 67, 1212–1218. [Google Scholar] [CrossRef]
  334. Lee, S.-Y.; Krochta, J. Accelerated shelf life testing of whey-protein-coated peanuts analyzed by static headspace gas chromatography. J. Agric. Food Chem. 2002, 50, 2022–2028. [Google Scholar] [CrossRef] [PubMed]
  335. Perez-Gago, M.; Serra, M.; Del Rio, M. Color change of fresh-cut apples coated with whey protein concentrate-based edible coatings. Postharvest Biol. Technol. 2006, 39, 84–92. [Google Scholar] [CrossRef]
  336. Gontard, N.; Marchesseau, S.; CUQ, J.L.; Guilbert, S. Water vapour permeability of edible bilayer films of wheat gluten and lipids. Int. J. Food Sci. Technol. 1995, 30, 49–56. [Google Scholar] [CrossRef]
  337. Gennadios, A.; Weller, C.L.; Testin, R.F. Modification of physical and barrier properties of edible wheat gluten-based films. Cereal Chem. 1993, 70, 426–429. [Google Scholar]
  338. Tanada-Palmu, P.S.; Grosso, C.R. Effect of edible wheat gluten-based films and coatings on refrigerated strawberry (fragaria ananassa) quality. Postharvest Biol. Technol. 2005, 36, 199–208. [Google Scholar] [CrossRef]
  339. Zhang, X.; Do, M.D.; Kurniawan, L.; Qiao, G.G. Wheat gluten-based renewable and biodegradable polymer materials with enhanced hydrophobicity by using epoxidized soybean oil as a modifier. Carbohydr. Res. 2010, 345, 2174–2182. [Google Scholar] [CrossRef] [PubMed]
  340. Rhim, J.-W.; Gennadios, A.; Weller, C.L.; Cezeirat, C.; Hanna, M.A. Soy protein isolate-dialdehyde starch films. Ind. Crops Prod. 1998, 8, 195–203. [Google Scholar] [CrossRef]
  341. Chao, Z.; Yue, M.; Xiaoyan, Z.; Dan, M. Development of soybean protein-isolate edible films incorporated with beeswax, span 20, and glycerol. J. Food Sci. 2010, 75, C493–C497. [Google Scholar] [CrossRef] [PubMed]
  342. Park, S.K.; Hettiarachchy, N.S.; Ju, Z.; Gennadios, A. Formation and properties of soy protein films and coatings. In Protein-Based Films and Coatings; CRC Press: New York, USA, 2002; pp. 978–1587. [Google Scholar]
  343. Gennadios, A.; Cezeirat, C.; Weller, C.L.; Hanna, M.A. Emulsified soy protein-lipid films. In Paradigm for Successful Utilization of Renewable Resources; AOCS Press: Champaign, USA, 1998; pp. 213–226. [Google Scholar]
  344. Cao, N.; Fu, Y.; He, J. Preparation and physical properties of soy protein isolate and gelatin composite films. Food Hydrocoll. 2007, 21, 1153–1162. [Google Scholar] [CrossRef]
  345. Were, L.; Hettiarachchy, N.; Coleman, M. Properties of cysteine-added soy protein-wheat gluten films. J. Food Sci. 1999, 64, 514–518. [Google Scholar] [CrossRef]
  346. Song, Y.; Zheng, Q.; Liu, C. Green biocomposites from wheat gluten and hydroxyethyl cellulose: Processing and properties. Ind. Crops Prod. 2008, 28, 56–62. [Google Scholar] [CrossRef]
  347. Schmid, M.; Dallmann, K.; Bugnicourt, E.; Cordoni, D.; Wild, F.; Lazzeri, A.; Noller, K. Properties of whey protein coated films and laminates as novel recyclable food packaging materials with excellent barrier properties. Int. J. Polym. Sci. 2012, 2012, 7. [Google Scholar] [CrossRef]
  348. Cinelli, P.; Schmid, M.; Bugnicourt, E.; Wildner, J.; Bazzichi, A.; Anguillesi, I.; Lazzeri, A. Whey protein layer applied on biodegradable packaging film to improve barrier properties while maintaining biodegradability. Polym. Degrad. Stab. 2014, 108, 151–157. [Google Scholar] [CrossRef]
  349. Schmid, M.; Krimmel, B.; Grupa, U.; Noller, K. Effects of thermally induced denaturation on technological-functional properties of whey protein isolate-based films. J. Dairy Sci. 2014, 97, 5315–5327. [Google Scholar] [CrossRef] [PubMed]
  350. Guo, M.; Wang, G. Whey protein polymerisation and its applications in environmentally safe adhesives. Int. J. Dairy Technol. 2016. [Google Scholar] [CrossRef]
  351. Liu, H.; Li, C.; Sun, X.S. Improved water resistance in undecylenic acid (UA)-modified soy protein isolate (SPI)-based adhesives. Ind. Crops Prod. 2015, 74, 577–584. [Google Scholar] [CrossRef]
  352. Nordqvist, P.; Nordgren, N.; Khabbaz, F.; Malmström, E. Plant proteins as wood adhesives: Bonding performance at the macro and nanoscale. Ind. Crops Prod. 2013, 44, 246–252. [Google Scholar] [CrossRef]
  353. Santoni, I.; Pizzo, B. Evaluation of alternative vegetable proteins as wood adhesives. Ind. Crop Prod. 2013, 45, 148–154. [Google Scholar] [CrossRef]
  354. Kumar, R.; Choudhary, V.; Mishra, S.; Varma, I.K.; Mattiason, B. Adhesives and plastics based on soy protein products. Ind. Crops Prod. 2002, 16, 155–172. [Google Scholar] [CrossRef]
  355. Schmid, M.; Zillinger, W.; Müller, K.; Sängerlaub, S. Permeation of water vapour, nitrogen, oxygen and carbon dioxide through whey protein isolate based films and coatings—Permselectivity and activation energy. Food Packag. Shelf Life 2015, 6, 21–29. [Google Scholar] [CrossRef]
  356. Patachia, S.; Croitoru, C. 14-Biopolymers for Wood Preservation. In Biopolymers and Biotech Admixtures for Eco-Efficient Construction Materials; Woodhead Publishing: Waltham, MA, USA, 2016; pp. 305–332. [Google Scholar]
  357. Verbeek, C.J.R.; Klunker, E. Thermoplastic protein nano-composites using bloodmeal and bentonite. J. Polym. Environ. 2013, 21, 963–970. [Google Scholar] [CrossRef]
  358. Song, Y.H.; Seo, J.H.; Choi, Y.S.; Kim, D.H.; Choi, B.-H.; Cha, H.J. Mussel adhesive protein as an environmentally-friendly harmless wood furniture adhesive. Int. J. Adhes. Adhes. 2016, 70, 260–264. [Google Scholar] [CrossRef]
  359. Latza, V.; Guerette, P.A.; Ding, D.; Amini, S.; Kumar, A.; Schmidt, I.; Keating, S.; Oxman, N.; Weaver, J.C.; Fratzl, P.; et al. Multi-scale thermal stability of a hard thermoplastic protein-based material. Nat. Commun. 2015, 6. [Google Scholar] [CrossRef] [PubMed]
  360. Torculas, M.; Medina, J.; Xue, W.; Hu, X. Protein-based bioelectronics. ACS Biomater. Sci. Eng. 2016, 2, 1211–1223. [Google Scholar] [CrossRef]
  361. Das, M.; Chowdhury, T. Heat sealing property of starch based self-supporting edible films. Food Packag. Shelf Life 2016, 9, 64–68. [Google Scholar] [CrossRef]
  362. Odila Pereira, J.; Soares, J.; Sousa, S.; Madureira, A.R.; Gomes, A.; Pintado, M. Edible films as carrier for lactic acid bacteria. LWT Food Sci. Technol. 2016, 73, 543–550. [Google Scholar] [CrossRef]
Figure 1. Effect of plasticizer content on thermoplastic processing, adapted from [4].
Figure 1. Effect of plasticizer content on thermoplastic processing, adapted from [4].
Ijms 17 01376 g001
Figure 2. Reductive methylation, adapted from [207].
Figure 2. Reductive methylation, adapted from [207].
Ijms 17 01376 g002
Figure 3. Acetylation of protein chains, adapted from [215].
Figure 3. Acetylation of protein chains, adapted from [215].
Ijms 17 01376 g003
Figure 4. Succinylation of protein chains, adapted from [215].
Figure 4. Succinylation of protein chains, adapted from [215].
Ijms 17 01376 g004
Figure 5. Reaction of proteins with fatty acids, adapted from [225].
Figure 5. Reaction of proteins with fatty acids, adapted from [225].
Ijms 17 01376 g005
Figure 6. Acidic and Basic Hydrolysis of Proteins, adapted from [30,229].
Figure 6. Acidic and Basic Hydrolysis of Proteins, adapted from [30,229].
Ijms 17 01376 g006
Figure 7. Quaternary structure of β-Lg as a function of pH at low temperature and low concentration, adapted from [234].
Figure 7. Quaternary structure of β-Lg as a function of pH at low temperature and low concentration, adapted from [234].
Ijms 17 01376 g007
Figure 8. Protein hydrolysis and synthesis of peptide bonds [30].
Figure 8. Protein hydrolysis and synthesis of peptide bonds [30].
Ijms 17 01376 g008
Figure 9. Reactions catalyzed by transglutaminase, adapted from [30].
Figure 9. Reactions catalyzed by transglutaminase, adapted from [30].
Ijms 17 01376 g009
Table 1. Prolamin and glutelin proteins of whey gluten divided into high, medium, and low molecular mass groups with their subunits and their wheat gluten and cysteine contents [62].
Table 1. Prolamin and glutelin proteins of whey gluten divided into high, medium, and low molecular mass groups with their subunits and their wheat gluten and cysteine contents [62].
GroupHigh MolecularMedium MolecularLow Molecular
HMW-Subunitsω-GliadinGliadinLMW-Subunit
x-typey-typeω5ω1,2αγ
Gluten protein content (%)4–93–43–64–728–3323–3119–25
Sum of cysteine4700688
Table 2. Amino acid composition and commonly used modifications [207,208].
Table 2. Amino acid composition and commonly used modifications [207,208].
Side ChainAmino AcidCommonly Used Modificationsβ-Lactoglobulin (Whey Protein) (mol %)β-Conglycinin (Soy Protein) (mol %)γ-Gliadins (Wheat Gluten) (mol %)
AminoLysineAlkylation, Acylation, TG 110.56.1-
Arginine2.58.31.8
CarboxylGlutamineAmidation, Esterification11.2--
Glutamic acidTG 16.224.545.8
Asparagine 3.112.02.9
Aspartic acid 6.9--
DisulfideCysteineReduction, Oxidation2.80.03-
ImiazoleHistidineAlkylation, Oxidation1.52.81.6
IndoleTryptophanAlkylation, Oxidation2.0--
PhenolicTyrosineAcylation, electrophilic3.63.53.5
TryptophanSubstitution2.0--
Histidine-1.52.81.6
Phenylalanine-3.25.45.2
SulfhydrylCysteineAlkylation, Oxidation2.80.03-
ThioetherMethionineAlkylation, Oxidation---
1 Transglutaminase crosslinking: Isopeptide bonding between glutamine and lysine.
Table 3. Mechanical properties of protein-based films.
Table 3. Mechanical properties of protein-based films.
Film ModificationTensile Strength (MPa)Elongation (%)Elastic Modulus (MPa)Source
Heating
Heat treatment of solution
WPI: 70 °C:5 min3.4 ± 0.47 ± 1156 ± 17[79]
WPI: 70 °C:10 min3.3 ± 0.18 ± 2141 ± 10[79]
WPI: 70 °C:15 min4.5 ± 0.29 ± 0.4194 ± 12[79]
WPI: 70 °C:20 min4.9 ± 0.217 ± 5192 ± 13[79]
WPI: 80 °C:5 min3.8 ± 0.37 ± 1159 ± 24[79]
WPI: 80 °C:10 min7 ± 218 ± 3299 ± 62[79]
WPI: 80 °C:15 min12 ± 217 ± 4346 ± 71[79]
WPI: 80 °C:20 min14 ± 218 ± 4460 ± 42[79]
WPI: 90 °C:5 min8 ± 23 ± 1327 ± 71[77]
WPI: 90 °C:10 min12 ± 214 ± 3429 ± 59[79]
WPI: 90 °C:15 min12 ± 216 ± 3427 ± 41[79]
WPI: 90 °C:20 min13 ± 216 ± 5472 ± 57[79]
WPI: 90 °C:30 min6.941199[75]
WPI: 100 °C:5 min8 ± 314 ± 3342 ± 32[79]
WPI: 100 °C:10 min10 ± 314 ± 4419 ± 53[79]
WPI: 100 °C:15 min12 ± 215 ± 5425 ± 28[79]
WPI: 100 °C:20 min9 ± 218 ± 3429 ± 28[79]
SPI control11.2 ± 2.010.2 ± 5.5928 ± 233[84]
SPI: 85 °C12.8 ± 2.716.8 ± 6.6992 ± 276[84]
Heat treatment of the film
SPI: 60 °C:24 h11 a180 an/a[23]
SPI: 70 °C:24 h9 a170 an/a[23]
SPI: 80 °C:24 h13 a160 an/a[23]
SPI control8.2 ± 0.230 ± 3.3n/a[78]
SPI: 90 °C:24 h14.7 ± 0.46.1 ± 0.7n/a[78]
WG control1.7 ± 0.3501 ± 4613 ± 3[27]
WG: 80 °C:15 min2.4 ± 0.4391 ± 5829 ± 13[27]
WG: 95 °C:15 min2.5 ± 0.3386 ± 7530 ± 11[27]
WG: 110 °C:15 min3.1 ± 0.4327 ± 5837 ± 10[27]
WG: 125 °C:15 min6.3 ± 0.5275 ± 1240 ± 14[27]
WG: 140 °C:15 min7.3 ± 1.2170 ± 2698 ± 16[27]
WG: 140 °C:1.5 min4.2 ± 1.1326 ± 4937 ± 8[27]
Ultrasound
WPC control3.36 ± 0.24n/an/a[151]
WPC: 168 kHz:1.9 W:0.5 h2.75 ±0.32n/an/a[151]
WPC: 168 kHz:1.9 W:1 h4.40 ± 0.20n/an/a[151]
WPC: 520 kHz:3 W:0.5 h3.67 ± 0.47n/an/a[151]
WPC: 520 kHz:3 W:1 h4.92 ± 0.72n/an/a[151]
WPC: 168 kHz:3.35 W:0.5 h1.75 ± 0.85n/an/a[151]
WPC: 168 kHz:3.35 W:1 h3.08 ± 0.28n/an/a[151]
WPC: 520 kHz:5.22 W:0.5 h2.75 ± 0.36n/an/a[151]
WPC: 520 kHz:5.22 W:0.5 h4.47 ± 0.14n/an/a[151]
WPI control1.1 an/a19 a[150]
WPI: amplitude 16 µm1.1 an/a18 a[150]
WPI: amplitude 80 µm1.1 an/a20 a[150]
WPI: amplitude 160 µm1.2 an/a17 a[150]
UV and γ-irradiation
UV irradiation of solution
WPI control4.68 ± 0.39114.0 ± 14.2n/a[173]
WPI: 324 J·cm−26.40 ± 0.30110.0 ± 15.5n/a[173]
SPI controln/an/a299 ± 13[174]
SPI: 125 W:2 hn/an/a309 ± 27[174]
SPI control3.7 a124.2 an/a[162]
SPI: 51.8 J·m−25.25 a113 an/a[162]
UV irradiation of the film
WPI control6.8 MD42 MD118 MD[13]
WPI: 2.3 J·cm−28.2 MD50 MD125 MD[13]
WPI: 10.2 J·cm−28.1 MD53 MD95 MD[13]
WPI: 19.0 J·cm−28 MD54 MD62 MD[13]
WPI: 31.4 J·cm−28.8 MD39 MD115 MD[13]
SPI controln/an/a299 ± 13[174]
SPI: 125 W:2 hn/an/a256 ± 42[174]
SPI control8.2 ± 0.230 ± 3.3n/a[78]
SPI: 51.8 J·m−210.0 ± 0.623.3 ± 5.6n/a[78]
WG control1.2 ± 0.2n/an/a[163]
WG: 51.8 J·m−22.0 ± 0.1n/an/a[163]
WG control1.7 ± 0.3501 ± 4613 ± 3[27]
WG: 0.25 J·cm−22.0 ± 0.3424 ± 9718 ± 6[27]
WG: 1 J·cm−22.0 ± 0.3478 ± 7015 ± 4[27]
γ-irradiation of the film
WG control2.1 ± 0.5384 ± 8229 ± 4[27]
WG: 10 kGy3.0 ± 0.7261 ± 8252 ± 15[27]
WG: 20 kGy2.6 ± 0.4344 ± 4536 ± 6[27]
WG: 40 kGy2.7 ± 0.4297 ± 5438 ± 7[27]
Compression Molding
WPI: 30% Gly:0.81 MPa:113 °C7.5 a35 a180 a[82]
WPI: 30% Gly:0.81 MPa:127 °C10.5 a35 a275 a[82]
WPI: 30% Gly:0.81 MPa:140 °C10 a38 a245 a[82]
WPI: 30% Gly:2.25MPa:113 °C5.5 a57 a125 a[82]
WPI: 30% Gly:2.25 MPa:127 °C10 a50 a250 a[82]
WPI: 30% Gly:2.25 MPa:140 °C10.5 a48 a250 a[82]
Chemical Modifications
Acylation
SPI control2.5n/an/a[218]
SPI:acetylated2.5n/an/a[218]
SPI:succinylated2.6n/an/a[218]
Alkylation
SPI:PGA:pH 80.779 ± 0.07618.3 ± 0.9n/a[209]
SPI:PGA:KOH0.848 ± 0.06922.8 ± 1.5n/a[209]
Hydrolysis
WPI control4.35 a75.8 a140 a[185]
WPI:5.5% DH0.5 a31.5 a3 a[185]
WPI:10% DH1.5 a7.7 a50 a[185]
pH alteration
WPI:pH 7n/a40 ± 478 ± 3[236]
WPI:pH 8n/a54 ± 574 ± 2[236]
WPI:pH 9n/a66 ± 764 ± 2[236]
SPI:pH 63.5 ± 0.272 ± 13.2n/a[232]
SPI:pH 83.6 ± 0.4139.5 ± 19.5n/a[232]
SPI:pH 103.6 ± 0.1169.3 ± 9.3n/a[232]
SPI:pH 121.3 ± 0.566.5 ± 31.6n/a[232]
Enzymatic Crosslinking
WPI control5.64 ± 0.64n/an/a[251]
WPI:TG12.53 ± 1.12n/an/a[251]
WPI:SPI 11S control6.26 ± 0.88n/an/a[251]
WPI:SPI 11S:TG17.86 ± 1.44n/an/a[251]
WPI control4 a25 an/a[261]
WPI:TG3.2 a105 an/a[261]
WPI:Chitosan 0.25 mg·cm−29.5 ± 0.614.1 ± 0.59n/a[262]
WPI:Chitosan:TG26.2 ± 0.93.1 ± 0.3n/a[262]
SPI 11S control7.61 ± 0.71n/an/a[251]
SPI 11S:TG16.4 ± 1.38n/an/a[251]
SPI control11.2 ± 2.010.2 ± 5.5928 ± 233[84]
SPI:Horseradish Peroxidasen/a1.1 ± 0.51503 ± 129[84]
SPI:Pectin control6.8 ± 0.9211.61 ± 1.09n/a[270]
SPI:Pectin:TG12.4 ± 1.057.2 ± 1.03n/a[270]
SPI:Gly control2.21 ± 0.25159.87 ± 9.20n/a[257]
SPI:Gly:mTG2.58 ± 0.28105.88 ± 9.20n/a[257]
SPI:Gly:Sorbitol control2.58 ± 0.11102.04 ± 13.68n/a[257]
SPI:Gly:Sorbitol:mTG3.10 ± 0.1780.04 ± 5.40n/a[257]
SPI:Sorbitol control4.16 ± 0.04101.77 ± 15.60n/a[257]
SPI:Sorbitol:mTG4.48 ± 0.3527.33 ± 3.61n/a[257]
SPI:mTG 0 U·g−13.12 ± 0.25167.1 ± 22.8n/a[271]
SPI:mTG 4 U·g−13.75 ± 0.30124.4 ± 11.8n/a[271]
SPI:mTG 10 U·g−13.98 ± 0.20108.1 ± 9.1n/a[271]
SPI:mTG 40 U·g−11.75 ± 0.2783.3 ± 12.1n/a[271]
SPI:mTG 60 U·g−10.95 ± 0.1857.4 ± 7.8n/a[271]
SPI:Gelatin control33.78 ± 5.8139.17 ± 6.76n/a[274]
SPI:Gelatin:TG 10%65.73 ± 7.0732.93 ± 7.04n/a[274]
SPI:Gelatin:TG 20%45.32 ± 5.7914.20 ± 3.66n/a[274]
SPI:Gelatin:TG 30%42.74 ± 4.1512.12 ± 4.29n/a[274]
WG:Putrescine control1.12 ± 0.11327 ± 326.76 ± 2.99[276]
WG:Putrescine 0.09 mol:mol TG1.61 ± 0.16582 ± 572.15 ± 0.62[276]
WG:Cadaverine control0.94 ± 0.06304 ± 739.51 ± 3.40[276]
WG:Cadaverine 0.09 mol:mol TG1.24 ± 0.15522 ± 832.04 ± 0.34[276]
WG:Diaminohexane control0.89 ± 0.06351 ± 816.50 ± 1.68[276]
WG:Diaminohexane 0.09 mol:mol TG1.87 ± 0.25527 ± 853.96 ± 0.91[276]
WG:Diaminooctance control0.75 ± 0.06413 ± 1184.18 ± 1.43[276]
WG:Diaminohectane 0.09 mol:mol TG1.53 ± 0.24501 ± 732.92 ± 0.15[276]
Composite Films and Bioactive Compounds
WPI control28.0 a1.8 a1675 a[324]
WPI:40% carnauba wax22.5 a1.5 a1700 a[324]
WPI:40% candelilla wax17.0 a1.0 a1800 a[324]
WPI:40% milk fat fraction19.0 a2.2 a975 a[324]
WPI:40% BW18.5 a2.0 a1200 a[324]
WPI:Gly:BW (60:20:20) PS 0.5 µm10.2 ± 0.8 a4.9 ± 1.3 a550 ± 20 a[325]
WPI:Gly:BW (30:10:60) PS 0.5 µm5.5 ± 0.2 a3 ± 0.2 a430 ± 30 a[325]
WPI:Gly:BW (30:10:60) PS 1.0 µm4.2 ± 0.3 a2.7 ± 0.3 a420 ± 10 a[325]
WPI:Gly:BW (30:10:60) PS 1.5 µm3.0 ± 0.2 a1.2 ± 0.2 a410 ± 30 a[325]
WPI:Gly:BW (30:10:60) PS 2.0 µm2.9 ± 0.05 a1.2 ± 0.2 a380 ± 15 a[325]
WPI control2.2 ± 0.1120 ± 396 ± 3.7[326]
WPI:Acetem1.0 ± 0.0829 ± 536 ± 3.3[326]
WPI control7.1 ± 1.829.8 ± 6.51.9 ± 0.9[328]
WPI:Almond oil 0.5%10.2 ± 1.921.9 ± 2.63.6 ± 0.7[328]
WPI:Almond oil 1.0%5.4 ± 0.853.7 ± 7.71.4 ± 0.2[328]
WPI:Walnut oil 0.5%11.8 ± 0.914.5 ± 4.53.8 ± 0.6[328]
WPI:Walnut oil 1.0%6.9 ± 1.524.9 ± 4.92.3 ± 0.6[328]
WPI:Gly:(1:1)2.9 ± 0.4118 ± 20 a41 ± 4[329]
WPI:BW:Gly (1:1:1)1.2 ± 0.130 ± 10 a39 ± 12[329]
WPI:CW:Gly (1:1:1)3.1 ± 0.110 ± 1 a124 ± 29[329]
WPI control2.0 ± 0.1 a60 ± 5 a48 ± 3 a[331]
WPI:Gly:CAN 7.5 g:100 g0.6 ± 0.1 a65 ± 5 a45 ± 3 a[331]
WPI:Gly extruded2.4 ± 0.1 a90 ± 5 a30 ± 1 a[331]
WPI:Gly:CAN 7.5 g:100 g extruded1.7 ± 0.2 a55 ± 5 a20 ± 1 a[331]
WPI:Gly compressed extruded2.1 ± 0.1 a60 ± 5 a15 ± 1 a[331]
WPI:Gly:CAN 7.5 g:100 g compressed extruded1.5 ± 0.1 a30 ± 5 a10 ± 2 a[331]
WPI:TiO2 0% w/w6 ± 1.5 a13 ± 1 an/a[281]
WPI:TiO2 1% w/w8.25 ± 0.25 a10 ± 2 an/a[281]
WPI:TiO2 0 wt %1.69 ± 0.0355.56 ± 1.0531.44 ± 3.03[283]
WPI:TiO2 1 wt %2.19 ± 3.0340.11 ± 1.0163.09 ± 1.98[283]
WPI:TiO2 4 wt %1.78 ± 0.0812.14 ± 0.2239.23 ± 3.65[283]
WPI control3.40 ± 0.5850.9 ± 12.5171.8 ± 14.3[284]
WPI:Cloisite Na+2.98 ± 0.2942.4 ± 7.6109.3 ± 18.0[284]
WPI:Cloisite 30B3.29 ± 0.1051.7 ± 4.8162.6 ± 37.9[284]
WPI:Cloisite 20A1.55 ± 0.3229.1 ± 9.0115.5 ± 13.5[284]
WPI:ZNP 0 w/w2.5 ± 0.150 ± 10 an/a[285]
WPI:ZNP 0.4 w/w3.9 ± 0.1 a50 ± 5 an/a[285]
WPI:ZNP 1.2 w/w10.2 ± 0.335 ± 5 an/a[285]
WPI control1.902 ± 0.123126.8 ± 3.920.21 ± 2.26[308]
WPI:lactoperoxidase system 0.7% w/w2.090 ± 0.156129.5 ± 3.422.57 ± 2.57[308]
WPI:Sorbitol (50:50)3.3227.6083.96[314]
WPI:Sorbitol:BW (50:35:15)3.9112.5129.72[314]
WPI:Sorbitol:Potassium sorbate (50:43.3:6.7)2.6168.928.47[314]
WPI:Sorbitol:BW:Potassium sorbate (50:35:11.7:3.3)3.8632.890.27[314]
SPI:WG (2:1):pH 7.04.85 ± 0.20n/an/a[345]
SPI:WG (2:1): cys:pH 7.04.85 ± 0.20n/an/a[345]
SPI:WG (3:1):pH 7.05.68 ± 0.20n/an/a[345]
SPI:WG (3:1):cys:pH 7.06.37 ± 0.20n/an/a[345]
SPI:WG (4:1):pH 7.05.04 ± 0.20n/an/a[345]
SPI:WG (4:1):cys:pH 7.06.87 ± 0.20n/an/a[345]
SPI:WG (1:0) :pH 7.03.70 ± 0.20n/an/a[345]
SPI:WG (1:0):cys:pH 7.06.72 ± 0.22n/an/a[345]
SPI: DAS 0%6.34 ± 0.0265.9 ± 25.3n/a[340]
SPI: DAS 10%7.84 ± 0.1959.8 ± 19.9n/a[340]
SPI:10% fatty acid4.4 ± 0.670.1 ± 9.8n/a[343]
SPI:10% lauric acid1.9 ± 0.1125.8 ± 25.0n/a[343]
SPI:10% myristic acid0.4 ± 0.141.0 ± 5.4n/a[343]
SPI:10% palmitic acid1.2 ± 0.139.0 ± 15.3n/a[343]
SPI:10% oleic acid2.2 ± 0.2228.1 ± 14.2n/a[343]
SPI:Gelatin (10:0)5 a2.30 ± 1.0 a750 a[344]
SPI:Gelatin (8:2)25.552.64 ± 0.41280[344]
SPI:Gelatin (6:4)30.00 a3.1 ± 0.4 a1500.00 a[344]
SPI:Gelatin (4:6)35.00 a3.1 ± 0.8 a1550.00 a[344]
SPI:Gelatin (2:8)44.63.231861.16[344]
SPI:Gelatin (0:10)75 a3.50 ± 0.1 a3000 a[344]
SPI:BW:Span90.825.8n/a[341]
WG control2.6 ± 0.2237.9 ± 21.9n/a[337]
WG:Mineral oil2.2 ± 0.3267.2 ± 40.1n/a[337]
WG:MMT 0 wt %1.86 ± 0.4858.4 ± 7.03.73 ± 0.48[300]
WG:MMT 2.5 wt %2.37 ± 0.4455.4 ± 15.45.58 ± 1.04[300]
WG:MMT 5.0 wt %4.70 ± 0.8416.0 ± 12.410.58 ± 3.43[300]
WG:MMT 7.5 wt %3.60 ± 0.4215.2 ± 2.311.44 ± 1.77[300]
WG:MMT 10.0 wt %n/an/an/a[300]
WG controln/an/an/a[301]
WG:MMT 2.5% C0n/an/an/a[301]
WG:MMT 5.0% C0n/an/an/a[301]
WG:HEC 0%1.20 ± 0.1 a210 ± 15 a10 a[346]
WG:HEC 5%1.10 ± 0.05 a160 ± 20 a10 a[346]
WG:HEC 15%1.10 ± 0.05 a120 ± 10 a13 a[346]
WG:HEC 31.8%2.4 ± 0.05 a45 ± 5 a65 a[346]
Nanotechnology
SPI control1.10 ± 0.2065.95 ± 17.7626.89 ± 11.21[290]
SPI:SNC 5%1.34 ± 0.0758.67 ± 9.8839.42 ± 9.93[290]
SPI:SNC 20%2.61 ± 0.2641.89 ± 8.61102.23 ± 14.93[290]
SPI:SNC 40%5.08 ± 0.4821.35 ± 10.54310.34 ± 21.55[290]
SPI:Cloisite 20A 0%2.26 ± 0.4811.85 ± 0.39n/a[295]
SPI:Cloisite 20A 5%12.40 ± 0.6542.80 ± 0.57n/a[295]
SPI:Cloisite 20A 10%14.15 ± 0.3371.00 ± 3.68n/a[295]
SPI:Cloisite 20A 15%13.66 ± 0.2822.80 ± 1.70n/a[295]
SPI:Cloisite 30B 0%2.26 ± 0.4811.85 ± 0.39n/a[295]
SPI:Cloisite 30B 5%15.11 ± 0.8681.60 ± 2.83n/a[295]
SPI:Cloisite 30B 10%16.19 ± 0.75103.60 ± 4.53n/a[295]
SPI:Cloisite 30B 15%18.64 ± 0.2354.80 ± 2.26n/a[295]
SPI control2.26 ± 0.4811.85 ± 0.39n/a[295]
SPI:MMT 5%6.28 ± 0.8864.60 ± 4.69n/a[295]
SPI:MMT 10%12.62 ± 0.5423.98 ± 5.02n/a[295]
SPI:MMT 15%15.60 ± 1.6917.80 ± 2.27n/a[295]
SPI:MMT 0 wt %8.5 a90 a180 a[297]
SPI:MMT 4 wt %10.5 a65 a260 a[297]
SPI:MMT 8 wt %12 a30 a330 a[297]
SPI:MMT 12 wt %14 a17 a410 a[297]
SPI:MMT 16 wt %15 a8 a520 a[297]
SPI:MMT 20 wt %14.5 a5 a590 a[297]
SPI:MMT 0%3.0 a37.5 a42 a[298]
SPI:MMT 2.5%4.5 a32.5 an/a[298]
SPI:MMT 5%5.5 a27.5 an/a[298]
SPI:MMT 7.5%8 a20.0 an/a[298]
SPI:MMT 10%8.5 a6.5 an/a[298]
SPI control1.35 a60 an/a[299]
SPI:PNP 0.5%2.00 a85 an/a[299]
SPI:PNP 1.0%2.10 a86 an/a[299]
SPI:PNP 2.0%3.55 a90 an/a[299]
SPI:PNP 4.0%2.70 a82 an/a[299]
Antimicrobial Films
SPI control1.08 ± 0.3424.63 ± 0.1322.80 ± 6.14[318]
SPI:PLA (60:40)8.57 ± 1.611.09 ± 0.091085 ± 134[318]
SPI:PLA (50:50)13.69 ± 0.941.25 ± 0.021579 ± 52[318]
SPI control9.3115.7n/a[322]
SPI:Rutin35.173.5n/a[322]
SPI:Epicatechin22.138.5n/a[322]
Some data are converted to the same units; a Numerical value estimated from graph; MD Measured in Machine Direction; n/a not available.
Table 4. Barrier properties of protein-based films.
Table 4. Barrier properties of protein-based films.
Film ModificationWater Vapor Permeability (g·m·m−2·s−1 Pa−1)Relative Humidity (%)Oxygen Permeability (cm3·m−2·d−1·bar−1)Source
Heating
Heat treatment of solution
WPI control1.41 × 10−9 an/an/a[75]
WPI: 90 °C:30 min1.38 × 10−9 an/an/a[75]
WPI controln/an/a(6.83 ± 0.35) × 10−13[79]
WPI: 90 °C:30 minn/an/a(9.03 ± 0.35) × 10−13[79]
SPI control5.06 × 10−10n/an/a[84]
SPI: 85 °C4.48 × 10−10n/an/a[84]
Heat treatment of the film
SPI: 60 °C:24 h2.3 × 10−9 a0–50n/a[23]
SPI: 70 °C:24 h2.3 × 10−9 a0–50n/a[23]
SPI: 80 °C:24 h1.8 × 10−9 a0–50n/a[23]
WG control(1.37 ± 0.18) × 10−100–100n/a[27]
WG: 80 °C:15 min(1.22 ± 0.07) × 10−100–100n/a[27]
WG: 95 °C:15 min(1.13 ± 0.13) × 10−100–100n/a[27]
WG: 110 °C:15 min(1.19 ± 0.14) × 10−100–100n/a[27]
WG: 125 °C:15 min(1.37 ± 0.07) × 10−100–100n/a[27]
WG: 140 °C:15 min(1.15 ± 0.04) × 10−100–100n/a[27]
WG: 140 °C:1.5 min(1.21 ± 0.23) × 10−100–100n/a[27]
UV and γ-Irradiation
UV Irradiation of Solution
WPI control(4.49 ± 0.38) × 10−100–90(1.04 ± 0.06) × 10−7[173]
WPI: 324 J·cm−2(4.98 ± 0.24) × 10−100–90(9.17 ± 0.09) × 10−8[173]
UV Irradiation of the Film
WPI control349 ± 9 b50–0n/a[13]
WPI: 2.3 J·cm−2511 ± 11 b50–0n/a[13]
WPI: 10.2 J·cm−2434 ± 23 b50–0n/a[13]
WPI: 19.0 J·cm−2383 ± 40 b50–0n/a[13]
WPI: 31.4 J·cm−2386 ± 19 b50–0n/a[13]
SPI control(2.0 ± 0.2) × 10−950–100n/a[162]
SPI: 51.8 J·cm−2(2.5 ± 0.8) × 10−950–100n/a[162]
WG control(4.44 ± 0.14) × 10−950–100n/a[163]
WG: 51.8 J·cm−2(4.19 ± 0.08) × 10−950–100n/a[163]
WG control(1.37 ± 0.18) × 10−100–100n/a[27]
WG: 0.25 J·cm−2(1.42 ± 0.09) × 10−100–100n/a[27]
WG: 1 J·cm−2(1.35 ± 0.07) × 10−100–100n/a[27]
γ-Irradiation of Solution
WPI control3.5 × 10−10 a56–100n/a[177]
WPI: 32 kGy2.6 × 10−10 a56–100n/a[177]
WPC control3.0 × 10−10 a56–100n/a[177]
WPC: 32 kGy1.7 × 10−1056–100n/a[177]
γ-Irradiation of the Film
WG control(1.15 ± 0.02) × 10−100–100n/a[27]
WG: 10 kGy(1.37 ± 0.04) × 10−100–100n/a[27]
WG: 20 kGy(1.37 ± 0.09) × 10−100–100n/a[27]
WG: 40 kGy(1.21 ± 0.07) × 10−100–100n/a[27]
Chemical Modifications
Acylation
SPI0.84 × 10−9100/600.7 × 10−16[218]
SPI: acetylated0.86 × 10−9100/600.9 × 10−16[218]
SPI: succinylated0.84 × 10−9100/600.7 × 10−16[218]
Hydrolysis
WPI control1.29 × 10−9 a50 ± 11.2 × 10−12 a[185]
WPI: 5.5% DH1.25 × 10−9 a50 ± 11.4 × 10−12 a[185]
WPI: 10% DH1.35 × 10−9 a50 ± 11 × 10−12 a[185]
pH Alteration
WPI:pH 4 (pI)1.39 × 10−9 an/an/a[237]
WPI:pH 5 (pI)1.89 × 10−9 an/an/a[237]
WPI:pH 61.36 × 10−9 an/an/a[237]
WPI:pH 71.25 × 10−9 an/an/a[237]
WPI:pH 81.36 × 10−9 an/an/a[237]
SPI:pH 6 (pI)3.04 × 10−9n/a1.0 × 10−16[232]
SPI:pH 81.90 × 10−9n/a0.59 × 10−16[232]
SPI:pH 102.54 × 10−9n/a0.43 × 10−16[232]
SPI:pH 121.79 × 10−9n/a0.37 × 10−16[232]
Enzymatic Crosslinking
WPI control(7.53 ± 1.47) × 10−10100–50n/a[251]
WPI:TG(1.36 ± 0.17) × 10−9100–50n/a[251]
WPI:SPI 11S control(1.36 ± 0.15) × 10−9100–50n/a[251]
WPI:SPI 11S:TG(2.13 ± 0.17) × 10−9100–50n/a[251]
WPI control4.5 × 10−12 an/an/a[261]
WPI:TG3.4 × 10−12 an/an/a[261]
WPI:Chitosan 0.25 mg·cm−23.24 ± 0.150(2.38 ± 0.08) × 10−10[262]
WPI:Chitosan:TG0.88 ± 0.060(9.03 ± 0.93) × 10−11[262]
WPI controln/a502.11 × 10−9[268]
WPI:TG 10 units·g−1n/a503.47 × 10−10[268]
WPI control(7.53 ± 1.47) × 10−10100–50n/a[251]
WPI:TG(1.36 ± 0.17) × 10−9100–50n/a[251]
WPI:SPI 11S control(1.36 ± 0.15) × 10−9100–50n/a[251]
WPI:SPI 11S:TG(2.13 ± 0.17) × 10−9100–50n/a[251]
SPI 11S control(1.33 ± 0.25) × 10−9100–50n/a[251]
SPI 11S:TG(2.17 ± 0.16) × 10−9100–50n/a[251]
SPI control5.06 × 10−10n/an/a[84]
SPI:Horseradish Peroxidasen/an/an/a[84]
SPI:Gly control(3.44 ± 0.03) × 10−10100–0n/a[257]
SPI:Gly:mTG(3.69 ± 0.14) × 10−10100–0n/a[257]
SPI:Gly:Sorbitol control(2.94 ± 0.14) × 10−10100–0n/a[257]
SPI:Gly:Sorbitol:mTG(3.78 ± 0.03) × 10−10100–0n/a[257]
SPI:Sorbitol control(3.25 ± 0.14) × 10−10100–0n/a[257]
SPI:Sorbitol:mTG(3.64 ± 0.00) × 10−10100–0n/a[257]
SPI:Gelatin control(1.84 ± 0.03) × 10−10100–0n/a[274]
SPI:Gelatin:TG 10%(1.78 ± 0.03) × 10−10100–0n/a[274]
SPI:Gelatin:TG 20%(1.69 ± 0.04) × 10−10100–0n/a[274]
SPI:Gelatin:TG 30%(1.63 ± 0.03) × 10−10100–0n/a[274]
Composite Films and Bioactive Compounds
WPI control5.21 × 10−10 a100–0n/a[324]
WPI: 40% carnauba wax(3.82 ± 0.30) × 10−10100–0n/a[324]
WPI: 40% candelilla wax(3.59 ± 0.14) × 10−10100–0n/a[324]
WPI: 40% milk fat fraction(2.53 ± 0.50) × 10−10100–0n/a[324]
WPI:40% BW(1.25 ± 0.36) × 10−10100–0n/a[324]
WPI:Gly:BW (60:20:20):PS 0.5 µm(4.17 ± 0.69) × 10−10 a0n/a[325]
WPI:Gly:BW (30:10:60):PS 0.5 µm(2.77 ± 1.39) × 10−10 a0n/a[325]
WPI:Gly:BW (30:10:60):PS 1.0 µm(3.33 ± 0.56) × 10−10 a0n/a[325]
WPI:Gly:BW (30:10:60):PS 1.5 µm(3.89 ± 0.56) × 10−10 a0n/a[325]
WPI:Gly:BW (30:10:60):PS 2.0 µm(4.17 ± 0.83) × 10−10 a0n/a[325]
WPI control(3.83 ± 4.72) × 10−9100–50n/a[326]
WPI:Acetem(5.55 ± 2.78)× 10−11100–50n/a[326]
WPI control (2.00 ± 0.03) × 10−10100–0(1.30 ± 0.28) × 10−9[328]
WPI:Almond oil 0.5%(1.56 ± 0.06) × 10−10100–0(1.55 ± 0.14) × 10−9[328]
WPI:Almond oil 1.0%(1.27 ± 0.19) × 10−10100–0(1.81 ± 0.27) × 10−9[328]
WPI:Walnut oil 0.5%(1.32 ± 0.19) × 10−10100–0(1.32 ± 0.16) × 10−9[328]
WPI:Walnut oil 1.0%(1.02 ± 0.09) × 10−10100–0(1.52 ± 0.66) × 10−9[328]
WPI:Gly (1:1)(3.32 ± 0.25) × 10−1150–20n/a[329]
WPI:BW:Gly (1:1:1)(2.58 ± 0.23) × 10−1150–20n/a[329]
WPI:CW:Gly (1:1:1)(2.62 ± 0.28) × 10−1150–20n/a[329]
WPI control(1.94 ± 0.28) × 10−9 a0(3.18 ± 0.12) × 10−9 a[331]
WPI:Gly:CAN:7.5 g:100 g(1.94 ± 0.28) × 10−9 a0(3.30 ± 0.12) × 10−9 a[331]
WPI:Gly extruded(1.25 ± 0.06) × 10−8 a0n/a[331]
WPI:Gly:CAN:7.5 g:100 g extruded(9.72 ± 0.28) × 10−9 a0n/a[331]
WPI:Gly compressed extruded(3.61 ± 0.28) × 10−9 a0(2.89 ± 0.58) × 10−10 a[331]
WPI:Gly:CAN:7.5 g:100 g compressed extruded(2.28 ± 0.56) × 10−9 a0(3.47 ± 0.58) × 10−10 a[331]
WPI:TiO2 0% w/w(3.19 ± 0.08) × 10−10n/an/a[281]
WPI:TiO2 1% w/w(2.89 ± 0.11) × 10−10n/an/a[281]
WPI:TiO2 0 wt %(2.78 ± 0.28) × 10−9 a100–50n/a[282]
WPI:TiO2 1 wt %(2.42 ± 0.28) × 10−9 a100–50n/a[282]
WPI:TiO2 4 wt %(1.02 ± 0.83) × 10−9 a100–50n/a[282]
WPI control(66.0 ± 3.6) × 10−968.8 ± 1.3n/a[284]
WPI:Cloisite Na+(47.1 ± 3.3) × 10−972.2 ± 0.9n/a[284]
WPI:Cloisite 30B(55.6 ± 7.5) × 10−970.1 ± 0.8n/a[284]
WPI:Cloisite 20A(64.8 ± 4.2) × 10−971.1 ± 1.4n/a[284]
WPI:ZNP 0 w/w(8.94 ± 0.25) × 10−1125 ± 1n/a[285]
WPI:ZNP 0.4 w/w(3.42 ± 0.33) × 10−1126 ± 1n/a[285]
WPI:ZNP 1.2 w/w(1.44 ± 0.56) × 10−1227 ± 1n/a[285]
WPI controln/a50 ± 1(2.53 ± 0.16) × 10−9[308]
WPI:lactoperoxidase system 0.7% w/wn/a50 ± 1(2.56 ± 0.19) × 10−9[308]
WPI:Sorbitol (50:50)2.68 × 10−950–25n/a[315]
WPI:Sorbitol:BW (50:35:15)1.49 × 10−950–25n/a[315]
WPI:Sorbitol:Potassium sorbate (50:43.3:6.7)3.03 × 10−950–25n/a[315]
WPI:Sorbitol:BW:Potassium sorbate (50:35:11.7:3.3)2.34 × 10−950–25n/a[315]
SPI:WG (2:1):pH 7.00.50 × 10−9 a552.43 × 10−11[345]
SPI:WG (2:1):cys:pH 7.00.90 × 10−9 a552.55 × 10−11[345]
SPI:WG (3:1):pH 7.00.85 × 10−9 a550.61 × 10−11[345]
SPI:WG (3:1) cys:pH 7.00.75 × 10−9 a551.13 × 10−11[345]
SPI:WG (4:1):pH 7.00.50 × 10−9 a551.57 × 10−11[345]
SPI:WG (4:1) cys:pH 7.00.55 × 10−9 a551.39 × 10−11[345]
SPI:WG (1:0):pH 7.02.7 × 10−9 a550.67 × 10−11[345]
SPI:WG (1:0) cys:pH 7.02.5 × 10−9 a550.57 × 10−11[345]
SPI:DAS 0%(15.0 ± 0.2) × 10−1073.1 ± 0.3n/a[340]
SPI:DAS 10%(16.5 ± 0.89) × 10−1072.7 ± 0.5n/a[340]
SPI:10% fatty acid(28 ± 0.94) × 10−10100–50n/a[343]
SPI:10% lauric acid(17.5 ± 1.11) × 10−10100–50n/a[343]
SPI:10% myristic acid(20.6 ± 0.42) × 10−10100–50n/a[343]
SPI:10% palmitic acid(22.4 ± 1.44) × 10−10100–50n/a[343]
SPI:10% oleic acid(18.5 ± 0.69) × 10−10100–50n/a[343]
SPI:BW:Span2.22 × 10−10540[341]
WG:Gly (83.3:16.7)(9.5 ± 0.9) × 10−11100–0n/a[3]
WG:Gly:BW (66.7:16.7:20)(3.5 ± 0.17) × 10−11100–0n/a[3]
WG:Gly:carnauba wax (66.7:16.7:20)(6.51 ± 0.43) × 10−11100–0n/a[3]
WG:Gly:refined paraffin (66.7:16.7:20)(6.9 ± 0.43) × 10−11100–0n/a[3]
WG:Gly:oleic acid (66.7:16.7:20)(7.8 ± 0.439) × 10−11100–0n/a[3]
WG:Gly:soy lecithin (66.7:16.7:20)(1.0 ± 0.1) × 10−10100–0n/a[3]
WG:Gly:acetic ester of monoglyceride (66.7:16.7:20)(1.2 ± 0.1) × 10−10100–0n/a[3]
WG:Gly:diacetyl tartaric ester of monoglyceride (66.7:16.7:20)(4.3 ± 0.1) × 10−11100–0n/a[3]
WG:Gly:sucroglyceride (66.7:16.7:20)(5.64 ± 0.17) × 10−11100–0n/a[3]
WG:Gly:stearic alcohol (66.7:16.7:20)(4.34 ± 0.17) × 10−11100n/a[3]
WG:MMT 5.0 wt %6.5 × 10−12 a1008.0 × 10−12 a[300]
WG:MMT 7.5 wt %7.0 × 10−12 a1008.1 × 10−12 a[300]
WG:MMT 10.0 wt %6.0 × 10−12 a1008.3 × 10−12 a[300]
WG control(3.0 ± 0.1) × 1011n/an/a[301]
WG:MMT 2.5% C0(2.2 ± 0.1) × 1011n/an/a[301]
WG:MMT 5.0% C0(1.8 ± 0.1) × 1011n/an/a[301]
Nanotechnology
SPI control(4.3 ± 0.2) × 10−1065n/a[290]
SPI:SNC 5%(4.8 ± 0.3) × 10−1065n/a[290]
SPI:SNC 20%(3.9 ± 0.1) × 10−1065n/a[290]
SPI:SNC 40%(3.57 ± 0.08) × 10−1065n/a[290]
SPI:Cloisite 20A 0%(10.56 ± 0.31) × 10−10100–65n/a[295]
SPI:Cloisite 20A 5%(7.31 ± 0.08) × 10−10100–65n/a[295]
SPI:Cloisite 20A 10%(6.00 ± 0.28) × 10−10100–65n/a[295]
SPI:Cloisite 20A 15%(5.69 ± 0.31) × 10−10100–65n/a[295]
SPI:Cloisite 30B 0%(10.56 ± 0.31) × 10−10100–65n/a[295]
SPI:Cloisite 30B 5%(8.58 ± 0.14) × 10−10100–65n/a[295]
SPI:Cloisite 30B 10%(7.42 ± 0.22) × 10−10100–65n/a[295]
SPI:Cloisite 30B 15%(6.47 ± 0.25) × 10−10100–65n/a[295]
SPI control(10.56 ± 0.31) × 10−10100–65n/a[295]
SPI:MMT 5%(8.22 ± 0.28) × 10−10100–65n/a[295]
SPI:MMT 10%(6.92 ± 0.22) x 10−10100–65n/a[295]
SPI:MMT 15%(6.03 ± 0.17) x 10−10100–65n/a[295]
SPI:MMT 0%(12 ± 0.8) × 10−1175n/a[298]
SPI:MMT 2.5%(11 ± 0.3) × 10−1175n/a[298]
SPI:MMT 5%(6.8 ± 0.2) × 10−1175n/a[298]
SPI:MMT 7.5%(5.0 ± 0.6) × 10−1175n/a[298]
SPI:MMT 10%(3.2 ± 0.9) × 10−1175n/a[298]
SPI control(1.21 ± 0.04) × 10−6100n/a[299]
SPI:PNP 0.5%(1.06 ± 0.04) × 10−6100n/a[299]
SPI:PNP 1.0%(1.00 ± 0.03) × 10−6100n/a[299]
SPI:PNP 2.0%(0.92 ± 0.02) × 10−6100n/a[299]
SPI:PNP 4.0%(0.83 ± 0.01) × 10−6100n/a[299]
Antimicrobial Films
SPI control(14.9 ± 0.5) × 10−1165n/a[318]
SPI:PLA (60:40)(3.4 ± 0.1) × 10−1165n/a[318]
SPI:PLA (50:50)(2.3 ± 0.1) × 10−1165n/a[318]
SPI control(0.47 ± 0.03) × 10−354n/a[322]
SPI:Rutin(0.33 ± 0.06) × 10−354n/a[322]
SPI:Epicatechin(0.64 ± 0.06) × 10−354n/a[322]
Some data are converted to the same units. a Numerical value estimated from graph; b Water Vapour Transmission Rate in g·m−2·d−1; n/a not available.

Share and Cite

MDPI and ACS Style

Zink, J.; Wyrobnik, T.; Prinz, T.; Schmid, M. Physical, Chemical and Biochemical Modifications of Protein-Based Films and Coatings: An Extensive Review. Int. J. Mol. Sci. 2016, 17, 1376. https://doi.org/10.3390/ijms17091376

AMA Style

Zink J, Wyrobnik T, Prinz T, Schmid M. Physical, Chemical and Biochemical Modifications of Protein-Based Films and Coatings: An Extensive Review. International Journal of Molecular Sciences. 2016; 17(9):1376. https://doi.org/10.3390/ijms17091376

Chicago/Turabian Style

Zink, Joël, Tom Wyrobnik, Tobias Prinz, and Markus Schmid. 2016. "Physical, Chemical and Biochemical Modifications of Protein-Based Films and Coatings: An Extensive Review" International Journal of Molecular Sciences 17, no. 9: 1376. https://doi.org/10.3390/ijms17091376

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop