Next Article in Journal
A Pandemic Early Warning System Decision Analysis Concept Utilizing a Distributed Network of Air Samplers via Electrostatic Air Precipitation
Next Article in Special Issue
Assessment of Manufacturing Parameters for New 3D-Printed Heating Circuits Based on CNT-Doped Nanocomposites Processed by UV-Assisted Direct Write
Previous Article in Journal
Mean Dynamic Topography Modeling Based on Optimal Interpolation from Satellite Gravimetry and Altimetry Data
Previous Article in Special Issue
Machine Learning-Based Prediction and Optimisation System for Laser Shock Peening
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Can Sustainable Packaging Help to Reduce Food Waste? A Status Quo Focusing Plant-Derived Polymers and Additives

1
Institute for Animal Sciences, Rheinische Friedrich Wilhelms-University Bonn, Katzenburgweg 7-9, D-53115 Bonn, Germany
2
Department of Fresh Produce Logistics, Hochschule Geisenheim University, Von-Lade-Straße 1, D-65366 Geisenheim, Germany
3
Departement of Social Sciences, Wageningen University, P.O. Box 8130, Bode 55, 6700 EW Wageningen, The Netherlands
4
Institue for Food and Resource Economics, Rheinische Friedrich Wilhelms-University Bonn, Meckenheimer Allee 174, D-53115 Bonn, Germany
5
Institute of Crop Science and Resource Conservation, INRES-Renewable Resources, Rheinische Friedrich Wilhelms-University Bonn, Campus Klein-Altendorf 1, D-53359 Rheinbach, Germany
6
Department of Chemistry, MS 015, Brandeis University, 415 South Street, Waltham, MA 02453, USA
7
Department of Natural Sciences, Institute of Technology, Resource and Energy-Efficient Engineering (TREE), Bonn-Rhein-Sieg University of Applied Sciences, von-Liebig-Strasse 20, D-53359 Rheinbach, Germany
*
Authors to whom correspondence should be addressed.
Submission received: 13 April 2021 / Revised: 25 April 2021 / Accepted: 4 June 2021 / Published: 7 June 2021
(This article belongs to the Special Issue 10th Anniversary of Applied Sciences: Invited Papers in Materials)

Abstract

:

Featured Application

Exploitation of renewable resources and plant extracts for the development of sustainable materials used for active food packaging.

Abstract

The promotion of sustainable packaging is part of the European Green Deal and plays a key role in the EU’s social and political strategy. One option is the use of renewable resources and biomass waste as raw materials for polymer production. Lignocellulose biomass from annual and perennial industrial crops and agricultural residues are a major source of polysaccharides, proteins, and lignin and can also be used to obtain plant-based extracts and essential oils. Therefore, these biomasses are considered as potential substitute for fossil-based resources. Here, the status quo of bio-based polymers is discussed and evaluated in terms of properties related to packaging applications such as gas and water vapor permeability as well as mechanical properties. So far, their practical use is still restricted due to lower performance in fundamental packaging functions that directly influence food quality and safety, the length of shelf life, and thus the amount of food waste. Besides bio-based polymers, this review focuses on plant extracts as active packaging agents. Incorporating extracts of herbs, flowers, trees, and their fruits is inevitable to achieve desired material properties that are capable to prolong the food shelf life. Finally, the adoption potential of packaging based on polymers from renewable resources is discussed from a bioeconomy perspective.

1. Introduction

The main functions of food packaging are protection/preservation, containment, communication/marketing, and convenience. Thereby, food safety and quality related properties as well as reducing food waste are targeted [1,2,3]. With an appropriate packaging solution, a high quality, safe (extended) shelf life along the entire supply chain can be ensured. Thus, possible waste of the food product prior to final consumption can be prevented [1,3,4,5,6].
There are different packaging strategies to achieve a longer shelf life. One of the main strategies is the application of materials with certain barrier functions like gas and water vapor permeability (WVP) guaranteed by various fossil-based materials [7,8] (for definition of terms such as fossil-based, bio-based, etc., see glossary at the end of the main body text). Many of these fossil-based foils are complex multi-layer materials that represent a large group of individual fossil-based polymers with different chemical and technical characteristics due to diverse requirements regarding food safety and waste reduction [8,9,10]. Another strategy is the use of active packaging [2,11]. Based on the European Union (EU) Guidance to the Commission Regulation No 450/2009, active materials are defined as: “[…] materials […] that are intended to extend the shelf life or to maintain or improve the condition of packaged food; they are designed to deliberately incorporated components that would release or absorb substances into or from the packaged food or the environment surrounding the food.” [12].
Today, around 40% of all plastics circulating are applied for packaging [13], of which approximately 60% are used for food and beverages while the rest covers non-food applications [14]. In Germany, the packaging consumption of glass reached 35.0 kg/head, for paper 98.5 kg/head and for plastic 39.0 kg/head in 2018 [15]. Plastic waste represents one of the most complex material mixtures from a recycling perspective [16]. Moreover, there are increasing issues concerning the harm caused to the environment mainly due to the manufacturing phase (i.e., oil refinery and material production), problematic end-of-life strategies, and adverse effects on human health [5,14]. Despite the negative environmental and health-related effects, fossil-based plastics are favored because of their lightweight nature and low costs [8,9]. As a result, the use of plastic packaging is growing [17] caused by the need to reduce food waste due to the steadily growing population and market expansion [18]. To deal with the issue of adverse impact of fossil-based plastics on human health and environment, the design and production of plastics and plastic products must take account of the end-of-use strategy such as reuse, repair, and recycling needs [14]. This leads to a paradigm shift from linear to circular economy. The core principles “take, make, dispose” of a linear economy are replaced by “take, make, reuse” in a circular economy. “Reuse” involves circular criteria like repair, refurbish, and recycle, as recently reported by Taleb et al. [19].
Within the last decade, the development and promotion of more sustainable materials became key roles on social and political levels in the EU [14]. Pursuing these strategies, the European Commission adopted a Circular Economy Action Plan in 2015 [20]. This action plan lays the foundation for a new plastics economy addressing the environmental issues concerning plastics and forcing the EU to move towards a more sustainable model for economic development [21]. The EU launched and developed the action plan in 2018 with the “European Strategy for Plastics in a Circular Economy”, the so called plastics strategy [14]. The plastics strategy forces the industry to rethink plastics design and their usage, disposal, and recycling within the entire value chain. The main goal is to achieve improvements in sustainability [8].
To reduce environmental impacts, one key requirement is the complete reusability and/or recyclability of all plastic packaging placed on the EU market by 2030 [14]. Moreover, Matthews et al. stated that innovations in food packaging have to focus on maintaining food safety and shelf life and reducing food waste [8].
Besides current political requirements, sustainable packaging is also an important aspect for consumers [14,22]. As consumer preferences shifted to high quality and safe products with enhanced shelf life, the development of various new trends in packaging systems has arisen [1]. There are different strategies for the development of sustainable packaging such as reducing packaging materials by decreasing the thickness and/or number of layers, applying bio-based and/or biodegradable materials, reducing the amount of layers, and using easy recyclable materials [22]. For example, material properties of simple mono-layer materials can be improved by design of packaging itself such as introducing packaging conditions through a modified atmosphere or by using active packaging solutions [23]. Recently, Pauer et al. reported that the environmental benefit of weight reduction is greater than the benefit from improved recyclability in terms of meat packaging [24].
Therefore, bio-based and biodegradable polymers represent a growing field in creating environmentally friendly materials [25]. The continued use of agricultural and industrial by-products and waste flows such as corn stover, wheat straw, and whey constitutes from dairy and cheese industries as raw materials would provide a significant ecological advantage and would reduce pressure on land use [26,27,28]. Plant-based (waste) materials such as wood and lignocellulosic residues from agriculture and forestry are a major source of polysaccharides; therefore, they are considered as sustainable alternatives. They have the potential to be used instead of fossil resources [29,30]. Concerns about greenhouse gas (GHG) emissions and the security of industrial feedstock supplies promote substituting conventional fossil-based feedstock in the production of synthetic materials with biomass [31]. Several studies considering cradle-to-grave life cycle analysis (LCA) of various bio- and fossil-based plastics showed that the production and use of plastics produced from renewable resources is generally advantageous in terms of saving fossil resources and reducing GHG emissions [28,32,33]. Biodegradable or compostable plastics can reduce the amount of waste sent to landfills [1,8]. Natural bio-based polymers such as agar, chitosan, cellulose, and starch represent the group of bio-based polymers in food packaging applications [34].
Although significant effort is currently being made to develop novel, sustainable materials, there are currently no competitive alternatives which offer the same level of protection to fossil-based multi-layer plastic packaging, especially for fresh products like meat [8]. Although research focuses on improving the bio-based film characteristics of packaging materials, their mechanical, thermal, and physical properties are still non-satisfactory, and their use in industrial applications is often restricted [2]. To counteract these disadvantages by replacing fossil-based plastic packaging, research, and developments include active packaging based on bioactive polymers and composites obtained from renewable resources [1,35]. There are several developments in this field, but up to now, a widespread use in the market is still missing. Today, a limited number of bio-based plastics with food packaging applications are commercially available mainly based on the following polymers: poly(lactic acid) (PLA), poly(hydroxyalkanoates) (PHA), pol(yethylene furanoate) (PEF), poly(butylene succinate) (PBS), thermoplastic cellulose derivatives, and starch-based films [1].
This review focuses on most recent developments of packaging materials (polymers and composites) that are produced from renewable resources and considered as promising alternatives for fossil-based plastics. Annual and perennial crops, herbs, flowers, lignocellulosic and agricultural residues are a major source of polysaccharides, proteins, and lignin. In addition, they can also be used to obtain plant extracts and essential oils (EO). Therefore, we consider their application in terms of sustainable packaging that contribute to the reduction of food waste. First, challenges along food supply chains related to packaging characteristics are discussed: advantages and disadvantages related to the characteristics of the most prominent fossil-based plastics packaging are compared. They are evaluated regarding food waste, recyclability, and sustainability issues (Section 2). Second, characteristics of selected packaging plastics (mainly plant-derived polymers) in terms of mechanical, thermal, and physical properties are highlighted. Biodegradability will not be addressed in detail but mainly properties, such as gas permeability with influence on the food quality, safety, and shelf life (Section 3). Third, natural additives with focus on plant extracts in the context of active food packaging are presented (Section 4). Incorporating extracts of herbs, flowers, trees, and their fruits is inevitable to achieve desired material properties that are capable to prolong shelf life, resulting in reduced food waste. Finally, in Section 5, the potential of renewable resources is evaluated from a bioeconomy perspective of a packaging core matrix as well as a source for natural additives for active packaging.

2. Challenges in Food Supply Chains Related to Packaging Characteristics

2.1. Current Packaging Characteristics

While bio-based systems are the focus of the review, we begin by discussing general issues of packaging and the reasons behind the current ubiquitousness of fossil-based food packaging. Intensive research and development efforts over a long period of time resulted in packaging solutions which are optimized with regard to various important parameters that guarantee their functionality including material weight per unit of packed volume [8,36], mechanical characteristics (impact and tear strength, stiffness, flexibility), durability, and many others [2,5,9,17,37,38]. Moreover, fossil-based plastics are less expensive per weight unit compared to most of the bio-based materials. Therefore, currently, alternative bio-based materials are more expensive because of poor commercial availability and lack of efficient production processes. In the future, the prices for fossil resources will rise due to the limited availability, and on the other hand, costs for bio-based plastics may drop due to improvements in production process efficiency [39]. The current focus of research activities is the reduction of packaging waste especially fossil-based material [5]. LCA showed that a reduction in environmental impact of the packaging itself can be best achieved by minimization of used materials (thinner layers) that retain mechanical and barrier properties rather than emphasizing end-of-life issues (such as recycling or disposal) [7].
Today, multi-layer materials are widely used for food packaging throughout the food industry. As multi-layers of plastic can be easily adjusted to the various requirements of different food types, this kind of packaging can offer effective solutions to maintain food safety and quality, achieve optimal shelf life, and minimize food waste [8]. Although from an environmental, sustainability, and biocompatibility perspective, the use of multi-layered packaging materials has to be reduced, their global use for food applications is growing. This development is based on the mechanical and barrier properties of multi-layer materials (higher resistance to water, gas, and aroma transfer) [1,7,8]. Especially, the barrier against oxygen is a key factor for food quality and safety [7]. According to McMillin, appearance, color, lipid stability, nutritive value, and palatability (texture, flavor, aroma) are significant factors that must be considered when choosing a packaging solution [40]. Changing to biobased packaging materials, these factors have to be considered to ensure the shelf life and quality of the food [41,42].
So far, a broad variety of bio-based materials have been investigated to meet the purpose and achieve the properties of commercial packaging plastics. Although many new materials are used, companies worry about their physical inferiority compared to conventional polymers. Currently, European legislation and regulations are forcing companies to eliminate or reduce conventional plastic in packaging (COM/2018/028 final) [8]. Even if there is a desire to change, there is a conflicting pressure that prohibits changes in packaging because of different attitudes along the supply chain. The companies are faced with a challenge of alternatives offering higher costs and lower functionality, existing infrastructure, and inconsistent legislation [43,44].
Today, many companies are trying to use materials that are more recyclable instead of using bio-based materials. According to the adopted “European strategy for plastics in a circular economy” (COM/2018/028 final) where sustainability is the underlying motivation, recycling of plastic packaging is a key factor [8]. Recycling is viewed as the primary mechanism to reduce the environmental and waste management issues that are related especially to the use of conventional plastic [45]. Therefore, sorting and recycling capacities have to be expanded and modernized. Industries are investing in research and innovation activities to develop new technologies that support and increase the recovery of plastic packaging material [14,38,46].
As recently reported, in a circular economy, resources, materials, and products remain as long as possible on the market, minimizing waste and resources. This results in major economic benefits, innovation, and growth. However, safe disposal and recycling of materials often remain challenging. Reasons are poor management and enforcement, regulatory disparities, lack of infrastructure, and high cost of waste recycle systems [19].
Although many polymers are recyclable, due to additives and related quality issues, recycling rates remain low [38]. Currently, the recycling rate of packaging waste in the EU reaches 67% in total [47], 42% of plastic (packaging waste) [48], and 72% of paper and cardboard [49], respectively. Until 2025, at least 65% by weight of all packaging waste has to be recycled. Regarding specific materials contained in packaging waste, 50% of plastics and 75% of paper and cardboard are the target rates for recycling. By 2030 the recycling rate of all packaging waste rises to 70%, and for plastic and paper and cardboard the targets are 55% and 85%, respectively [50].
Moreover, the multi-layer structure makes plastic waste one of the most complex material mixture from a recycling perspective [16]. Recycling of these materials is accompanied by either high costs, technical difficulties regarding the separation process of the different polymers, or the inability to recycle mixed polymers [8,45]. Current recycling technologies for processing and handling solid plastic waste streams include gasification as a thermo-chemical conversion process for the recycling of polymeric composites, pyrolysis, fluid catalytic cracking, hydrogen technologies (hydrocracking and IH2 process), and the catalytic pressure-less depolymerization process [16,51,52,53,54,55].
The purpose of future redesign and waste management is the reduction of the amount of plastic that is accumulated in the environment and disposed on landfills, especially in developing countries [56,57]. So, further innovations in both recyclable packaging designs and corresponding cost-effective technologies are needed—independent of the material origin (natural or artificial polymers) [8].

2.2. Food Waste and the Meaning of Packaging

Food wastage and loss describe a major concern in the food supply chain that takes all of the involved stakeholders into consideration [58]. There is an increasing concern about the amount of food waste in Europe as wasted food has a significant impact on the use of natural resources and the environment [59]. The United Nations set a target of halving the actual amount of global food waste per capita at retail and consumer levels and reducing food losses along production and food supply chains as part of their sustainable development goals for 2030 (goal 12.3 of the UN General Assembly) [59].
Differentiating between avoidable (edible) food waste and unavoidable (non-edible) food waste, proper waste management, and recycling strategies is required to reduce unavoidable food waste. Various chemical and biological processes can be used to convert food waste into bio-commodity chemicals and bio-energy [60].
Packaging prevents avoidable food waste and has the potential to further decrease it [42]. According to a study of Bruckner et al., the shelf life of poultry under aerobe packaging conditions at 4 °C accounts for 98.6 h [61]. At 4 °C, the shelf life of poultry packed under modified atmosphere packaging (70% oxygen (O2)/30% carbon dioxide (CO2)) is prolonged to 228 h [62]. The kind of packaging has a high impact on the shelf life of poultry and can more than double it.
Caldeira et al. focused on food waste generated in the EU for the major food groups: sugar beets, oil crops, potatoes, vegetables, fruit, cereals, meat, fish, dairy, and eggs. The food waste generated at each stage of the food supply chain was quantified. In total around 638 mega tons (Mt) primary foods result in approximately 129 Mt of food waste generated along the food supply chain. Fruit and vegetables were the food groups presenting the highest amount of food waste overall, with similar amounts generated at the primary production and consumption stages [63]. Products of food categories with a relatively short shelf life, like fresh meat, tend to be the most wasteful products [64,65].
The amount of wasted food means that not only the products themselves are lost but also a high amount of primary resources of fuel, land, and water, including resources needed for breeding and fattening of animals, for cultivation of plants, and raw materials for processing and packaging during production as well as along the entire supply chain [65,66]. Therefore, for a packaging system, it is important to find a balance between the environmental impact of the package itself, on the one hand, and the impact originating from the potential loss of the packaged product, on the other [66,67].
Considering the environmental impact, this is much higher for producing the food itself than the (multi-layer) plastic packaging. Therefore, if food waste occurs, the negative overall environmental impact rises with every step in the supply chain due to more used resources [8,66,68]. An analysis of the food supply chain and the points where food waste is generated showed that reducing packaging is important, but it must still fulfil its duty of protection as the main criteria for sustainability. Otherwise, the supply chain overall will be less sustainable [24,68,69].
Steinbuchel reported that the production of starch plastic granules requires 25–75% less energy, and GHG emissions are reduced about 20–80% compared to poly(ethylene) (PE) [70]. Weiss et al. reviewed the environmental impacts of bio-based materials in a meta-analysis of LCA data. Therefore, one metric ton (t) of bio-based materials saves 55 ± 34 gigajoules (GJ) of primary energy and 3 ± 1 t carbon dioxide equivalents of GHG emissions relative to conventional materials [31].
Conte et al. assessed the environmental impact of single-layer and multi-layer conventional packaging. The results show that multi-layer surpass single-layer materials by environmental impact when food waste is included in the system boundaries [71].
Pettersen et al. studied the possibility of packaging chicken fillets in recyclable mono-layer materials (high-density polyethylene (HDPE)) instead of complex multi-layered materials (amorphous polyethylene terephthalate (APET)/(PE)) as a replacement for more sustainable packaging system without decreasing the quality of fresh chicken fillets. The results show that a competitive quality and shelf life can be obtained [72].
Zhang et al. focused on a case study, based on LCA data, where the ability of active packaging to minimize food losses by using thymol/carvacrol-enabled active packaging for fresh beef was investigated. Different scenarios have been considered in terms of overall environmental performance of the food and packaging system, including the effect of food loss reduction by using active packaging. It was shown that a breakeven point can be achieved considering the evaluated impact categories in the scenario using the best-performance active packaging whereas differences were observed between the impact categories. The breakeven point can be achieved as early as 0.1% food loss elimination occurs, whereas in the case of cumulative energy demand (fossil), it required more effort to reach the breakeven point. In this case, the active packaging performance needs to reduce food losses at least by 0.6% [73].

3. Plastics Used for Food Packaging

Within the last decades, a broad variety of polymers prepared from renewable resources have been studied as potential substitute for conventional packaging plastics [44]. The European Committee of the Regions stated that further research on the relation between packaging and food preservation on a life cycle basis is needed, and that possible alternative approaches to prevent food waste without the use of fossil-based (complex) plastics has to be investigated [74]. Increasing the exploitation efficiency of natural resources plays an important part on the way to a circular economy [75]. Next to circularity, sustainable packaging should be safe for the environment and humans [76]. The idea of biodegradable polymers, particularly obtained from renewable sources, stems from the need to close the natural cycle of matter [77]. Bio-based applications can be a useful replacement considering the biodegradability, biocompatibility, and recyclability [45,78,79].
Different studies show a limited but growing number of natural polymers used as films and coatings applied for food packaging [80,81]. So far, their practical use depending on the material is restricted due to lower performance in fundamental packaging functions. Challenges such as relatively poor thermal, mechanical, and rheological properties; higher costs; lack of compatibility with the processing and recycling systems currently available; or perceived environmental issues of natural polymers must be overcome. In addition, the barrier properties of natural polymers, especially the moisture barrier properties due to the hydrophilic nature of these polymers, are detrimental to existing packaging materials [3,82,83,84].
The WVP is important for fresh foods where dehydration and absorption of moisture should be avoided [85,86]. In general, the water vapor permeability is affected by several factors: chemical structure of macromolecules, degree of cross-linking, crystallinity and porosity, comparative humidity, and the addition of a plasticizer [87]. Oxygen permeability is another fundamental parameter of food packaging material. Low values in oxygen permeability are aimed to prevent deterioration in food quality [88].
To contribute to the reduction of packaging waste by preservation of fresh foods and to enhance their applications, currently, most natural polymers are mixed or blended with synthetic compounds such as PLA, poly(caprolactone) (PCL), and poly(hydroxybutyrate) (PHB). Furthermore, lightweight polysaccharide-based nanomaterials that could replace traditional plastic packaging are shown to improve antimicrobial activity, thermal, mechanical, and gas barrier properties while retaining the biodegradable and non-toxic characteristics of polysaccharides such as chitosan, carboxymethyl cellulose (CMC), and starch [37,83,89,90,91,92,93].

3.1. Classification of Plastics

Currently, the food packaging industry depends on fossil-based plastics which in turn originate from a finite raw material feedstock [8]. The finite resources issue induces a movement towards reducing the usage of virgin plastics towards a plastic production based on alternative raw materials such as renewable resources and biomass waste that have the potential to become plastic alternatives [44,94].
Next to conventional plastics that are fossil-based and non-biodegradable (e.g., PE, PP, poly(ethylene terephthalate) (PET), “bioplastics” were developed. According to the European Bioplastics association, “a plastic material is defined as a bioplastic if it is either bio-based, biodegradable, or features both properties” [95]. Thus, bioplastics involve a range of materials that show different properties and applications [1]. Figure 1 illustrates the categories of the plastics used for food packaging applications.
Thus, biodegradability (and even more compostability) is considered as a useful characteristic providing one option to reduce plastic waste. Biodegradation occurs when a product undergoes a significant change in chemical structure under specific environmental conditions. Biodegradable polymers can, for example, be decomposed to natural substances (CO2 or methane (CH4) and water (H2O)) by microorganisms that are found in the environment like algae, fungi, and bacteria [1,8,89,96]. As biodegradation depends on the chemical structure of the material compound rather than on its origin, the basis of biodegradable plastics are not necessarily renewable resources [1,97,98].
Biopolymers such as proteins, polysaccharides, and lignin are natural polymers produced by the cells of living organisms (e.g., forestry and agricultural crops, terrestrial and marine animals) [29,30]. These biopolymers can be used for the manufacturing of packaging materials and therefore have a high potential to replace synthetic plastics [99,100]. Most common biopolymers currently used for food packaging applications are synthesized [34,83,101,102].
In the following subchapters, the review focuses on different groups of bio-based polymers and their characteristics regarding food packaging applications starting with synthetically manufactured polymers using natural monomers (Section 3.2.) followed by polymers isolated from renewable resources (Section 3.3).

3.2. Selected Biodegradable Synthetically Manufactured Polymers

3.2.1. Biomass-derived Chemically Manufactured Polymers

One commercially manufactured example should be discussed here: poly (lactic acid) (PLA) and related products. PLA is a thermoplastic and biodegradable aliphatic polyester (Figure 2). As it has its roots in the aliphatic class of polyesters, PLA can be created either by chemical processing of lactic acid monomer or by fermentation of a carbohydrate [102].
PLA is the first polymer synthesized from bio-based monomers commercialized on a large scale and can be shaped into injection molded objects, films, coatings, and 3D printed materials [103]. Next to PHAs, starch, and PCL, PLA is the primary biodegradable polymer used for monolayer and some multilayer applications [83,104]. So, PLA films are applied as thermoformed trays, cups, bowls, bags, or jars for packaging of fresh salads, ready-to-eat meals, deli products, beverages, potato chips, and yoghurt among other uses [105,106]. Renewability, biodegradability, and biocompatibility are attributes that make PLA one of the best polymeric substitutes for various fossil-based polymers [83]. However, the PLA synthesis—and in turn the corresponding products—are still rather expensive [39].
So far, PLA and corresponding copolymers are used to substitute polyolefins as high-density poly(ethylene) (HDPE), low-density poly(ethylene) (LDPE), poly(propylene) (PP), PET, and poly(styrene) (PS) as packaging materials [37,83] due to comparable mechanical properties like stiffness and tensile strength, gas permeability, and transparency (Table 1) [107,108,109].
The production of composites by adding nanofillers is a way to extend and improve the properties of PLA [83,104,109]. The addition of many nanofillers (three-dimensional spherical and polyhedral, two-dimensional nanofibers or one-dimensional sheet-like nanoparticles) has been studied and lead to satisfactory achievements in the design of PLA nanocomposites [116].
Panseri et al. studied the effectiveness of PLA-based packaging solutions compared to a conventional reference package consisting of APET/PET trays wrapped in plastic films of poly(vinyl chloride) (PVC) to store red fresh meat during its refrigerated shelf life. By using PLA packaging in combination with a gas mixture of 66% O2, 25% CO2, and 9% N2, it was possible to maintain an optimum red color together with a reduced content of volatile compounds associated to off-flavors of meat samples [117].
Marra et al. investigated biocomposite films of PLA with zinc oxide regarding mechanical, barrier, and antimicrobial properties. The results showed that PLA films with 5 wt% of zinc oxide exhibit good mechanical properties related to a high modulus and stress at yielding, decrease of permeability to CO2 and O2 carbon dioxide and oxygen, and a slight increase of water vapor permeability. Furthermore, the incorporation of 5% zinc oxide leads to an antimicrobial activity against E. coli after 24 h with a reduction value of 99.99% [37]. Vanitha and Kavitha incorporated cellulose natural fibers from palm sprouts in a PLA matrix. The results showed that the mechanical resistance increased, and the water absorption rate decreased significantly with the optimum concentration of palm sprouts fiber in the PLA-film. Interactions between palm sprout and PLA restrict the water infiltration [118].
The incorporation of lignin in PLA films via simple blending results in a small but significant increase of the oxygen barrier properties, as well as an improved antiradical efficiency that increases with the severity of the heat treatment of the blends [119]. Moreover, the water sorption capacity decreased with an increase of lignin loading from 7 wt% to 15 wt% while tensile strength increased, as shown in a study of Spiridon et al. [120]. Gordobil et al. used commercial alkaline lignin and organosolv lignin from almond shells as PLA filler, which greatly improved the thermal stability and increased the elongation at break. Low percentages up to 1% unmodified lignin did not affect the maximum strain, while it was decreased with increasing lignin content at percentages greater than 5% [121]. In addition, kraft and organosolv lignin were examined as nucleating agents, showing that both lignins induce heterogeneous nucleation and increase the crystallization rate in PLA by shortening the crystallization half time and increasing the degree of crystallinity in PLA, while not affecting the processing window of the polymer [122]. One problem when incorporating lignin is its compatibility with PLA, which can be overcome with the addition of triallyl isocyanurate (TAIC), leading to the formation of PLA-TAIC-lignin crosslinked structures as interface, improving the compatibility in the blend and thus the mechanical, thermal, and hydrolytic degradation properties [123]. However, using TAIC, the biocompatibility has to be studied before applying it for food packaging. Another possibility is the introduction of lignin nanoparticles (LNP) into PLA using a Pickering emulsion template method where lignin acts as stabilizer. According to the results of this study, lignin could increase the decomposition temperature by approx. 10%, reduce the light transmission in the UV region, and increase the Young’s modulus but also decrease the tensile strength and elongation at break. Moreover, the crystallinity of PLA could be improved with the addition of lignin [124]. In another approach, LNP in PLA films could inhibit the growth of bacterial plant pathogens and showed a high antioxidant activity, while migration values remained below the legislative limits, suggesting the exploitation of LNP-PLA films as food packaging material [125].
The combination of LNPs with another lignocellulosic nanofiller, namely, cellulose nanocrystals (CNC), in PLA films can in fact improve UV light blocking capability and strength and modulus values compared to neat PLA or PLA binary systems, confirming a synergic effect of LNP and CNC [126]. This was also reported in another study, where Young’s modulus, elongation at break, and toughness of neat PLA films were improved by 14%, 77%, and 30%, respectively, by incorporation of high lignin-containing cellulose nanocrystals. In contrast, commercial lignin-coated CNCs showed inferior crystallinity, smaller surface area, and a higher degree of agglomeration, concluding that the presence of LNPs is important for the compatibility between the PLA polymer matrix and CNCs [127].

3.2.2. Polymers Produced by Microorganisms

Here, we discuss the most prominent representatives of polymers produced by microorganisms: PHA and PHB and corresponding copolymers or composites. In the group of PHAs, more than 100 known bio-derived polymers exist. The most common ones are PHB and corresponding copolymers such as poly(3-hydroxybutyrate-co-3-hydroxy-valerate) (PHBV) and poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx) (Figure 3) [110,128].
In general, PHAs can be used to coat paper or paperboard to produce water-resistant surfaces, making the coated material completely biodegradable. PHB-coated paperboard has been used for packaging of ready meals, while PHBV-coated paperboard has been used for dry products, dairy products, and beverages [129]. In addition to functionalizing the surface of fiber-based materials, PHAs can also functionalize paper and board’s grease resistance and sealability [110]. PHAs involve a range of biodegradable thermoplastic polymers that are produced through fermentation by different microorganisms [99]. These polymers are characterized by thermomechanical properties that are similar to synthetic polymers such as PP [130]. PHAs can be processed into different products including films, trays, and coatings on other bio-based materials (e.g., paperboard) [110].
Initially, PHAs were used to make everyday articles like shampoo bottles. Moreover, they are used to produce carrier bags, containers and paper coatings, biodegradable bags, and lids. Currently, their use in terms of packaging applications is restricted as they are not transparent [106,110]. Their gas and water vapor permeability offer opportunities to be applied as food packaging materials. Copolymerization as well as blending are used to improve physical-mechanical properties of PHAs [110].
Several studies have shown that PHB, PHBV, and PHBHHx films are promising materials for food packaging due to their good barrier properties. The oxygen permeability of PHAs is comparable to PET and PLA. The values are much lower compared to conventional polymers such as PE and PP. The water vapor permeability of PHAs is similar to materials such as PET and PLA but slightly higher than more apolar polymers such as PE and PP. The carbon dioxide permeability of PHAs is higher compared to PET but substantially lower than for common packaging materials such as PP and PE (Table 2) [131,132,133,134,135,136].
Dilkes-Hoffman et al. summarized that a combination of PHA with thermoplastic starch (TPS), which provides one of the best oxygen barriers of all polymeric materials, seems to have the potential to lower food spoilage rates compared to conventional packaging materials according to good barrier properties [45].
PHB shows a high crystallinity and a high melting point. Therefore, PHB is often blended with PLA. This results in materials of improved mechanical, thermal, and physical properties compared to neat PLA [128]. Arrieta et al. figured out that blending PLA with 25% (w/w) PHB resulted in improved oxygen and barrier properties, whilst the inherent transparency of PLA was reduced [138].
Kovalcik et al. studied the melting and crystallization behavior, thermo-oxidative stability, mechanical and viscoelastic properties, and permeability for oxygen and carbon dioxide of composite materials of microbial PHBHV with methanol fractionated kraft lignin. The results showed that a concentration of already 1 wt% of methanol-extracted kraft lignin can act as an active agent for decreasing the oxygen as well as carbon dioxide permeability of PHBHV films. The gas permeability was decreased for oxygen by 77% and by 91% for carbon dioxide, respectively, compared to the native PHBHV film. The low thermo-oxidative stability of pure PHBHV was increased for the lignin-containing films. Based on this results, methanol-extracted kraft lignin is suggested as a suitable active additive in PHBHV films for applications, especially in the field of food packaging [133].
Although, high production costs limit the competitiveness in commercial applications, PHAs might have high potential as bio-based and biodegradable plastic packaging materials in the transition towards a circular economy [110].

3.3. Selected Plant-derived Polymers

3.3.1. Lignocellulosic Biomass and Lignin

Lignocellulosic biomass is the major structural component of plants, mainly consisting of cellulose (40–60%), hemicellulose (10–40%), and lignin (15–30%), whereby the latter one is the most complex constituent [139]. Lignin is a randomly crosslinked macromolecule composed of the three monolignols p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol, which form the residues p-hydroxyphenyl (H), guaiacyl (G), and syringyl (S), respectively, as shown in Figure 4 [140]. It can be obtained from woody biomass (e.g., pine, poplar, birch), annual plants (e.g., wheat straw, miscanthus, switchgrass), or agricultural residues (e.g., sugarcane bagasse) by various extraction processes. The molecular structure of lignins strongly depends on the botanical origin but also on the growing site, season, and isolation process [141,142]. There are different types of technical lignins that can either be classified as sulfur-containing or sulfur-free. The most common ones are lignosulfonates, kraft lignin, organosolv lignin, and soda lignin [139]. Most industrial lignins originate from the pulp and paper industry with up to 90 million tons of kraft lignin released per year worldwide, though only 2% of it are used commercially for value-added products [143]. One reason for that might be the deficient quality or missing specifications of technical lignins, as they are rather undefined products with a complex composition and impurities from the pulping process (such as remaining sugars or thiol groups). So far, this restricts their industrial exploitation, as resulting products have varying properties, which are inferior to fossil-based products.
Nevertheless, in consideration of the global energy crisis and the depletion of fossil fuels and petrochemicals, the potential of lignin has been a key topic in biorefinery research [139,144,145,146]. As polyphenols, lignins possess numerous interesting functional properties, such as antioxidant activity, and thus, they are investigated as active packaging materials for the protection of light- or oxygen-sensitive goods [119]. Antioxidant polymers are a field of great interest, as the use of macromolecular antioxidants is related to the possibility to produce materials with long-term stability. Due to its cross-linked 3D structure and enzymatic resilience, lignin possesses a higher thermal and biological stability compared to low molecular weight compounds and thus could be used in special fields where the exploitation of low-molecular antioxidant substances would be inefficient due to their higher diffusion rates. Moreover, carcinogenic effects have been observed for synthetic antioxidants: butylated hydroxyanisole (BHA) and butylated hydroxytoluene (BHT), for example, are able to cause cytotoxicity and carcinogenesis as shown in in vivo studies [147]. Azadfar et al. have already shown that lignin has the potential to serve as raw material for antioxidants like guaiacol and 4-vinylguaiacol, whereby their antioxidant activity is comparable to that of commercial antioxidants [148].
Consequently, lignin has gained increasing interest as environmentally benign antioxidant and its ability to improve mechanical, thermal, and barrier properties when incorporated in conventional packaging films. In general, a reduction in WVP of lignin-based films is explained by the hydrophobicity of lignin. It acts as a barrier in the polymeric matrix and increases the path for the diffusion of water vapor, resulting in lower permeation of water molecules through the polymer film [149]. Next to the utilization of lignin as additive/blend in different polymer matrices, it could also be used as raw material for the development of polymeric packaging materials, e.g., as polyol substitute in polyurethanes or polyesters or as phenol substitute in resins. Hult et al. investigated softwood lignin esterified with tall oil fatty acid as coating on paper board. The thermoplastic properties of lignin were enhanced, and the water vapor and oxygen transmission rate decreased while tensile strength was not affected [150]. Polyurethanes of high transperency and flexibility for construction or packaging applications were prepared by Klein et al., where petroleum based polyols could be substituted with kraft lignin up to 80 wt% [151]. In addition, demethylated lignins were also used to enhance the reaction selectivity towards polyurethane formation [152]. In a recently published study by Hao et al. thermoset coatings with integrated self-healing and removal properties were investigated. They investigated a kraft lignin functionalized with carboxylic acid groups as curing agent with poly(ethylene glycol) diglycidyl ether, which resulted in a crosslinked structure [153]. Hambardzumyan et al. designed novel nanocomposite films of lignin and CNC that could be self-supported or used as coatings. This combination of the antibacterial properties of lignin and oxygen barrier properties of CNC films are promising as food packaging material [154].
Rastogi et al. summarized different possibilities for bio-based paper coatings. Next to polyester and polysaccharides discussed so far, also lipids and proteins could be used [155]. Nevertheless, a full exploitation at industrial scale is not possible, due to different crystallization behaviour, brittleness, or melt instabilities that lead to difficulties in processing of these biopolymers. Blending the lipids and proteins with other biopolymers, such as lignin, may provide a route to overcoming this obstacle.
Lignin cannot only be used in paper coatings but also can be used to improve the strength of paperboard. Flory et al. developed a green binder system with Salix lignin that was equal to the wet tear strength of the commercial vinyl acetate binder [156]. Inspired by the reinforcement principle of lignin and cellulose in wood, Jiang et al. developed a cellulose fiber scaffold with lignin as reinforced matrix via successive infiltration and mechanical hot-pressing treatments. The resulting composite shows a high isotropic tensile strength of 200 MPa, compared to 40 MPa of conventional cellulose paper, and Young’s modulus of 10 GPa, which is even higher than many fossil-based plastics. In addition, also the thermostability and UV-blocking performance is enhanced due to the lignin addition [157].
In conclusion, lignins are promising candidates for environmentally benign antioxidants with a great abundance [158,159,160]. Regarding the studies summarized in this review, lignin use as an additive seems more favorable than its copolymerization, as copolymerization usually requires at least one functionalization (such as demethylation or carboxylation). Thus, further effort is necessary to get a deeper understanding of the lignin structure and the processing conditions required to maintain and enhance the antioxidant and antimicrobial properties.

3.3.2. Protein-Based Polymers

Research in food packaging has also focused on protein-based films due to their good film-forming properties, low cost, and biodegradable nature [83]. Materials synthesized from proteins exhibit desirable film-forming and barrier properties, which are often comparable to fossil-based products [90]. Proteins from different sources have been investigated for the synthesis of bioplastic films for packaging applications including collagen, gelatin, caseins, soy/whey/quinoa protein, egg white protein, myofibrillar protein, corn zein, wheat gluten, and keratin [83,161,162].
Among all the protein sources, soy proteins got great attention as a potential source for bio-based packaging materials. This development is based on excellent film-forming and oxygen barrier properties of films produced from soy protein isolate. However, these materials cannot meet the requirements of a film with mechanical and water barrier properties guaranteed by conventional plastics [83,163]. Compared to films from other proteins, soy protein-based films are characterized by transparency, flexibility and cost-effectiveness [162]. Furthermore, they show good oxygen barrier properties under low moisture conditions [164]. A disadvantage that limits their use beside low mechanical strength is a lack of heat stability compared to LDPE [165,166,167,168].
Whey proteins are able to form elastic films [169], which are transparent, flexible, and exhibit good oil and oxygen barrier properties at low humidity. A disadvantage is a moderate moisture permeability. Nevertheless, whey proteins have been intensively studied as raw material for biodegradable packaging [170].
Collagen and gelatin are proteins originating from animal sources acquired by a controlled hydrolysis reaction. In nature, collagen is the most abundant occurring protein [171]. Collagen-based bioplastic films are characterized by good mechanical properties, [172] and therefore, are suitable for various applications [173]. In contrast to collagen-based films, gelatin films show poor mechanical and barrier properties according to their hydrophilic nature [174]. Biscarat et al. determined functional properties of gelatin-based films. Compared to synthetic polymers, good gas barrier properties were reached by gelatin films cross-linked with ferulic acid. Gelatin films with poly(ethylene glycol) (PEG) 200 showed high gas barrier properties and high permselectivity towards carbon dioxide and oxygen [80].
Gelatin has been introduced in the manufacturing of packaging films due to its low cost and abundance [175]. Furthermore, gelatin is used to produce biodegradable packaging materials due to its good properties, such as low melting and gelling points, good capacity of oxygen barrier, biodegradability, and excellent film formation [176]. The use of gelatin-based composite films incorporating other materials like chitosan, sunflower oil, and corn oil to enhance the barrier and mechanical properties of these films was studied by different authors [177]. In addition to gelatin, gluten is used to prepare films of high homogeneity, excellent gas barrier properties, and mechanical strength [178].
Kanatt developed a new intelligent-active food packaging film using poly(vinyl alcohol) (PVA) and gelatin incorporated with Amaranthus leaf extract to monitor freshness and increase the shelf life of fish and chicken meat. Incorporation of Amaranthus leaf extract improved its mechanical and water vapor barrier properties next to active functions. The decrease in solubility enables the use for packaging of flesh foods. Samples packed in neat films had a shelf life of 3 days while those in active films spoiled after 12 days. The results of the study suggest the application of Amaranthus leaf extract containing PVA-gelatin films being both active and intelligent ensuring quality and safety of flesh foods [179].
For comparative purposes, Table 3 shows the discussed barrier as well as mechanical properties of protein-based polymers.

3.3.3. Polysaccharides

Packaging films based on carbohydrate sources are generally transparent and homogeneous films with effective oxygen barriers at intermediate to low humidity and good mechanical properties. According to their hydrophilic character, they have poor water vapor barrier qualities, and they are relatively sensitive to moisture. Furthermore, the films obtained from several polysaccharides are brittle usually due to interactions between the polymer chains. This leads to limited applications of polysaccharide- and protein-based coatings and films [2,106,184]. To meet the required properties, a pretreatment such as plasticization with small molecular weight–compatible constituents, blending, or chemical modification is needed [130]. While considering mechanical properties, the tensile strength of polysaccharide-based films are similar to those of synthetic polymers; differences are observed in elongation at break (Table 4) [82].
For several polysaccharides, the film-forming properties and especially their potential for edible packaging has been studied, including starch, cellulose, and its derivatives, alginate and chitosan (Figure 5) [161,185,186,187].

Starch

Starch is a natural polysaccharide easily available on an industrial scale. Many plant-based polysaccharides and cost-effective starch-based materials have been extensively investigated as an alternative material for fossil-based food packaging applications due to their environmental compatibility and biodegradability [198]. Although starch-based packaging, which has good film-forming properties and excellent oxygen barrier, is already in wide use, this material still has some disadvantages such as the poor vapor and oxygen moisture barrier and poor mechanical properties compared to conventional non-biodegradable plastics used in food packaging industry (Table 1 and Table 4) [198,199,200,201].
The oxygen barrier properties are correlated to a high-ordered hydrogen-bonded network structure. The barrier properties can be improved by increasing the crystallinity or a higher content of amylopectin [199]. The poor moisture barrier is caused by a strong hydrophilic behavior [200]. A higher crystalline structure in starch-based films leads to less sensitivity of moisture and to environmental relative humidity. Considering the poor mechanical properties, starch-based films show a relatively high tensile strength while the elongation percentage is low [199]. The high tensile strength is attributed to the extensive intra-molecular hydrogen bonds between amylose, amylopectin, and amylose-amylopectin molecules. Amorphous regions in starch-based films formed by amylose cause brittleness and thereby influence the poor mechanical properties [189].
As starch-based films are odorless, colorless, and tasteless, starch is used in either pure or blended form as a biodegradable coating or packaging film. The coatings of edible starch are also applied for other kinds of foods to maintain quality and to extend the shelf life of products [200,202,203,204]. However, as starch blends may contain additives like compatibilizers and plasticizers that can migrate out of the matrix, only some starch blends are suitable for food packaging applications [106].
Starch has already been combined with lignin to improve its poor thermo-mechanical properties while simultaneously decreasing its water vapor permeability significantly [205]. Miranda et al. confirmed these findings in their studies, showing that the presence of lignin in combination with CNC increased maximum stress and modulus of elasticity, barrier properties, and the thermal stability of the material [206,207]. Javed et al. published a study concerning starch-based coatings for paper packaging materials with lignin, investigating the self-supporting films regarding their mechanical properties and chemical stability in water as well as their barrier properties when used as coating on paper board. When lignin is added, the dissolution of starch from the composites could be significantly decreased. The addition of ammonium zirconium carbonate (AZC) leads to further improvement of the storage modulus, indicating that crosslinking had occurred [208].

Cellulose and Derivatives

Cellulose represents the most abundant renewable polymer source available in nature. Biodegradable films made out of this raw material are characterized by renewability, low cost, non-toxicity, biocompatibility, biodegradability, and chemical stability [209]. For example, cellulose films (known as Cellophane®) are used for wrapping fruits in bio-based trays [105].
Films made of cellulose exhibit good toughness, tensile strength, high surface gloss, and good transparency [210]. Whereas hemicellulose-based films are brittle, the flexibility, toughness, and oxygen permeability can be improved by addition of plasticizers [211,212]. A disadvantage of cellulose films is their poor water vapor barrier. This is caused by the underlying hydrophilic nature of polysaccharides [83]. Soaking cellulose with alkali to swell the structure followed by different derivatization reactions, CMC, methylcellulose (MC), hydroxypropyl cellulose (HPC), and hydroxypropylmethyl cellulose (HPMC) are available. They are used as raw materials to prepare biodegradable films, which are transparent, water soluble, odorless, tasteless, and flexible and have moderate strength and resistance to lipid compounds (Table 4) [213].
CMC is most often used for biodegradable film production. It is highly soluble (in water) and crystalline and can build solid and flexible films [214,215]. Beneficial characteristics include film-forming properties, good mechanical and gas barrier properties, transparency, ease of processing, and low price [216]. Next to its good film-forming properties, CMC has been studied as antibacterial food packaging in composites with chitosan [217,218] and pectin [219]. Michelin et al. investigated the incorporation of organosolv lignin from corncob in CMC-based films, which leads to an improved water resistance of approx. 60% and reduction in the water vapor permeability of 20%, while also enhancing the thermal stability and antioxidant activity [149].
As mentioned before, mechanical and barrier properties of cellulose-based films can be improved by the production of nanocomposites. Moura et al. proposed nanocomposites using chitosan as nanofiller in HPMC to enhance mechanical and film barrier attributes. HPMC films containing different concentrations of chitosan as nanoparticles were analyzed for mechanical properties, water vapor permeability, and oxygen permeability. They realized that chitosan nanoparticles tended to fill up poriferous spaces in the HPMC-matrix. This improves film tensile properties and water vapor permeability, concluding that a HPMC-chitson nanocomposite could be a possible material for food-packaging applications to extend the shelf life of food [158,220].
Another approach is the preparation of HPMC/chitosan films [221]. Alzagameem et al. examined the incorporation of different lignins in HPMC and HPMC/chitosan films [158,160]. Results show that lignins are in general more active against Gram-positive bacteria than against Gram-negative bacteria and that films with organosolv lignin possess a higher activity against S. aureus than films with kraft lignin. It was shown that biomass as well extraction process influence the properties of the films and that the antioxidant activity of lignins correlates with different parameters such as genotype and phenotype of biomass, pulping and purification of lignin and the resulting heterogeneity [159].
Cellulose acetate (CA), cellulose acetate propionate (CAP), and cellulose acetate butyrate (CAB) are thermoplastic cellulosic-derived materials developed through esterification of cellulose. CA gets great attention because of its biodegradable nature, excellent optical clarity, and greater toughness [83].

Alginates

Alginate and corresponding derivatives are one of the most promising carbohydrates for packaging applications, especially for foods that are sensitive to gas permeation [222]. Alginates are naturally occurring indigestible polysaccharides. They are commonly produced from various genera of brown algae [223]. A lot of research on alginate has focused on edible coatings or to improve color and flavors [224,225].
Ahmed et al. evaluated the barrier, mechanical, and oil resistance of paper sheets coated with a novel cost-saving coating prepared at different temperatures to be used in packaging purposes. The coating is based on sodium alginate (SA) with new core-shell inorganic particles composed of waste silica fume core covered with cobalt(II) oxide/zinc oxide (CoO.ZnO) oxides. SA is broadly used due to high oil resistance and enhanced greaseproof properties.
Very recently, a binary oxide CoO.ZnO on the surface of silica was shown to decrease the penetration of oil and grease through the paper pores leading to lower oil absorption and enhanced mechanical properties. Tensile strength was decreased whereas stiffness showed slight increase in case of paper sheets coated with SA-CoO.ZnO/SiO2. Incorporation of SA-SiO2 and SA-CoO.ZnO/SiO2 in the fiber matrices improved the tear and burst indices properties. The network created by SA film blended with (CoO.ZnO/SiO2) pigments on paper sheet substrate drastically changed their mechanical and barrier properties. They are highly dense and organized, causing the creation of smooth distribution of nanoparticles and a strong surface that fills the pores in the paper matrices and increases the air, water vapor, and oil resistance of the samples coated with SA-CoO.ZnO/SiO2 pigments compared with uncoated paper [223].

Chitosan

Chitosan is the second most widely found amino-polysaccharide in nature [226]. It is a biopolymer with antibacterial properties that can be found in fungi, insects, crabs, and shrimps and makes up a significant part of the crustacean waste that is discarded every year [227]. Chitosan is intensively investigated for biomedical applications [228] and bioactive packaging solutions [229]. Properties like being microbe resistant, biocompatible, and biodegradable make chitosan attractive for research and use in various applications [230]. In food packaging, potential applications of chitosan blend-based films are for fresh products (vegetable, meat, and fish) and foods with short to medium shelf life [231]. Moreover, chitosan is widely used as a material for nanofibers production due to its ability to form films with antibacterial properties [232,233]. Moreover, chitosan-based materials exhibit good mechanical properties and a selective permeability to carbon dioxide and oxygen [101,234]. The tensile strength and elongation at break values are comparable to appropriate values of HDPE and LDPE (Table 4).
However, the use of chitosan to produce flexible packaging is still limited due to high sensitivity to humidity and moisture and high oxygen permeability. As chitosan films are highly permeable to water vapor, their use in food packaging is restricted [234,235,236]. To overcome these limitations of this otherwise excellent polysaccharide biomaterial, several attempts have been made such as blending the films with other natural or synthetic polymers and the addition of several active and functional substances like fillers, plasticizers, crosslinkers, and natural oils [234].
Several studies are published that focus on improving the barrier and mechanical properties of chitosan-based films. Kamdem et al. developed a composite flexible film using chitosan as base polymer matrix, xylan to improve mechanical properties, and carvacrol (a monoterpenoid phenol) to control microbial protection. The results show that adding xylan significantly increases the elongation at break of the composite films and exhibits higher tensile strength and Young’s modulus. The incorporation of carvacrol and xylan in the composite films was not effective in terms of antimicrobial activity [101].
Chitosan is another potential packaging material that is investigated with lignin as additive. The addition of lignin to chitosan improves the tensile strength, storage modulus, glass transition temperature, and degradation temperature, compared to pure chitosan films [237]. Moreover, lignin confers the scavenging properties to the chitosan films: this antioxidant activity is highly dependent on the film structure, functional properties, and surface activity, which is, in turn, dependent on moisture [238] and the homogenization process of the film-forming suspension [239].
Despite favorable properties of chitosan like its antioxidant and antimicrobial activity as well as a good biocompatibility and biodegradability, there are some drawbacks such as its dissolution in acidic media and poor thermal properties. Thus, the combination of chitosan with other polymers in binary or ternary films is intensively studied for various systems. Yang et al. investigated films based on PVA, chitosan, and LNPs for food packaging applications and found that lignin improved tensile strength and Young’s modulus as well as the thermal stability of the systems [240].
In conclusion, Table 5 summarizes advantages and disadvantages of the polymers discussed in Section 3.

4. Natural Additives in the Context of Active Food Packaging

In the packaging and food industry, active packaging that prolongs shelf life and reduces food losses has already been widely used [67,241]. Currently, these packaging types are still mainly based on polymers from fossil-based resources [35]. However, as described previously, several other strategies in the field of bio-based materials are under development.
Currently, active packaging solutions follow different approaches [1,242]:
  • Addition of absorbers and scavengers of gases, off flavors, moisture, taints, UV light;
  • Removal of catalyzing undesired food components;
  • Addition of emitters/generators of gases and flavors;
  • Release of antioxidant and/or antimicrobial compounds;
  • Temperature controlled systems (insulting materials; self-heating or cooling).
Thus, basic barrier properties (such as oxygen or moisture) can be improved by adding active ingredients in the packaging system and/or using functionalized polymers. Polymers used in films and coatings that are inherently antimicrobial are chitosan, poly(L-lysine), calcium alginate, acrylic polymers, and sustainable active microbiocidal (SAM) polymers [11,243].
A promising technology that is presented and discussed in this chapter is the integration of antimicrobial and antioxidative substances [244,245]. Antimicrobial systems target the control or reduction of microbial growth which often results in the extension of the lag phase or in a reduced growth rate in the exponential phase [246,247]. Different antimicrobial strategies using plant extracts are based on absorption, release, and immobilization systems [244].
There are different opportunities to implement antimicrobials in packaging materials. In terms of time-releasing killing, either a volatile or non-volatile antimicrobial agent is temporarily trapped within the backbone material and released from the polymer to the environment. Volatile antimicrobial agents are released through evaporation or diffusion into the headspace of food in most cases without direct contact. Non-volatile antimicrobial agents are released by direct contact through diffusion into the food surface [243,248]. Another approach is the permanent immobilization of a non-volatile antimicrobial agent to a polymer backbone. The integration of antimicrobial agents in packaging can be realized by direct incorporation of the antimicrobial into the packaging material or by coating the packaging material with antimicrobial agents [243,244,248].
Natural antimicrobial, antioxidative, and photostabilizing agents used for the preservation of food are bacteriocins and extracts from biomass of animals, plants, and microorganisms. These include enzymes, EOs, and natural extracts from different plant sources [249,250]. Phenolic compounds and terpenoids are antioxidative and antimicrobial agents that occur in EOs extracted from different plants. Eos are the most abundant source of bioactive compounds [251]. The effect of EOs on the microbial cells depends on different mechanisms such as disrupting the enzyme structures, damaging the phospholipid bilayer of cell membrane, and compromising the genetic makeup of microbes [98]. Antioxidants are compounds that react with free radicals, neutralizing them and thereby preventing or reducing their damaging effects. Aromatic plants are a source of natural antioxidants because of the activity of secondary metabolites such as phenylpropanoids and EOs [252]. The antioxidant capacity of plant extracts is strongly related to the phenolic content [253,254]. Due to their redox property, they can act as reducing agents, hydrogen donors, singlet oxygen quenchers, metal chelating agents, and suppressors of free radicals [159,255]. The antioxidant activity is not a property of a single phenolic compound, but it is widely distributed among the phenolic phytochemical constituents. So, anthocyans, flavonoids, phenolic acids, phenolic terpenes, and volatile oils are particularly interesting as antioxidants in food packaging [159,256].
The search for appropriate substitutional bio-based core materials is important for tackling environmental issues. Currently, those materials include, but are not limited to, chitosan, starch, CMC, PLA, whey proteins, and combinations thereof. As they represent a comparably small amount of the packaging itself when compared to core materials, additives tend to be neglected. However, they are often critical to achieve the desired properties in packaging materials and can be capable of prolonging the shelf life of both packaging materials and packed goods, leading to reduced food loss.
In the following subchapters, the review summarizes applications of plant EOs and plant extracts in the context of active food packaging.

4.1. Plant Essential Oils

Plant-based active stabilizers can be isolated by extracting the appropriate biomasses by steam distillation using various solvents. Generally, EOs are complex mixtures containing over 300 different polar and non-polar volatile organic compounds, usually of low molecular weight (below 300) at quite different concentrations [257]. Often, two or three major components exist at relatively high concentrations (20–70%) while others are present in trace amounts. The major constituents of EOs are terpenoids and phenylpropanoids which provide the characteristic aroma and biological properties. Both families are comprised of phenolic compounds. The sufficiently high vapor pressure of EOs, in general, at atmospheric pressure and room temperature causes them to be found partially in the vapor state [257,258,259].
In comparison to producing plant extracts, a higher amount of plant biomass must be processed to obtain EOs. However, this can prove worthwhile as EOs contain active components of biomasses in particularly high concentrations and are therefore typically highly effective in different applications. This allows manufacturers to obtain great effects with applying only small amounts of oil. Prominent examples of EO applications for active food packaging are listed in Table 6. Results obtained so far show that plant EOs are interesting components for active food packaging. However, the corresponding materials have to be specified regarding water resistance; water vapor permeability; mechanical properties; and enhancement of antimicrobial, antioxidant, photostabilizing, and light-absorbing properties. Particular research interest lies on EOs obtained from food sources such as cinnamon, thyme, and rosemary or food production byproducts as apricot kernels and banana leaves.

4.2. Plant Extracts of Various Biomasses

Quality and quantity of plant extracts strongly depend on their biomass origin and type of extraction processes. The most common extracts have been obtained by conventional solvent extraction methods (infusion, decoction, digestion, maceration, and percolation) using solvents such as water, ethanol, methanol, chloroform, or dimethyl-sulfoxide [289,290,291].
Such extracts typically retard bacterial growth and can also introduce antioxidant and photoabsorbing effects. Those properties are reported for various different biomasses including (but not limited to) herbs, flowers, trees, and their fruits [292,293,294,295,296,297,298]. In contrast to essential oils, the effects tend to be less extensive; however, less biomass must be processed to obtain such extracts. Furthermore, as high-concentrated essential oils provide lipophilic surroundings for active substances, the application of polar active substances is only possible by (hydrophilic) extraction, thus introducing a whole new group of active compounds. The specific characteristics observed in plant extracts can be utilized by incorporating them in food packaging materials to positively affect the packed food. Prominent examples for this approach are documented in Table 7. The studies using plant extracts show similar improvements in packaging characteristics to plant essential oils. Decreased water vapor permeability; enhanced moisture and oil resistance; improved mechanical properties; and enhanced antimicrobial, antioxidant, photoabsorbing, and UV-stability are reported. While it is challenging to directly compare the obtained data with each other due to a variety of tests and extraction methods used, most researchers claim a relevant potential exists for plant-based stabilizers in food packaging applications. Again, plant-based active packaging research is typically focused on biomasses that represents foodstuff.

4.3. Encapsulated Plant Essential Oils

Advanced methods use plant essential oils after encapsulation by a variety of different techniques, including formation of nanofibers, nanotubes, and nanoparticles [308]. This way, the essential oils are more resistant against thermal influences [309,310,311]. The incorporation of encapsulated essential oil typically also improves the mechanical properties of packaging materials [312,313]. Encapsulation furthermore facilitates gradual release of active ingredients, leading to a more durable protection of the packed foodstuff. Encapsulated essential oils are also under investigation in other fields and applications including bio-based insecticides and cleaning agents [314,315]. Recent studies utilizing the encapsulation of essential oils in the context of food packaging are presented in Table 8. The results confirm that encapsulated plant oils are able to improve water vapor permeability, transparency, and tensile strength as well as antioxidant and antimicrobial effects.

5. Adoption Potential of Bio-Based (Active) Packaging along the Value Chain

Active packaging based on biopolymers is identified as the more sustainable alternative compared to conventional packaging. In addition, the integration of natural additives has positive effects on the quality and shelf life of the packaged product [39,329,330]. However, in order to be successful in the market, this innovative concept needs to be adopted along the whole agricultural food value chain [331,332].
More specifically, farmers need to collect, process, and deliver raw materials such as annual plants (e.g., miscanthus) or residues from agricultural production (e.g., sugarcane bagasse) [333]. The packaging industry must adopt renewable resources as raw materials and might also need to adjust their production processes for the application of bio-based and/or biodegradable polymers as a packaging core matrix and to integrate natural additives into the packaging materials. Food companies need to be willing to pay more for the material to pack their products, and the consumer needs to accept the concept of bio-based active packaging [334,335]. As the implementation of active packaging based on bio-based polymers entails several changes for farmers, industry, and consumers, the remaining section reviews extant studies exploring the adoption decisions of these value chain actors.
Existing literature looking at the adoption behavior of farmers finds that these value chain actors are generally skeptical towards innovations related to the bioeconomy. Therefore, monetary incentives and assistance with the novel practices and processes might be necessary [336]. In addition to farmers, food processing companies might also serve as the provider of by-products as raw materials to produce active bio-based packaging. However, there is currently a lack of research regarding the adoption decisions of managers in those companies. Today, a growing number of farmers are interested in adopting practices to valorize by-products [337]. Therefore, these farmers need to be targeted by policy initiatives and could then serve as opinion leaders to positively influence the adoption decisions’ of their communities [338,339].
Focusing on industry representatives such as packaging producers and food companies, exploratory studies identify several factors driving their adoption decision. First of all, the market prices and the availability of renewable resources for the production of bio-based polymers and natural additives are relevant for the adoption decisions of packaging producers [334]. Moreover, relevant policy instruments need to be implemented to foster research and development of bio-based polymers with natural additives (e.g., subsidies) or even to ban conventional (multi-layer) plastics [14,334,340]. This would increase the competitiveness of bio-based and/or biodegradable packaging with conventional plastics [334]. However, even when policy instruments are in place, the level of consumer demand is the most important factor driving the adoption decision of industry representatives [340].
Consumer studies indicate that the final actors in the value chain have both positive and negative associations with bio-based products. They may misunderstand the concept of ‘bio-based’ [341,342,343]. However, the majority of consumers seem to believe that sustainable packaging is important and useful [344]. Moreover, results of two studies provide evidence that bio-based packaging seems to increase the preferences for the packaged product [345,346]. In fact, empirical results from a discrete choice experiment indicate that consumers are willing to pay a price premium for bio-based plastic packaging [335]. Considering the calculations by van den Oever et al., this price premium even covers the additional costs for bio-based and/or biodegradable plastics compared to conventional materials [39]. Food companies could therefore switch to bioplastic packaging without expecting any lost profits [335]. Moreover, especially those consumers with high levels of environmental awareness and innovativeness seem to prefer bio-based plastics over conventional plastic products [347,348,349]. Products packaged with bio-based materials thus need to be presented in retail locations which are preferred by this type of consumers such as organic stores [335]. However, as bio- and fossil-based plastic packaging are not easy to be distinguished by consumers, the packaging needs to be labelled accordingly [350].
After its use as packaging material, end-of-life solutions also need to be considered for active packaging derived from bio-based and/or biodegradable polymers from renewable resources [335]. Depending on consumers’ disposal behaviors, bio-based bioplastics can be decomposed given the right conditions (in case of biodegradable and/or compostable compounds) or the material can be used to generate renewable energy. Thus, it is very important that the disposal options are clearly communicated to the consumers [351]. In fact, consumers are even willing to pay a price premium for biodegradable and recyclable packaging [352,353].

6. Conclusions

Besides current political requirements that aim to improve sustainability aspects, the development and promotion of more sustainable materials have gained more importance due to consumer interests. As customers preferences shifted to high quality and safe products with enhanced shelf life, the development of various new trends in packaging systems has arisen. Research focuses on improving the characteristics of bio-based packaging materials, in particular mechanical, thermal, and physical properties. Although bio-based polymers provide significant opportunities in terms of sustainability and biocompatibility, their use in industrial applications is often restricted due to lower performance in fundamental packaging functions. Companies are faced with a challenge of alternatives offering higher costs, limited functionality, existing infrastructure, and inconsistent legislation. Furthermore, a lack of compatibility with conventional processing technologies has to be overcome. Food companies need to be willing to pay more for the material to pack their products and the consumer needs to accept the novel concept of (active) packaging. The proof-of-concept is shown by a few commercially available biopolymers with food applications such as PLA, PHAs, PEF, PBS, and thermoplastic cellulose or starch-based films. In the future, the market for sustained active packaging will certainly increase due to enhanced efforts and innovations in material development and processing technologies.

Author Contributions

Conceptualization and writing—original draft preparation, I.K.; writing—original draft preparation, J.N.F., T.H., J.R., J.W., and M.S. (Margit Schulze); writing—review and editing, project administration and funding acquisition, S.B., J.K., R.P., M.S. (Michaela Schmitz), and M.S. (Margit Schulze). C.K. (native speaker) performed final English revision. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by EFRE/NRW “Biobasierte Produkte” (Grant EFRE 0500035). J.R. gratefully acknowledges a scholarship given by the Graduate Institute of the Bonn-Rhein-Sieg University of Applied Sciences.

Institutional Review Board Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

List of Abbreviations

APETAmorphous poly(ethylene terephthalate)
AZCAmmonium zirconium carbonate
CACellulose acetate
CABCellulose acetate butyrate
CAPCellulose acetate propionate
CMCCarboxymethyl cellulose
CoO.ZnO Cobalt(II) oxide/zinc oxide
EOEssential oil
GHGGreenhouse gas
HDPEHigh density poly(ethylene)
HLCNCHigh lignin-containing cellulose nanocrystals
HPMCHydroxypropylmethyl cellulose
LCALife cycle analysis
LDPELow density poly(ethelene)
PAPoly(amide)
PBATPoly(butylene adipate terephthalate)
PBSPoly(butylene succinate)
PCLPoly(caprolactone)
PEPoly(ethylene)
PHAPoly(hydroxyalkanoate)
PHB Poly(3-hydroxybutyrate)
PEFPoly(ethylene furanoate)
PET Poly(ethylene terephthalate)
PHBHHxPoly(3-hydroxybutyrate-co-3-hydroxyhexanoate)
PHBVPoly(3-hydroxybutyrate-co-3-hydroxyvalerate)
PLAPoly(lactic acid)
PPPoly(propylene)
PS Poly(styrene)
PTTPoly(trimethylene terephthalate)
PVAPoly(vinyl alcohol)
PVCPoly(vinyl chloride) (PVC)
SASodium alginate
SAMSustainable active microbiocidal (SAM)
SiO2Silicon dioxide
TAICTriallyl isocyanurate
WVPWater vapor permeability

Glossary

Active packagingMaterials designed to deliberately incorporated components that would release or absorb substances into or from the packaged food or the environment surrounding the food.
BioactiveCompound that has an effect on a living organism, tissue/cell.
Bio-basedCompound that is composed (in whole or in significant part) of biological products or renewable domestic agricultural or forestry materials (including plant, animal, and marine materials).
BiodegradableDegradability achieved via microorganisms.
BioplasticsPlastics that are either bio-base, biodegradable, or features both properties.
BiopolymersNatural polymers produced by the cells of living organisms (e.g., forestry and agricultural crops, terrestrial and marine animals), examples are polysaccharides, proteins, and lignin.
CompostableCompounds approved to be degradable by microorganisms at defined conditions (e.g., temperature, humidity, time).
Edible packing Compounds approved to be metabolized by humans.
End-of-life optionsIncluding re-use, recycling, recovery, disposal, and others (such as littering, ingestion).
European Green DealAction plan to boost the efficient use of resources by moving to a clean, circular economy, restore biodiversity, and cut pollution.
Fossil-basedCompounds obtained from crude oil, natural gas, brown or hard coal.
Renewable resourceResource which will replenish to replace the portion depleted by usage and consumption, either through natural reproduction or other recurring processes in a finite amount of time in a human time scale.

References

  1. Sharma, R.; Ghoshal, G. Emerging trends in food packaging. Nutr. Food Sci. 2018, 48, 764–779. [Google Scholar] [CrossRef]
  2. Youssef, A.M.; El-Sayed, S.M. Bionanocomposites materials for food packaging applications: Concepts and future outlook. Carbohydr. Polym. 2018, 193, 19–27. [Google Scholar] [CrossRef]
  3. Aggarwal, A.; Langowski, H.-C. Packaging functions and their role in technical development of food packaging systems: Functional equivalence in yoghurt packaging. Procedia CIRP 2020, 90, 405–410. [Google Scholar] [CrossRef]
  4. Kreyenschmidt, J.; Albrecht, A.; Braun, C.; Herbert, U.; Mack, M.; Rossaint, S.; Ritter, G.; Teitscheid, P.; Ilg, Y. Food waste in der Fleisch verarbeitenden Kette. Fleischwirtschaft 2013, 10, 57–63. [Google Scholar]
  5. Accorsi, R. A support-design procedure for sustainable food product-packaging systems. In Sustainable Food Supply Chains; Elsevier: Amsterdam, The Netherlands, 2019; pp. 61–81. ISBN 9780128134115. [Google Scholar]
  6. Coelho, P.M.; Corona, B.; Klooster, R.T.; Worrell, E. Sustainability of reusable packaging–Current situation and trends. Resour. Conserv. Recycl. X 2020, 6. [Google Scholar] [CrossRef]
  7. Barlow, C.Y.; Morgan, D.C. Polymer film packaging for food: An environmental assessment. Resour. Conserv. Recycl. 2013, 78, 74–80. [Google Scholar] [CrossRef]
  8. Matthews, C.; Moran, F.; Jaiswal, A.K. A review on European union’s strategy for plastics in a circular economy and its impact on food safety. J. Clean. Prod. 2021, 283, 125263. [Google Scholar] [CrossRef]
  9. Ellen MacArthur Foundation. The New Plastics Economy: Rethinking the Future of Plastics; Ellen MacArthur Foundation: Cowes, UK, 2016. [Google Scholar]
  10. Faraca, G.; Astrup, T. Plastic waste from recycling centres: Characterisation and evaluation of plastic recyclability. Waste Manag. 2019, 95, 388–398. [Google Scholar] [CrossRef]
  11. Ilg, Y.; Kreyenschmidt, J. Review: Nutzen und Risiken der Anwendung antimikrobieller Werkstoffe in der Lebensmittelkette. J. Food Saf. Food Qual. 2012, 63, 28–34. [Google Scholar]
  12. European Union. Commission Regulation (EC) No 450/2009 of 29 May 2009 on active and intelligent materials and articles intended to come into contact with food. Off. J. Eur. Union 2009, L135, 3. [Google Scholar]
  13. PlasticsEurope. Plastics—The Facts 2019: An Analysis of European Plastics Production, Demand and Waste Data; PlasticsEurope: Brussels, Belgium, 2019. [Google Scholar]
  14. European Commission. A European Strategy for Plastics in a Circular Economy; European Commission: Brussels, Belgium, 2018. [Google Scholar]
  15. Umweltbundesamt. Aufkommen und Verwertung von Verpackungsabfällen in Deutschland im Jahr 2018: Abschlussbericht; Umweltbundesamt: Vienna, Austria, 2020; Volume 166. [Google Scholar]
  16. Ragaert, K.; Delva, L.; van Geem, K. Mechanical and chemical recycling of solid plastic waste. Waste Manag. 2017, 69, 24–58. [Google Scholar] [CrossRef] [PubMed]
  17. Van Eygen, E.; Feketitsch, J.; Laner, D.; Rechberger, H.; Fellner, J. Comprehensive analysis and quantification of national plastic flows: The case of Austria. Resour. Conserv. Recycl. 2017, 117, 183–194. [Google Scholar] [CrossRef]
  18. Sohail, M.; Sun, D.-W.; Zhu, Z. Recent Developments in Intelligent Packaging for Enhancing Food Quality and Safety. Crit. Rev. Food Sci. Nutr. 2018, 58, 2650–2662. [Google Scholar] [CrossRef]
  19. Taleb, M.A.; Al Farooque, O. Towards a circular economy for sustainable development: An application of full cost accounting to municipal waste recyclables. J. Clean. Prod. 2021, 280. [Google Scholar] [CrossRef]
  20. European Commission. Closing the Loop—An EU Action Plan for the Circular Economy; European Commission: Brussels, Belgium, 2015. [Google Scholar]
  21. European Commission. On the Implementation of the Circular Economy Action Plan; European Commission: Brussels, Belgium, 2019. [Google Scholar]
  22. Pro Carton. Wie Wichtig Ist Nachhaltige Verpackung?: Die Einstellung von Konsumenten zu Verpackung und Nachhaltigkeit; Pro Carton: Vienna, Austria, 2010; Volume 3. [Google Scholar]
  23. Yildirim, S.; Röcker, B.; Pettersen, M.K.; Nilsen-Nygaard, J.; Ayhan, Z.; Rutkaite, R.; Radusin, T.; Suminska, P.; Marcos, B.; Coma, V. Active packaging applications for food. Compr. Rev. Food Sci. Food Saf. 2018, 17, 165–199. [Google Scholar] [CrossRef] [Green Version]
  24. Pauer, E.; Tacker, M.; Gabriel, V.; Krauter, V. Sustainability of flexible multilayer packaging: Environmental impacts and recyclability of packaging for bacon in block. Clean. Environ. Syst. 2020, 1. [Google Scholar] [CrossRef]
  25. Fabra, M.J.; López-Rubio, A.; Lagaron, J.M. Biopolymers for food packaging applications. In Smart Polymers and Their Applications; Elsevier: Amsterdam, The Netherlands, 2014; pp. 476–509. ISBN 9780857096951. [Google Scholar]
  26. Koller, M.; Sandholzer, D.; Salerno, A.; Braunegg, G.; Narodoslawsky, M. Biopolymer from industrial residues: Life cycle assessment of poly(hydroxyalkanoates) from whey. Resour. Conserv. Recycl. 2013, 73, 64–71. [Google Scholar] [CrossRef]
  27. Narodoslawsky, M. LCA of PHA Production—Identifying the Ecological Potential of Bio-plastic. Chem. Biochem. Eng. Q. 2015, 29, 299–305. [Google Scholar] [CrossRef]
  28. Ahorsu, R.; Medina, F.; Constantí, M. Significance and challenges of biomass as a suitable feedstock for bioenergy and biochemical production: A review. Energies 2018, 11. [Google Scholar] [CrossRef] [Green Version]
  29. Xu, C.; Arancon, R.A.D.; Labidi, J.; Luque, R. Lignin depolymerisation strategies: Towards valuable chemicals and fuels. Chem. Soc. Rev. 2014, 43, 7485–7500. [Google Scholar] [CrossRef] [PubMed]
  30. Daioglou, V.; Wicke, B.; Faaij, A.P.C.; van Vuuren, D.P. Competing uses of biomass for energy and chemicals: Implications for long-term global CO2 mitigation potential. GCB Bioenergy 2015, 7, 1321–1334. [Google Scholar] [CrossRef] [Green Version]
  31. Weiss, M.; Haufe, J.; Carus, M.; Brandão, M.; Bringezu, S.; Hermann, B.; Patel, M.K. A Review of the environmental impacts of biobased materials. J. Ind. Ecol. 2012, 16, 169–181. [Google Scholar] [CrossRef]
  32. Zhu, Y.; Romain, C.; Williams, C.K. Sustainable polymers from renewable resources. Nature 2016, 540, 354–362. [Google Scholar] [CrossRef]
  33. Broeren, M.L.; Kuling, L.; Worrell, E.; Shen, L. Environmental impact assessment of six starch plastics focusing on wastewater-derived starch and additives. Resour. Conserv. Recycl. 2017, 127, 246–255. [Google Scholar] [CrossRef]
  34. Rhim, J.-W.; Park, H.-M.; Ha, C.-S. Bio-nanocomposites for food packaging applications. Prog. Polym. Sci. 2013, 38, 1629–1652. [Google Scholar] [CrossRef]
  35. De Oliveira Filho, J.G.; Rodrigues, J.M.; Valadares, A.C.F.; de Almeida, A.B.; de Lima, T.M.; Takeuchi, K.P.; Alves, C.C.F.; de Figueiredo Sousa, H.A.; Da Silva, E.R.; Dyszy, F.H.; et al. Active food packaging: Alginate films with cottonseed protein hydrolysates. Food Hydrocoll. 2019, 92, 267–275. [Google Scholar] [CrossRef]
  36. Van Sluisveld, M.A.; Worrell, E. The paradox of packaging optimization—A characterization of packaging source reduction in the Netherlands. Resour. Conserv. Recycl. 2013, 73, 133–142. [Google Scholar] [CrossRef]
  37. Marra, A.; Silvestre, C.; Duraccio, D.; Cimmino, S. Polylactic acid/zinc oxide biocomposite films for food packaging application. Int. J. Biol. Macromol. 2016, 88, 254–262. [Google Scholar] [CrossRef]
  38. Hahladakis, J.N.; Iacovidou, E. Closing the loop on plastic packaging materials: What is quality and how does it affect their circularity? Sci. Total Environ. 2018, 630, 1394–1400. [Google Scholar] [CrossRef]
  39. Van den Oever, M.; Molenveld, K.; van der Zee, M.; Bos, H. Bio-Based and Biodegradable Plastics: Facts and Figures; Wageningen Food & Biobased Research: Wageningen, The Netherlands, 2017. [Google Scholar] [CrossRef] [Green Version]
  40. McMillin, K.W. Advancements in meat packaging. Meat Sci. 2017, 132, 153–162. [Google Scholar] [CrossRef]
  41. Molina-Besch, K.; Wikström, F.; Williams, H. The environmental impact of packaging in food supply chains—does life cycle assessment of food provide the full picture? Int. J. Life Cycle Assess 2019, 24, 37–50. [Google Scholar] [CrossRef] [Green Version]
  42. Wikström, F.; Verghese, K.; Auras, R.; Olsson, A.; Williams, H.; Wever, R.; Grönman, K.; Kvalvåg Pettersen, M.; Møller, H.; Soukka, R. Packaging Strategies That Save Food: A Research Agenda for 2030. J. Ind. Ecol. 2019, 23, 532–540. [Google Scholar] [CrossRef]
  43. Garcia-Arce, J.; Garrido, A.T.G.-P.; Prado-Prado, J.C. Implementing sustainable packaging logistics. An analysis in liquid detergents. Dir. Organ. 2016, 60, 47–56. [Google Scholar]
  44. Ma, X.; Park, C.; Moultrie, J. Factors for eliminating plastic in packaging: The European FMCG experts’ view. J. Clean. Prod. 2020, 256. [Google Scholar] [CrossRef]
  45. Dilkes-Hoffman, L.S.; Lane, J.L.; Grant, T.; Pratt, S.; Lant, P.A.; Laycock, B. Environmental impact of biodegradable food packaging when considering food waste. J. Clean. Prod. 2018, 180, 325–334. [Google Scholar] [CrossRef]
  46. Mulakkal, M.C.; Castillo Castillo, A.; Taylor, A.C.; Blackman, B.R.; Balint, D.S.; Pimenta, S.; Charalambides, M.N. Advancing mechanical recycling of multilayer plastics through finite element modelling and environmental policy. Resour. Conserv. Recycl. 2021, 166. [Google Scholar] [CrossRef]
  47. Statista. Recyclingquote Verschiedener Abfallarten in der EU im Jahr 2017; Statista: Hamburg, Germany, 2020. [Google Scholar]
  48. Statista. Recycling Rate of Plastic Packaging Waste in the European Union (EU-28) from 2006 to 2017; Statista: Hamburg, Germany, 2019. [Google Scholar]
  49. Statista. Recyclingquoten von Altpapier in Europa in den Jahren 1991 bis 2019; Statista: Hamburg, Germany, 2020. [Google Scholar]
  50. European Union. Directive (EU) 2018/852 of the European Parliament and of the Council of 30 May 2018 amending Directive 94/62/EC on packaging and packaging waste. Off. J. Eur. Union 2018, 150, 141. [Google Scholar]
  51. Sun, H.; Guo, G.; Memon, S.A.; Xu, W.; Zhang, Q.; Zhu, J.-H.; Xing, F. Recycling of carbon fibers from carbon fiber reinforced polymer using electrochemical method. Compos. Part A Appl. Sci. Manuf. 2015, 78, 10–17. [Google Scholar] [CrossRef]
  52. Deng, J.; Xu, L.; Liu, J.; Peng, J.; Han, Z.; Shen, Z.; Guo, S. Efficient method of recycling carbon fiber from the waste of carbon fiber reinforced polymer composites. Polym. Degrad. Stab. 2020, 182. [Google Scholar] [CrossRef]
  53. Mumbach, G.D.; Bolzan, A.; Machado, R.A.F. A closed-loop process design for recycling expanded polystyrene waste by dissolution and polymerization. Polymer 2020, 209. [Google Scholar] [CrossRef]
  54. Abdou, T.R.; Botelho, A.B., Jr.; Espinosa, D.C.R.; Tenório, J.A.S. Recycling of polymeric composites from industrial waste by pyrolysis: Deep evaluation for carbon fibers reuse. Waste Manag. 2021, 120, 1–9. [Google Scholar] [CrossRef]
  55. Khui, P.L.N.; Rahman, M.R.; Jayamani, E.; Bin Bakri, M.K. Recycling of sustainable polymers and composites. In Advances in Sustainable Polymer Composites; Elsevier: Amsterdam, The Netherlands, 2021; pp. 267–282. ISBN 9780128203385. [Google Scholar]
  56. Ingrao, C.; Gigli, M.; Siracusa, V. An attributional Life Cycle Assessment application experience to highlight environmental hotspots in the production of foamy polylactic acid trays for fresh-food packaging usage. J. Clean. Prod. 2017, 150, 93–103. [Google Scholar] [CrossRef]
  57. White, A.; Lockyer, S. Removing plastic packaging from fresh produce—What’s the impact? Nutr. Bull. 2020, 45, 35–50. [Google Scholar] [CrossRef]
  58. Bhat, R.; Jõudu, I. Emerging issues and challenges in agri-food supply chain. In Sustainable Food Supply Chains; Elsevier: Amsterdam, The Netherlands, 2019; pp. 23–37. ISBN 9780128134115. [Google Scholar]
  59. United Nations. Transforming our world: The 2030 Agenda for Sustainable Development. In General Assembly; A/Res/70/1; United Nations: New York, NY, USA, 2015. [Google Scholar]
  60. Mak, T.M.W.; Xiong, X.; Tsang, D.C.W.; Yu, I.K.M.; Poon, C.S. Sustainable food waste management towards circular bioeconomy: Policy review, limitations and opportunities. Bioresour. Technol. 2020, 297. [Google Scholar] [CrossRef]
  61. Bruckner, S.; Albrecht, A.; Petersen, B.; Kreyenschmidt, J. Characterization and comparison of spoilage processes in fresh pork and poultry. J. Food Qual. 2012, 35, 372–382. [Google Scholar] [CrossRef]
  62. Herbert, U.; Albrecht, A.; Kreyenschmidt, J. Definition of predictor variables for MAP poultry filets stored under different temperature conditions. Poult. Sci. 2015, 94, 424–432. [Google Scholar] [CrossRef]
  63. Caldeira, C.; de Laurentiis, V.; Corrado, S.; van Holsteijn, F.; Sala, S. Quantification of food waste per product group along the food supply chain in the European Union: A mass flow analysis. Resour. Conserv. Recycl. 2019, 149, 479–488. [Google Scholar] [CrossRef]
  64. Mena, C.; Adenso-Diaz, B.; Yurt, O. The causes of food waste in the supplier–retailer interface: Evidences from the UK and Spain. Resour. Conserv. Recycl. 2011, 55, 648–658. [Google Scholar] [CrossRef]
  65. Rossaint, S.; Kreyenschmidt, J. Intelligent label—A new way to support food waste reduction. Waste Resour. Manag. 2014, 168, 63–71. [Google Scholar] [CrossRef]
  66. Scherhaufer, S.; Moates, G.; Hartikainen, H.; Waldron, K.; Obersteiner, G. Environmental impacts of food waste in Europe. Waste Manag. 2018, 77, 98–113. [Google Scholar] [CrossRef] [PubMed]
  67. Williams, H.; Wikström, F. Environmental impact of packaging and food losses in a life cycle perspective: A comparative analysis of five food items. J. Clean. Prod. 2011, 19, 43–48. [Google Scholar] [CrossRef]
  68. Heller, M.C.; Selke, S.E.M.; Keoleian, G.A. Mapping the influence of food waste in food packaging environmental performance assessments. J. Ind. Ecol. 2019, 23, 480–495. [Google Scholar] [CrossRef] [Green Version]
  69. Verghese, K.; Lewis, H.; Lockrey, S.; Williams, H. Packaging’s role in minimizing food loss and waste across the supply chain. Packag. Technol. Sci. 2015, 28, 603–620. [Google Scholar] [CrossRef]
  70. Steinbuchel, A. Biopolymers: General Aspects and Special Applications; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar]
  71. Conte, A.; Cappelletti, G.M.; Nicoletti, G.M.; Russo, C.; Del Nobile, M.A. Environmental implications of food loss probability in packaging design. Food Res. Int. 2015, 78, 11–17. [Google Scholar] [CrossRef]
  72. Pettersen, M.K.; Grøvlen, M.S.; Evje, N.; Radusin, T. Recyclable mono materials for packaging of fresh chicken fillets: New design for recycling in circular economy. Packag. Technol. Sci. 2020, 33, 485–498. [Google Scholar] [CrossRef]
  73. Zhang, H.; Hortal, M.; Dobon, A.; Bermudez, J.M.; Lara-Lledo, M. The Effect of Active Packaging on Minimizing Food Losses: Life Cycle Assessment (LCA) of Essential Oil Component-enabled Packaging for Fresh Beef. Packag. Technol. Sci. 2015, 28, 761–774. [Google Scholar] [CrossRef]
  74. Van de Nadort, A. Opinion of the European Committee of the Regions—Communication on a European Strategy for Plastics in a circular economy. Off. J. Eur. Union 2018, 461, 30. [Google Scholar]
  75. Worrell, E.; Allwood, J.; Gutowski, T. The Role of Material Efficiency in Environmental Stewardship. Annu. Rev. Environ. Resour. 2016, 41, 575–598. [Google Scholar] [CrossRef] [Green Version]
  76. Verghese, K.; Lewis, H.; Fitzpatrick, L. Packaging for Sustainability; Springer: London, UK, 2012. [Google Scholar] [CrossRef]
  77. Popović, S.Z.; Lazić, V.L.; Hromiš, N.M.; Šuput, D.Z.; Bulut, S.N. Biopolymer Packaging materials for food shelf-life prolongation. In Biopolymers for Food Design; Elsevier: Amsterdam, The Netherlands, 2018; pp. 223–277. ISBN 9780128114490. [Google Scholar]
  78. Xia, C.; Wang, W.; Wang, L.; Liu, H.; Xiao, J. Multilayer zein/gelatin films with tunable water barrier property and prolonged antioxidant activity. Food Packag. Shelf Life 2019, 19, 76–85. [Google Scholar] [CrossRef]
  79. Wang, W.; Gu, F.; Deng, Z.; Zhu, Y.; Zhu, J.; Guo, T.; Song, J.; Xiao, H. Mulitlayer surface construction for enhancing barrier properties of cellulose-based packaging. Carbohydr. Polym. 2021, 255. [Google Scholar] [CrossRef]
  80. Biscarat, J.; Charmette, C.; Sanchez, J.; Pochat-Bohatier, C. Development of a new family of food packaging bioplastics from cross-linked gelatin based films. Can. J. Chem. Eng. 2015, 93, 176–182. [Google Scholar] [CrossRef]
  81. Asgher, M.; Urooj, Y.; Qamar, S.A.; Khalid, N. Improved exopolysaccharide production from Bacillus licheniformis MS3: Optimization and structural/functional characterization. Int. J. Biol. Macromol. 2020, 151, 984–992. [Google Scholar] [CrossRef] [PubMed]
  82. Cazón, P.; Velazquez, G.; Ramírez, J.A.; Vázquez, M. Polysaccharide-based films and coatings for food packaging: A review. Food Hydrocoll. 2017, 68, 136–148. [Google Scholar] [CrossRef]
  83. Kumar, N.; Kaur, P.; Bhatia, S. Advances in bio-nanocomposite materials for food packaging: A review. Nutr. Food Sci. 2017, 47, 591–606. [Google Scholar] [CrossRef]
  84. Thakur, S.; Chaudhary, J.; Sharma, B.; Verma, A.; Tamulevicius, S.; Thakur, V.K. Sustainability of bioplastics: Opportunities and challenges. Curr. Opin. Green Sustain. Chem. 2018, 13, 68–75. [Google Scholar] [CrossRef] [Green Version]
  85. Vejdan, A.; Ojagh, S.M.; Adeli, A.; Abdollahi, M. Effect of TiO2 nanoparticles on the physico-mechanical and ultraviolet light barrier properties of fish gelatin/agar bilayer film. LWT Food Sci. Technol. 2016, 71, 88–95. [Google Scholar] [CrossRef]
  86. Bahrami, A.; Rezaei Mokarram, R.; Sowti Khiabani, M.; Ghanbarzadeh, B.; Salehi, R. Physico-mechanical and antimicrobial properties of tragacanth/hydroxypropyl methylcellulose/beeswax edible films reinforced with silver nanoparticles. Int. J. Biol. Macromol. 2019, 129, 1103–1112. [Google Scholar] [CrossRef]
  87. Davachi, S.M.; Shekarabi, A.S. Preparation and characterization of antibacterial, eco-friendly edible nanocomposite films containing Salvia macrosiphon and nanoclay. Int. J. Biol. Macromol. 2018, 113, 66–72. [Google Scholar] [CrossRef]
  88. Saral Sarojini, K.; Indumathi, M.P.; Rajarajeswari, G.R. Mahua oil-based polyurethane/chitosan/nano ZnO composite films for biodegradable food packaging applications. Int. J. Biol. Macromol. 2019, 124, 163–174. [Google Scholar] [CrossRef]
  89. Ma, P.; Hristova-Bogaerds, D.G.; Goossens, J.; Spoelstra, A.B.; Zhang, Y.; Lemstra, P.J. Toughening of poly(lactic acid) by ethylene-co-vinyl acetate copolymer with different vinyl acetate contents. Eur. Polym. J. 2012, 48, 146–154. [Google Scholar] [CrossRef]
  90. 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] [Green Version]
  91. Youssef, A.M.; Malhat, F.M.; Abdel Hakim, A.A.; Dekany, I. Synthesis and utilization of poly (methylmethacrylate) nanocomposites based on modified montmorillonite. Arab. J. Chem. 2017, 10, 631–642. [Google Scholar] [CrossRef] [Green Version]
  92. Malathi, A.N.; Santhosh, K.S.; Nidoni, U. Recent trends of Biodegradable polymer: Biodegradable films for Food Packaging and application of Nanotechnology in Biodegradable Food Packaging. Curr. Trends Technol. Sci. 2014, 3, 73–79. [Google Scholar]
  93. Asgher, M.; Arshad, S.; Qamar, S.A.; Khalid, N. Improved biosurfactant production from Aspergillus niger through chemical mutagenesis: Characterization and RSM optimization. SN Appl. Sci. 2020, 2. [Google Scholar] [CrossRef] [Green Version]
  94. Ahmed, T.; Shahid, M.; Azeem, F.; Rasul, I.; Shah, A.A.; Noman, M.; Hameed, A.; Manzoor, N.; Manzoor, I.; Muhammad, S. Biodegradation of plastics: Current scenario and future prospects for environmental safety. Environ. Sci. Pollut. Res. 2018, 25, 7287–7298. [Google Scholar] [CrossRef] [PubMed]
  95. European Bioplastics. What are Bioplastics?: Material Types, Terminology, and Labels—An Introduction. Available online: https://docs.european-bioplastics.org/publications/fs/EuBP_FS_What_are_bioplastics.pdf (accessed on 26 February 2021).
  96. WRAP. Understanding Plastic Packaging and the Language We Use to Describe It; WRAP: Banbury, UK, 2018. [Google Scholar]
  97. Lambert, S.; Wagner, M. Environmental performance of bio-based and biodegradable plastics: The road ahead. Chem. Soc. Rev. 2017, 46, 6855–6871. [Google Scholar] [CrossRef] [PubMed]
  98. Asgher, M.; Qamar, S.A.; Bilal, M.; Iqbal, H.M.N. Bio-based active food packaging materials: Sustainable alternative to conventional petrochemical-based packaging materials. Food Res. Int. 2020, 137. [Google Scholar] [CrossRef]
  99. Peelman, N.; Ragaert, P.; Meulenaer, B.d.; Adons, D.; Peeters, R.; Cardon, L.; van Impe, F.; Devlieghere, F. Application of bioplastics for food packaging. Trends Food Sci. Technol. 2013, 32, 128–141. [Google Scholar] [CrossRef] [Green Version]
  100. Saha, T.; Hoque, M.E.; Mahbub, T. Biopolymers for Sustainable Packaging in Food, Cosmetics, and Pharmaceuticals. In Advanced Processing, Properties, and Applications of Starch and Other Bio-Based Polymers; Elsevier: Amsterdam, The Netherlands, 2020; pp. 197–214. ISBN 9780128196618. [Google Scholar]
  101. Kamdem, D.P.; Shen, Z.; Nabinejad, O.; Shu, Z. Development of biodegradable composite chitosan-based films incorporated with xylan and carvacrol for food packaging application. Food Packag. Shelf Life 2019, 21. [Google Scholar] [CrossRef]
  102. Tawakkal, I.S.M.A.; Cran, M.J.; Miltz, J.; Bigger, S.W. A review of poly(lactic acid)-based materials for antimicrobial packaging. J. Food Sci. 2014, 79, 1477–1490. [Google Scholar] [CrossRef]
  103. Rasal, R.M.; Janorkar, A.V.; Hirt, D.E. Poly(lactic acid) modifications. Prog. Polym. Sci. 2010, 35, 338–356. [Google Scholar] [CrossRef]
  104. Byun, Y.; Kim, Y.T. Utilization of Bioplastics for Food Packaging Industry. In Innovations in Food Packaging; Elsevier: Amsterdam, The Netherlands, 2014; pp. 369–390. ISBN 9780123946010. [Google Scholar]
  105. Jabeen, N.; Majid, I.; Nayik, G.A.; Yildiz, F. Bioplastics and food packaging: A review. Cogent Food Agric. 2015, 1. [Google Scholar] [CrossRef]
  106. Zinoviadou, K.G.; Gougouli, M.; Biliaderis, C.G. Innovative biobased materials for packaging sustainability. In Innovation Strategies in the Food Industry; Elsevier: Amsterdam, The Netherlands, 2016; pp. 167–189. ISBN 9780128037515. [Google Scholar]
  107. Auras, R.; Harte, B.; Selke, S. An overview of polylactides as packaging materials. Macromol. Biosci. 2004, 4, 835–864. [Google Scholar] [CrossRef]
  108. Benetto, E.; Jury, C.; Igos, E.; Carton, J.; Hild, P.; Vergne, C.; Di Martino, J. Using atmospheric plasma to design multilayer film from polylactic acid and thermoplastic starch: A screening Life Cycle Assessment. J. Clean. Prod. 2015, 87, 953–960. [Google Scholar] [CrossRef]
  109. Ahmadzadeh, S.; Khaneghah, A.M. Role of Green Polymers in Food Packaging. In Encyclopedia of Renewable and Sustainable Materials; Choudhury, I., Hashmi, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 305–319. ISBN 9780128131961. [Google Scholar]
  110. Ragaert, P.; Buntinx, M.; Maes, C.; Vanheusden, C.; Peeters, R.; Wang, S.; D’hooge, D.R.; Cardon, L. Polyhydroxyalkanoates for Food Packaging Applications. In Reference Module in Food Science; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 9780081005965. [Google Scholar]
  111. Crompton, T.R. Physical Testing of Plastics; Smithers Rapra: Shawbury, UK, 2012; ISBN 978-1-84735-487-7. [Google Scholar]
  112. Lange, J.; Wyser, Y. Recent innovations in barrier technologies for plastic packaging—A review. Packag. Technol. Sci. 2003, 16, 149–158. [Google Scholar] [CrossRef]
  113. Massey, L.K. Permeability Properties of Plastics and Elastomers: A Guide to Packaging and Barrier Materials, 2nd ed.; Elsevier Science: Norwich, UK, 2003; ISBN 9780815518518. [Google Scholar]
  114. Drieskens, M.; Peeters, R.; Mullens, J.; Franco, D.; Lemstra, P.J.; Hristova-Bogaerds, D.G. Structure versus properties relationship of poly(lactic acid). I. Effect of crystallinity on barrier properties. J. Polym. Sci. Part B Polym. Phys. 2009, 47, 2247–2258. [Google Scholar] [CrossRef]
  115. Rhim, J.-W.; Hong, S.-I.; Ha, C.-S. Tensile, water vapor barrier and antimicrobial properties of PLA/nanoclay composite films. LWT Food Sci. Technol. 2009, 42, 612–617. [Google Scholar] [CrossRef]
  116. Raquez, J.-M.; Habibi, Y.; Murariu, M.; Dubois, P. Polylactide (PLA)-based nanocomposites. Prog. Polym. Sci. 2013, 38, 1504–1542. [Google Scholar] [CrossRef]
  117. Panseri, S.; Martino, P.A.; Cagnardi, P.; Celano, G.; Tedesco, D.; Castrica, M.; Balzaretti, C.; Chiesa, L.M. Feasibility of biodegradable based packaging used for red meat storage during shelf-life: A pilot study. Food Chem. 2018, 249, 22–29. [Google Scholar] [CrossRef]
  118. Vanitha, R.; Kavitha, C. Development of natural cellulose fiber and its food packaging application. Mater. Today Proc. 2021, 36, 903–906. [Google Scholar] [CrossRef]
  119. Domenek, S.; Louaifi, A.; Guinault, A.; Baumberger, S. Potential of lignins as antioxidant additive in active biodegradable packaging materials. J. Polym. Environ. 2013, 21, 692–701. [Google Scholar] [CrossRef] [Green Version]
  120. Spiridon, I.; Tanase, C.E. Design, characterization and preliminary biological evaluation of new lignin-PLA biocomposites. Int. J. Biol. Macromol. 2018, 114, 855–863. [Google Scholar] [CrossRef]
  121. Gordobil, O.; Egüés, I.; Llano-Ponte, R.; Labidi, J. Physicochemical properties of PLA lignin blends. Polym. Degrad. Stab. 2014, 108, 330–338. [Google Scholar] [CrossRef]
  122. Kovalcik, A.; Pérez-Camargo, R.A.; Fürst, C.; Kucharczyk, P.; Müller, A.J. Nucleating efficiency and thermal stability of industrial non-purified lignins and ultrafine talc in poly(lactic acid) (PLA). Polym. Degrad. Stab. 2017, 142, 244–254. [Google Scholar] [CrossRef]
  123. Kumar, A.; Tumu, V.R.; Chowdhury, S.R.; Ramana Reddy, S.V.S. A green physical approach to compatibilize a bio-based poly (lactic acid)/lignin blend for better mechanical, thermal and degradation properties. Int. J. Biol. Macromol. 2019, 121, 588–600. [Google Scholar] [CrossRef] [PubMed]
  124. Li, X.; Hegyesi, N.; Zhang, Y.; Mao, Z.; Feng, X.; Wang, B.; Pukánszky, B.; Sui, X. Poly(lactic acid)/lignin blends prepared with the Pickering emulsion template method. Eur. Polym. J. 2019, 110, 378–384. [Google Scholar] [CrossRef] [Green Version]
  125. Yang, W.; Fortunati, E.; Dominici, F.; Giovanale, G.; Mazzaglia, A.; Balestra, G.M.; Kenny, J.M.; Puglia, D. Effect of cellulose and lignin on disintegration, antimicrobial and antioxidant properties of PLA active films. Int. J. Biol. Macromol. 2016, 89, 360–368. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, W.; Fortunati, E.; Dominici, F.; Giovanale, G.; Mazzaglia, A.; Balestra, G.M.; Kenny, J.M.; Puglia, D. Synergic effect of cellulose and lignin nanostructures in PLA based systems for food antibacterial packaging. Eur. Polym. J. 2016, 79, 1–12. [Google Scholar] [CrossRef]
  127. Wei, L.; Agarwal, U.P.; Matuana, L.; Sabo, R.C.; Stark, N.M. Performance of high lignin content cellulose nanocrystals in poly(lactic acid). Polymer 2018, 135, 305–313. [Google Scholar] [CrossRef]
  128. Armentano, I.; Fortunati, E.; Burgos, N.; Dominici, F.; Luzi, F.; Fiori, S.; Jiménez, A.; Yoon, K.; Ahn, J.; Kang, S.; et al. Bio-based PLA_PHB plasticized blend films: Processing and structural characterization. LWT Food Sci. Technol. 2015, 64, 980–988. [Google Scholar] [CrossRef] [Green Version]
  129. Andersson, C. New ways to enhance the functionality of paperboard by surface treatment—A review. Packag. Technol. Sci. 2008, 21, 339–373. [Google Scholar] [CrossRef]
  130. Alavi, S.; Thomas, S.; Sandeep, K.P.; Kalarikkal, N.; Varghese, J.; Yaragalla, S. Polymers for Packaging Applications; Apple Academic Press: Palm Bay, FL, USA, 2014. [Google Scholar]
  131. Siracusa, V.; Blanco, I.; Romani, S.; Tylewicz, U.; Rocculi, P.; Rosa, M.D. Poly(lactic acid)-modified films for food packaging application: Physical, mechanical, and barrier behavior. J. Appl. Polym. Sci. 2012, 125, E390–E401. [Google Scholar] [CrossRef]
  132. Vandewijngaarden, J.; Murariu, M.; Dubois, P.; Carleer, R.; Yperman, J.; Adriaensens, P.; Schreurs, S.; Lepot, N.; Peeters, R.; Buntinx, M. Gas Permeability Properties of Poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). J. Polym. Environ. 2014, 22, 501–507. [Google Scholar] [CrossRef]
  133. Kovalcik, A.; Machovsky, M.; Kozakova, Z.; Koller, M. Designing packaging materials with viscoelastic and gas barrier properties by optimized processing of poly(3-hydroxybutyrate-co-3-hydroxyvalerate) with lignin. React. Funct. Polym. 2015, 94, 25–34. [Google Scholar] [CrossRef]
  134. Farmahini-Farahani, M.; Khan, A.; Lu, P.; Bedane, A.H.; Eic, M.; Xiao, H. Surface morphological analysis and water vapor barrier properties of modified Cloisite 30B/poly(3-hydroxybutyrate-co-3-hydroxyvalerate) composites. Appl. Clay Sci. 2017, 135, 27–34. [Google Scholar] [CrossRef]
  135. Siracusa, V.; Ingrao, C.; Karpova, S.G.; Olkhov, A.A.; Iordanskii, A.L. Gas transport and characterization of poly(3 hydroxybutyrate) films. Eur. Polym. J. 2017, 91, 149–161. [Google Scholar] [CrossRef]
  136. Maes, C.; Luyten, W.; Herremans, G.; Peeters, R.; Carleer, R.; Buntinx, M. Recent updates on the barrier properties of ethylene vinyl alcohol copolymer (EVOH): A review. Polym. Rev. 2018, 58, 209–246. [Google Scholar] [CrossRef] [Green Version]
  137. Anjum, A.; Zuber, M.; Zia, K.M.; Noreen, A.; Anjum, M.N.; Tabasum, S. Microbial production of polyhydroxyalkanoates (PHAs) and its copolymers: A review of recent advancements. Int. J. Biol. Macromol. 2016, 89, 161–174. [Google Scholar] [CrossRef]
  138. Arrieta, M.P.; Samper, M.D.; López, J.; Jiménez, A. Combined effect of poly(hydroxybutyrate) and plasticizers on polylactic acid properties for film intended for food packaging. J. Polym. Environ. 2014, 22, 460–470. [Google Scholar] [CrossRef]
  139. Schutyser, W.; Renders, T.; van den Bosch, S.; Koelewijn, S.-F.; Beckham, G.T.; Sels, B.F. Chemicals from lignin: An interplay of lignocellulose fractionation, depolymerisation, and upgrading. Chem. Soc. Rev. 2018, 47, 852–908. [Google Scholar] [CrossRef]
  140. Rumpf, J.; Do, X.T.; Burger, R.; Monakhova, Y.B.; Schulze, M. Extraction of high-purity lignins via catalyst-free organosolv pulping from low-input crops. Biomacromolecules 2020, 21, 1929–1942. [Google Scholar] [CrossRef]
  141. Rinaldi, R.; Jastrzebski, R.; Clough, M.T.; Ralph, J.; Kennema, M.; Bruijnincx, P.C.A.; Weckhuysen, B.M. Paving the way for lignin valorisation: Recent advances in bioengineering, biorefining and catalysis. Angew. Chem. Int. Ed. 2016, 55, 8164–8215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Xu, C.; Ferdosian, F. Conversion of Lignin into Bio-Based Chemicals and Materials; Springer: Berlin/Heidelberg, Germany, 2017; ISBN 978-3-662-54957-5. [Google Scholar]
  143. Tribot, A.; Amer, G.; Abdou Alio, M.; Baynast, H.d.; Delattre, C.; Pons, A.; Mathias, J.-D.; Callois, J.-M.; Vial, C.; Michaud, P.; et al. Wood-lignin: Supply, extraction processes and use as bio-based material. Eur. Polym. J. 2019, 112, 228–240. [Google Scholar] [CrossRef]
  144. Kim, K.H.; Jeong, K.; Zhuang, J.; Jeong, H.J.; Kim, C.S.; Koo, B.; Yoo, C.G. Tandem conversion of lignin to catechols via demethylation and catalytic hydrogenolysis. Ind. Crop. Prod. 2021, 159. [Google Scholar] [CrossRef]
  145. Hassan, S.S.; Williams, G.A.; Jaiswal, A.K. Moving towards the second generation of lignocellulosic biorefineries in the EU: Drivers, challenges, and opportunities. Renew. Sustain. Energy Rev. 2019, 101, 590–599. [Google Scholar] [CrossRef] [Green Version]
  146. Renders, T.; van den Bossche, G.; Vangeel, T.; van Aelst, K.; Sels, B. Reductive catalytic fractionation: State of the art of the lignin-first biorefinery. Curr. Opin. Biotechnol. 2019, 56, 193–201. [Google Scholar] [CrossRef] [PubMed]
  147. Espinoza-Acosta, J.L.; Torres-Chávez, P.I.; Ramírez-Wong, B.; López-Saiz, C.M.; Montaño-Leyva, B. Antioxidant, antimicrobial, and antimutagenic properties of technical lignins and their applications. BioResources 2016, 11, 5452–5481. [Google Scholar] [CrossRef]
  148. Azadfar, M.; Gao, A.H.; Bule, M.V.; Chen, S. Structural characterization of lignin: A potential source of antioxidants guaiacol and 4-vinylguaiacol. Int. J. Biol. Macromol. 2015, 75, 58–66. [Google Scholar] [CrossRef]
  149. Michelin, M.; Marques, A.M.; Pastrana, L.M.; Teixeira, J.A.; Cerqueira, M.A. Carboxymethyl cellulose-based films: Effect of organosolv lignin incorporation on physicochemical and antioxidant properties. J. Food Eng. 2020, 285. [Google Scholar] [CrossRef]
  150. Hult, E.-L.; Ropponen, J.; Poppius-Levlin, K.; Ohra-Aho, T.; Tamminen, T. Enhancing the barrier properties of paper board by a novel lignin coating. Ind. Crop. Prod. 2013, 50, 694–700. [Google Scholar] [CrossRef]
  151. Klein, S.E.; Rumpf, J.; Alzagameem, A.; Rehahn, M.; Schulze, M. Antioxidant activity of unmodified kraft and organosolv lignins to be used as sustainable components for polyurethane coatings. J. Coat. Technol. Res. 2019, 55, 97. [Google Scholar] [CrossRef]
  152. Klein, S.E.; Alzagameem, A.; Rumpf, J.; Korte, I.; Kreyenschmidt, J.; Schulze, M. Antimicrobial activity of lignin-derived polyurethane coatings prepared from unmodified and demethylated lignins. Coatings 2019, 9, 494. [Google Scholar] [CrossRef] [Green Version]
  153. Hao, C.; Liu, T.; Zhang, S.; Brown, L.; Li, R.; Xin, J.; Zhong, T.; Jiang, L.; Zhang, J. A High-lignin-content, removable, and glycol-assisted repairable coating based on dynamic covalent bonds. ChemSusChem 2019, 1049–1058. [Google Scholar] [CrossRef] [PubMed]
  154. Hambardzumyan, A.; Foulon, L.; Bercu, N.B.; Pernes, M.; Maigret, J.E.; Molinari, M.; Chabbert, B.; Aguié-Béghin, V. Organosolv lignin as natural grafting additive to improve the water resistance of films using cellulose nanocrystals. Chem. Eng. J. 2015, 264, 780–788. [Google Scholar] [CrossRef]
  155. Rastogi, V.; Samyn, P. Bio-based coatings for paper applications. Coatings 2015, 5, 887–930. [Google Scholar] [CrossRef] [Green Version]
  156. Flory, A.R.; Vicuna Requesens, D.; Devaiah, S.P.; Teoh, K.T.; Mansfield, S.D.; Hood, E.E. Development of a green binder system for paper products. BMC Biotechnol. 2013, 13, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Jiang, B.; Chen, C.; Liang, Z.; He, S.; Kuang, Y.; Song, J.; Mi, R.; Chen, G.; Jiao, M.; Hu, L. Lignin as a wood-inspired binder enabled strong, water stable, and biodegradable paper for plastic replacement. Adv. Funct. Mater. 2020, 30, 1906307. [Google Scholar] [CrossRef]
  158. Alzagameem, A.; El Khaldi-Hansen, B.; Kamm, B.; Schulze, M. Lignocellulosic biomass for energy, biofuels, biomaterials, and chemicals. In Biomass and Green Chemistry: Building a Renewable Pathway; Vaz, S., Jr., Ed.; Springer International Publishing: Basel, Switzerland, 2018; pp. 95–132. ISBN 978-3-319-66736-2. [Google Scholar]
  159. Alzagameem, A.; Khaldi-Hansen, B.E.; Büchner, D.; Larkins, M.; Kamm, B.; Witzleben, S.; Schulze, M. Lignocellulosic biomass as source for lignin-based environmentally benign antioxidants. Molecules 2018, 23. [Google Scholar] [CrossRef] [Green Version]
  160. Alzagameem, A.; Klein, S.E.; Bergs, M.; Do, X.T.; Korte, I.; Dohlen, S.; Hüwe, C.; Kreyenschmidt, J.; Kamm, B.; Larkins, M.; et al. Antimicrobial activity of lignin and lignin-derived cellulose and chitosan composites against selected pathogenic and spoilage microorganisms. Polymers 2019, 11. [Google Scholar] [CrossRef] [Green Version]
  161. Galus, S.; Kadzińska, J. Food applications of emulsion-based edible films and coatings. Trends Food Sci. Technol. 2015, 45, 273–283. [Google Scholar] [CrossRef]
  162. Otoni, C.G.; Avena-Bustillos, R.J.; Olsen, C.W.; Bilbao-Sáinz, C.; McHugh, T.H. Mechanical and water barrier properties of isolated soy protein composite edible films as affected by carvacrol and cinnamaldehyde micro and nanoemulsions. Food Hydrocoll. 2016, 57, 72–79. [Google Scholar] [CrossRef]
  163. Cho, S.Y.; Lee, S.Y.; Rhee, C. Edible oxygen barrier bilayer film pouches from corn zein and soy protein isolate for olive oil packaging. LWT Food Sci. Technol. 2010, 43, 1234–1239. [Google Scholar] [CrossRef]
  164. Denavi, G.; Tapia-Blácido, D.R.; Añón, M.C.; Sobral, P.; Mauri, A.N.; Menegalli, F.C. Effects of drying conditions on some physical properties of soy protein films. J. Food Eng. 2009, 90, 341–349. [Google Scholar] [CrossRef]
  165. Popović, S.; Peričin, D.; Vaštag, Ž.; Lazić, V.; Popović, L. Pumpkin oil cake protein isolate films as potential gas barrier coating. J. Food Eng. 2012, 110, 374–379. [Google Scholar] [CrossRef]
  166. Umaraw, P.; Verma, A.K. Comprehensive review on application of edible film on meat and meat products: An eco-friendly approach. Crit. Rev. Food Sci. Nutr. 2017, 57, 1270–1279. [Google Scholar] [CrossRef]
  167. Zhang, Y.; Liu, Q.; Rempel, C. Processing and characteristics of canola protein-based biodegradable packaging: A review. Crit. Rev. Food Sci. Nutr. 2018, 58, 475–485. [Google Scholar] [CrossRef] [PubMed]
  168. Acquah, C.; Zhang, Y.; Dubé, M.A.; Udenigwe, C.C. Formation and characterization of protein-based films from yellow pea (Pisum sativum) protein isolate and concentrate for edible applications. Curr. Res. Food Sci. 2020, 2, 61–69. [Google Scholar] [CrossRef]
  169. Wang, Y.; Xiong, Y.L.; Rentfrow, G.K.; Newman, M.C. Oxidation promotes cross-linking but impairs film-forming properties of whey proteins. J. Food Eng. 2013, 115, 11–19. [Google Scholar] [CrossRef]
  170. Ramos, Ó.L.; Silva, S.I.; Soares, J.C.; Fernandes, J.C.; Poças, M.F.; Pintado, M.E.; Malcata, F.X. Features and performance of edible films, obtained from whey protein isolate formulated with antimicrobial compounds. Food Res. Int. 2012, 45, 351–361. [Google Scholar] [CrossRef] [Green Version]
  171. Fratzl, P. Collagen: Structure and Mechanics, an Introduction. In Collagen; Fratzl, P., Ed.; Springer: New York, NY, USA, 2008; pp. 1–13. ISBN 978-0-387-73906-9. [Google Scholar]
  172. Fadini, A.L.; Rocha, F.S.; Alvim, I.D.; Sadahira, M.S.; Queiroz, M.B.; Alves, R.; Silva, L.B. Mechanical properties and water vapour permeability of hydrolysed collagen–cocoa butter edible films plasticised with sucrose. Food Hydrocoll. 2013, 30, 625–631. [Google Scholar] [CrossRef]
  173. Oechsle, A.M.; Bugbee, T.J.; Gibis, M.; Kohlus, R.; Weiss, J. Modification of extruded chicken collagen films by addition of co-gelling protein and sodium chloride. J. Food Eng. 2017, 207, 46–55. [Google Scholar] [CrossRef]
  174. Ciannamea, E.M.; Castillo, L.A.; Barbosa, S.E.; Angelis, M.G.d. Barrier properties and mechanical strength of bio-renewable, heat-sealable films based on gelatin, glycerol and soybean oil for sustainable food packaging. React. Funct. Polym. 2018, 125, 29–36. [Google Scholar] [CrossRef]
  175. Chentir, I.; Kchaou, H.; Hamdi, M.; Jridi, M.; Li, S.; Doumandji, A.; Nasri, M. Biofunctional gelatin-based films incorporated with food grade phycocyanin extracted from the Saharian cyanobacterium Arthrospira sp. Food Hydrocoll. 2019, 89, 715–725. [Google Scholar] [CrossRef]
  176. Amjadi, S.; Hamishehkar, H.; Ghorbani, M. A novel smart PEGylated gelatin nanoparticle for co-delivery of doxorubicin and betanin: A strategy for enhancing the therapeutic efficacy of chemotherapy. Mater. Sci. Eng. C 2019, 97, 833–841. [Google Scholar] [CrossRef] [PubMed]
  177. Nur Hanani, Z.A.; Roos, Y.H.; Kerry, J.P. Use and application of gelatin as potential biodegradable packaging materials for food products. Int. J. Biol. Macromol. 2014, 71, 94–102. [Google Scholar] [CrossRef]
  178. Mojumdar, S.C.; Moresoli, C.; Simon, L.C.; Legge, R.L. Edible wheat gluten (WG) protein films. J. Therm. Anal. Calorim. 2011, 104, 929–936. [Google Scholar] [CrossRef]
  179. Kanatt, S.R. Development of active/intelligent food packaging film containing Amaranthus leaf extract for shelf life extension of chicken/fish during chilled storage. Food Packag. Shelf Life 2020, 24, 100506. [Google Scholar] [CrossRef]
  180. Li, Y.; Chen, H.; Dong, Y.; Li, K.; Li, L.; Li, J. Carbon nanoparticles/soy protein isolate bio-films with excellent mechanical and water barrier properties. Ind. Crop. Prod. 2016, 82, 133–140. [Google Scholar] [CrossRef]
  181. Zhang, X.; Zhao, Y.; Li, Y.; Zhu, L.; Fang, Z.; Shi, Q. Physicochemical, mechanical and structural properties of composite edible films based on whey protein isolate/psyllium seed gum. Int. J. Biol. Macromol. 2020, 153, 892–901. [Google Scholar] [CrossRef]
  182. Syahida, N.; Fitry, I.; Zuriyati, A.; Hanani, N. Effects of palm wax on the physical, mechanical and water barrier properties of fish gelatin films for food packaging application. Food Packag. Shelf Life 2020, 23, 100437. [Google Scholar] [CrossRef]
  183. Nur Hanani, Z.A.; O’Mahony, J.A.; Roos, Y.H.; Oliveira, P.M.; Kerry, J.P. Extrusion of gelatin-based composite films: Effects of processing temperature and pH of film forming solution on mechanical and barrier properties of manufactured films. Food Packag. Shelf Life 2014, 2, 91–101. [Google Scholar] [CrossRef]
  184. Han, J.H. Edible Films and Coatings. In Innovations in Food Packaging; Elsevier: Amsterdam, The Netherlands, 2014; pp. 213–255. ISBN 9780123946010. [Google Scholar]
  185. Jiménez, A.; Fabra, M.J.; Talens, P.; Chiralt, A. Edible and biodegradable starch films: A review. Food Bioprocess Technol. 2012, 5, 2058–2076. [Google Scholar] [CrossRef]
  186. 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]
  187. Xu, Q.; Chen, C.; Rosswurm, K.; Yao, T.; Janaswamy, S. A facile route to prepare cellulose-based films. Carbohydr. Polym. 2016, 149, 274–281. [Google Scholar] [CrossRef]
  188. Witzler, M.; Büchner, D.; Shoushrah, S.H.; Babczyk, P.; Baranova, J.; Witzleben, S.; Tobiasch, E.; Schulze, M. Polysaccharide-Based Systems for Targeted Stem Cell Differentiation and Bone Regeneration. Biomolecules 2019, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Muscat, D.; Adhikari, B.; Adhikari, R.; Chaudhary, D.S. Comparative study of film forming behaviour of low and high amylose starches using glycerol and xylitol as plasticizers. J. Food Eng. 2012, 109, 189–201. [Google Scholar] [CrossRef]
  190. Dang, K.M.; Yoksan, R. Development of thermoplastic starch blown film by incorporating plasticized chitosan. Carbohydr. Polym. 2015, 115, 575–581. [Google Scholar] [CrossRef] [PubMed]
  191. Dang, K.M.; Yoksan, R. Morphological characteristics and barrier properties of thermoplastic starch/chitosan blown film. Carbohydr. Polym. 2016, 150, 40–47. [Google Scholar] [CrossRef] [PubMed]
  192. Nazan Turhan, K.; Şahbaz, F. Water vapor permeability, tensile properties and solubility of methylcellulose-based edible films. J. Food Eng. 2004, 61, 459–466. [Google Scholar] [CrossRef]
  193. Hay, W.T.; Fanta, G.F.; Peterson, S.C.; Thomas, A.J.; Utt, K.D.; Walsh, K.A.; Boddu, V.M.; Selling, G.W. Improved hydroxypropyl methylcellulose (HPMC) films through incorporation of amylose-sodium palmitate inclusion complexes. Carbohydr. Polym. 2018, 188, 76–84. [Google Scholar] [CrossRef]
  194. Jannatyha, N.; Shojaee-Aliabadi, S.; Moslehishad, M.; Moradi, E. Comparing mechanical, barrier and antimicrobial properties of nanocellulose/CMC and nanochitosan/CMC composite films. Int. J. Biol. Macromol. 2020, 164, 2323–2328. [Google Scholar] [CrossRef] [PubMed]
  195. Shankar, S.; Reddy, J.P.; Rhim, J.-W. Effect of lignin on water vapor barrier, mechanical, and structural properties of agar/lignin composite films. Int. J. Biol. Macromol. 2015, 81, 267–273. [Google Scholar] [CrossRef]
  196. Roy, S.; Rhim, J.-W. Effect of CuS reinforcement on the mechanical, water vapor barrier, UV-light barrier, and antibacterial properties of alginate-based composite films. Int. J. Biol. Macromol. 2020, 164, 37–44. [Google Scholar] [CrossRef]
  197. Kurek, M.; Galus, S.; Debeaufort, F. Surface, mechanical and barrier properties of bio-based composite films based on chitosan and whey protein. Food Packag. Shelf Life 2014, 1, 56–67. [Google Scholar] [CrossRef]
  198. Hassan, B.; Chatha, S.A.S.; Hussain, A.I.; Zia, K.M.; Akhtar, N. Recent advances on polysaccharides, lipids and protein based edible films and coatings: A review. Int. J. Biol. Macromol. 2018, 109, 1095–1107. [Google Scholar] [CrossRef] [PubMed]
  199. Molavi, H.; Behfar, S.; Ali Shariati, M.; Kaviani, M.; Atarod, S. A review on biodegradable starch based film. J. Microbiol. Biotechnol. Food Sci. 2015, 4, 456–461. [Google Scholar] [CrossRef] [Green Version]
  200. Lumdubwong, N. Applications of Starch-Based Films in Food Packaging. In Reference Module in Food Science; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 9780081005965. [Google Scholar]
  201. Do Val Siqueira, L.; La Arias, C.I.F.; Maniglia, B.C.; Tadini, C.C. Starch-based biodegradable plastics: Methods of production, challenges and future perspectives. Curr. Opin. Food Sci. 2021, 38, 122–130. [Google Scholar] [CrossRef]
  202. Fakhouri, F.M.; Martelli, S.M.; Caon, T.; Velasco, J.I.; Mei, L.H.I. Edible films and coatings based on starch/gelatin: Film properties and effect of coatings on quality of refrigerated Red Crimson grapes. Postharvest Biol. Technol. 2015, 109, 57–64. [Google Scholar] [CrossRef]
  203. Thakur, R.; Pristijono, P.; Scarlett, C.J.; Bowyer, M.; Singh, S.P.; Vuong, Q.V. Starch-based films: Major factors affecting their properties. Int. J. Biol. Macromol. 2019, 132, 1079–1089. [Google Scholar] [CrossRef]
  204. Ojogbo, E.; Ogunsona, E.O.; Mekonnen, T.H. Chemical and physical modifications of starch for renewable polymeric materials. Mater. Today Sustain. 2020, 7, 100028. [Google Scholar] [CrossRef]
  205. Bhat, R.; Abdullah, N.; Din, R.H.; Tay, G.-S. Producing novel sago starch based food packaging films by incorporating lignin isolated from oil palm black liquor waste. J. Food Eng. 2013, 119, 707–713. [Google Scholar] [CrossRef]
  206. Miranda, C.S.; Ferreira, M.S.; Magalhães, M.T.; Bispo, A.P.G.; Oliveira, J.C.; Silva, J.B.; José, N.M. Starch-based films plasticized with glycerol and lignin from piassava fiber reinforced with nanocrystals from eucalyptus. Mater. Today Proc. 2015, 2, 134–140. [Google Scholar] [CrossRef]
  207. Miranda, C.S.; Ferreira, M.S.; Magalhães, M.T.; Santos, W.J.; Oliveira, J.C.; Silva, J.B.; José, N.M. Mechanical, thermal and barrier properties of starch-based films plasticized with glycerol and lignin and reinforced with cellulose nanocrystals. Mater. Today Proc. 2015, 2, 63–69. [Google Scholar] [CrossRef]
  208. Javed, A.; Ullsten, H.; Rättö, P.; Järnström, L. Lignin-containing coatings for packaging materials. Nord. Pulp Pap. Res. J. 2018, 33, 548–556. [Google Scholar] [CrossRef]
  209. Wang, S.; Lu, A.; Zhang, L. Recent advances in regenerated cellulose materials. Prog. Polym. Sci. 2016, 53, 169–206. [Google Scholar] [CrossRef]
  210. Guzman-Puyol, S.; Ceseracciu, L.; Tedeschi, G.; Marras, S.; Scarpellini, A.; Benítez, J.J.; Athanassiou, A.; Heredia-Guerrero, J.A. Transparent and robust all-cellulose nanocomposite packaging materials prepared in a mixture of trifluoroacetic acid and trifluoroacetic anhydride. Nanomaterials 2019, 9. [Google Scholar] [CrossRef] [Green Version]
  211. Mendes, F.R.; Bastos, M.S.; Mendes, L.G.; Silva, A.R.; Sousa, F.D.; Monteiro-Moreira, A.C.; Cheng, H.N.; Biswas, A.; Moreira, R.A. Preparation and evaluation of hemicellulose films and their blends. Food Hydrocoll. 2017, 70, 181–190. [Google Scholar] [CrossRef] [Green Version]
  212. Zhang, X.; Xiao, N.; Chen, M.; Wei, Y.; Liu, C. Functional packaging films originating from hemicelluloses laurate by direct transesterification in ionic liquid. Carbohydr. Polym. 2020, 229, 115336. [Google Scholar] [CrossRef]
  213. Dhall, R.K. Advances in edible coatings for fresh fruits and vegetables: A review. Crit. Rev. Food Sci. Nutr. 2013, 53, 435–450. [Google Scholar] [CrossRef]
  214. Bifani, V.; Ramírez, C.; Ihl, M.; Rubilar, M.; García, A.; Zaritzky, N. Effects of murta (Ugni molinae Turcz) extract on gas and water vapor permeability of carboxymethylcellulose-based edible films. LWT Food Sci. Technol. 2007, 40, 1473–1481. [Google Scholar] [CrossRef]
  215. Farshchi, E.; Pirsa, S.; Roufegarinejad, L.; Alizadeh, M.; Rezazad, M. Photocatalytic/biodegradable film based on carboxymethyl cellulose, modified by gelatin and TiO2-Ag nanoparticles. Carbohydr. Polym. 2019, 216, 189–196. [Google Scholar] [CrossRef]
  216. Hasheminya, S.-M.; Rezaei Mokarram, R.; Ghanbarzadeh, B.; Hamishekar, H.; Kafil, H.S. Physicochemical, mechanical, optical, microstructural and antimicrobial properties of novel kefiran-carboxymethyl cellulose biocomposite films as influenced by copper oxide nanoparticles (CuONPs). Food Packag. Shelf Life 2018, 17, 196–204. [Google Scholar] [CrossRef]
  217. Arnon, H.; Zaitsev, Y.; Porat, R.; Poverenov, E. Effects of carboxymethyl cellulose and chitosan bilayer edible coating on postharvest quality of citrus fruit. Postharvest Biol. Technol. 2014, 87, 21–26. [Google Scholar] [CrossRef]
  218. Lan, W.; He, L.; Liu, Y. Preparation and properties of sodium carboxymethyl cellulose/sodium alginate/chitosan composite film. Coatings 2018, 8, 291. [Google Scholar] [CrossRef] [Green Version]
  219. Šešlija, S.; Nešić, A.; Škorić, M.L.; Krušić, M.K.; Santagata, G.; Malinconico, M. Pectin/carboxymethylcellulose films as a potential food packaging material. Macromol. Symp. 2018, 378. [Google Scholar] [CrossRef]
  220. Moura, M.R.d.; Avena-Bustillos, R.J.; McHugh, T.H.; Krochta, J.M.; Mattoso, L.H.C. Properties of novel hydroxypropyl methylcellulose films containing chitosan nanoparticles. J. Food Sci. 2008, 73, 31–37. [Google Scholar] [CrossRef] [PubMed]
  221. Sebti, I.; Chollet, E.; Degraeve, P.; Noel, C.; Peyrol, E. Water sensitivity, antimicrobial, and physicochemical analyses of edible films based on HPMC and/or chitosan. J. Agric. Food Chem. 2007, 55, 693–699. [Google Scholar] [CrossRef] [PubMed]
  222. Küçük, G.S.; Çelik, Ö.F.; Mazi, B.G.; Türe, H. Evaluation of alginate and zein films as a carrier of natamycin to increase the shelf life of kashar cheese. Packag. Technol. Sci. 2020, 33, 39–48. [Google Scholar] [CrossRef]
  223. Ahmed, N.M.; Adel, A.M.; Diab, M.A. Packaging paper with improved mechanical and oil absorption properties based on novel ingredients. Packag. Technol. Sci. 2020, 33, 303–320. [Google Scholar] [CrossRef]
  224. Gammariello, D.; Incoronato, A.L.; Conte, A.; Del Nobile, M.A. Effect of sodium alginate coating with ascorbic acid on shelf life of raw pork meat. J. Food Technol. Res. 2016, 3, 1–11. [Google Scholar] [CrossRef] [Green Version]
  225. Sangsuwan, J.; Sutthasupa, S. Effect of chitosan and alginate beads incorporated with lavender, clove essential oils, and vanillin against Botrytis cinerea and their application in fresh table grapes packaging system. Packag. Technol. Sci. 2019, 32, 595–605. [Google Scholar] [CrossRef]
  226. Raafat, D.; Sahl, H.-G. Chitosan and its antimicrobial potential—A critical literature survey. Microb. Biotechnol. 2009, 2, 186–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Morganti, P.; Stoller, M. Chitin and lignin: Natural ingredients from waste materials to make innovative and healthy products for humans and plant. Chem. Eng. Trans. 2017, 60, 319–324. [Google Scholar] [CrossRef]
  228. Witzler, M.; Alzagameem, A.; Bergs, M.; Khaldi-Hansen, B.E.; Klein, S.E.; Hielscher, D.; Kamm, B.; Kreyenschmidt, J.; Tobiasch, E.; Schulze, M. Lignin-derived biomaterials for drug release and tissue engineering. Molecules 2018, 23. [Google Scholar] [CrossRef] [Green Version]
  229. Rai, S.; Dutta, P.K.; Mehrotra, G.K. Lignin incorporated antimicrobial chitosan film for food packaging application. J. Polym. Mater. 2017, 34, 171–183. [Google Scholar]
  230. Shahid-Ul-Islam; Butola, B.S. Recent advances in chitosan polysaccharide and its derivatives in antimicrobial modification of textile materials. Int. J. Biol. Macromol. 2019, 121, 905–912. [Google Scholar] [CrossRef]
  231. Haghighi, H.; Licciardello, F.; Fava, P.; Siesler, H.W.; Pulvirenti, A. Recent advances on chitosan-based films for sustainable food packaging applications. Food Packag. Shelf Life 2020, 26, 100551. [Google Scholar] [CrossRef]
  232. Lin, L.; Xue, L.; Duraiarasan, S.; Haiying, C. Preparation of ε-polylysine/chitosan nanofibers for food packaging against Salmonella on chicken. Food Packag. Shelf Life 2018, 17, 134–141. [Google Scholar] [CrossRef]
  233. Sharaf, O.M.; Al-Gamal, M.S.; Ibrahim, G.A.; Dabiza, N.M.; Salem, S.S.; El-Ssayad, M.F.; Youssef, A.M. Evaluation and characterization of some protective culture metabolites in free and nano-chitosan-loaded forms against common contaminants of Egyptian cheese. Carbohydr. Polym. 2019, 223. [Google Scholar] [CrossRef]
  234. Priyadarshi, R.; Rhim, J.-W. Chitosan-based biodegradable functional films for food packaging applications. Innov. Food Sci. Emerg. Technol. 2020, 62. [Google Scholar] [CrossRef]
  235. Shen, Z.; Kamdem, D.P. Development and characterization of biodegradable chitosan films containing two essential oils. Int. J. Biol. Macromol. 2015, 74, 289–296. [Google Scholar] [CrossRef]
  236. Aider, M. Chitosan application for active bio-based films production and potential in the food industry: Review. LWT Food Sci. Technol. 2010, 43, 837–842. [Google Scholar] [CrossRef]
  237. Chen, L.; Tang, C.; Ning, N.; Wang, C.; Fu, Q.; Zhang, Q. Preparation and properties of chitosan/lignin composite films. Chin. J. Polym. Sci. 2009, 27, 739. [Google Scholar] [CrossRef]
  238. Crouvisier-Urion, K.; Bodart, P.R.; Winckler, P.; Raya, J.; Gougeon, R.D.; Cayot, P.; Domenek, S.; Debeaufort, F.; Karbowiak, T. Bio-based composite films from chitosan and lignin: Antioxidant activity related to structure and moisture. ACS Sustain. Chem. Eng. 2016, 4, 6371–6381. [Google Scholar] [CrossRef]
  239. Crouvisier-Urion, K.; Lagorce-Tachon, A.; Lauquin, C.; Winckler, P.; Tongdeesoontorn, W.; Domenek, S.; Debeaufort, F.; Karbowiak, T. Impact of the homogenization process on the structure and antioxidant properties of chitosan-lignin composite films. Food Chem. 2017, 236, 120–126. [Google Scholar] [CrossRef]
  240. Yang, W.; Owczarek, J.S.; Fortunati, E.; Kozanecki, M.; Mazzaglia, A.; Balestra, G.M.; Kenny, J.M.; Torre, L.; Puglia, D. Antioxidant and antibacterial lignin nanoparticles in polyvinyl alcohol/chitosan films for active packaging. Ind. Crop. Prod. 2016, 94, 800–811. [Google Scholar] [CrossRef]
  241. Soltani Firouz, M.; Mohi-Alden, K.; Omid, M. A critical review on intelligent and active packaging in the food industry: Research and development. Food Res. Int. 2021, 141. [Google Scholar] [CrossRef]
  242. Mousavi Khaneghah, A.; Hashemi, S.M.B.; Limbo, S. Antimicrobial agents and packaging systems in antimicrobial active food packaging: An overview of approaches and interactions. Food Bioprod. Process. 2018, 111, 1–19. [Google Scholar] [CrossRef]
  243. Appendini, P.; Hotchkiss, J.H. Review of antimicrobial food packaging. Innov. Food Sci. Emerg. Technol. 2002, 3, 113–126. [Google Scholar] [CrossRef]
  244. Han, J.H. Antimicrobial food packaging. In Novel Food Packaging Techniques; Elsevier: Amsterdam, The Netherlands, 2003; pp. 50–70. ISBN 9781855736757. [Google Scholar]
  245. Han, J.H. Antimicrobial packaging systems. In Innovations in Food Packaging; Han, J.H., Ed.; Elsevier: Amsterdam, The Netherlands, 2005; ISBN 0123116325. [Google Scholar]
  246. Coma, V. Bioactive packaging technologies for extended shelf life of meat-based products. Meat Sci. 2008, 78, 90–103. [Google Scholar] [CrossRef]
  247. Lavoine, N.; Givord, C.; Tabary, N.; Desloges, I.; Martel, B.; Bras, J. Elaboration of a new antibacterial bio-nano-material for food-packaging by synergistic action of cyclodextrin and microfibrillated cellulose. Innov. Food Sci. Emerg. Technol. 2014, 26, 330–340. [Google Scholar] [CrossRef]
  248. Cooksey, K. Effectiveness of antimicrobial food packaging materials. Food Addit. Contam. 2005, 22, 980–987. [Google Scholar] [CrossRef]
  249. Pereira, M.C.; Hill, L.E.; Zambiazi, R.C.; Mertens-Talcott, S.; Talcott, S.; Gomes, C.L. Nanoencapsulation of hydrophobic phytochemicals using poly (dl-lactide-co-glycolide) (PLGA) for antioxidant and antimicrobial delivery applications: Guabiroba fruit (Campomanesia xanthocarpa O. Berg) study. LWT Food Sci. Technol. 2015, 63, 100–107. [Google Scholar] [CrossRef] [Green Version]
  250. Qamar, S.A.; Asgher, M.; Bilal, M. Immobilization of alkaline protease from bacillus brevis using ca-alginate entrapment strategy for improved catalytic stability, silver recovery, and dehairing potentialities. Catal. Lett. 2020, 150, 3572–3583. [Google Scholar] [CrossRef]
  251. Ruiz-Navajas, Y.; Viuda-Martos, M.; Sendra, E.; Perez-Alvarez, J.A.; Fernández-López, J. In vitro antibacterial and antioxidant properties of chitosan edible films incorporated with Thymus moroderi or Thymus piperella essential oils. Food Control 2013, 30, 386–392. [Google Scholar] [CrossRef]
  252. Burt, S. Essential oils: Their antibacterial properties and potential applications in foods—A review. Int. J. Food Microbiol. 2004, 94, 223–253. [Google Scholar] [CrossRef] [PubMed]
  253. Wang, S.Y.; Stretch, A.W. Antioxidant capacity in cranberry is influenced by cultivar and storage temperature. J. Agric. Food Chem. 2001, 49, 969–974. [Google Scholar] [CrossRef]
  254. Zheng, W.; Wang, S.Y. Oxygen radical absorbing capacity of phenolics in blueberries, cranberries, chokeberries, and lingonberries. J. Agric. Food Chem. 2003, 51, 502–509. [Google Scholar] [CrossRef] [PubMed]
  255. Srivastava, J.; Vankar, P.S. Principal phenolic phytochemicals and antioxidant property in Eucalyptus bark. Nutr. Food Sci. 2012, 42, 412–421. [Google Scholar] [CrossRef]
  256. Carpena, M.; Nuñez-Estevez, B.; Soria-Lopez, A.; Garcia-Oliveira, P.; Prieto, M.A. Essential Oils and Their Application on Active Packaging Systems: A Review. Resources 2021, 10, 7. [Google Scholar] [CrossRef]
  257. Dhifi, W.; Bellili, S.; Jazi, S.; Bahloul, N.; Mnif, W. Essential oils’ chemical characterization and investigation of some biological activities: A critical review. Medicines 2016, 3. [Google Scholar] [CrossRef] [Green Version]
  258. Raut, J.S.; Karuppayil, S.M. A status review on the medicinal properties of essential oils. Ind. Crop. Prod. 2014, 62, 250–264. [Google Scholar] [CrossRef]
  259. Nieto, G. Biological activities of three essential oils of the lamiaceae family. Medicines 2017, 4. [Google Scholar] [CrossRef] [Green Version]
  260. Priyadarshi, R.; Sauraj, K.B.; Deeba, F.; Kulshreshtha, A.; Negi, Y.S. Chitosan films incorporated with Apricot (Prunus armeniaca) kernel essential oil as active food packaging material. Food Hydrocoll. 2018, 85, 158–166. [Google Scholar] [CrossRef]
  261. Kamari, A.; Halim, A.L.A.; Yusoff, S.N.M.; Ishak, S. Gelatin film incorporated with banana leaf essential oil for food preservation. J. Phys. Conf. Ser. 2018, 1097. [Google Scholar] [CrossRef]
  262. Qin, Y.; Li, W.; Liu, D.; Yuan, M.; Li, L. Development of active packaging film made from poly (lactic acid) incorporated essential oil. Prog. Org. Coat. 2017, 103, 76–82. [Google Scholar] [CrossRef]
  263. Noshirvani, N.; Ghanbarzadeh, B.; Gardrat, C.; Rezaei, M.R.; Hashemi, M.; Le Coz, C.; Coma, V. Cinnamon and ginger essential oils to improve antifungal, physical and mechanical properties of chitosan-carboxymethyl cellulose films. Food Hydrocoll. 2017, 70, 36–45. [Google Scholar] [CrossRef]
  264. Kim, H.; Beak, S.-E.; Song, K.B. Development of a hagfish skin gelatin film containing cinnamon bark essential oil. LWT Food Sci. Technol. 2018, 96, 583–588. [Google Scholar] [CrossRef]
  265. Beak, S.; Kim, H.; Song, K.B. Sea Squirt Shell Protein and Polylactic Acid Laminated Films Containing Cinnamon Bark Essential Oil. J. Food Sci. 2018, 83, 1896–1903. [Google Scholar] [CrossRef] [PubMed]
  266. Xu, T.; Gao, C.; Feng, X.; Yang, Y.; Shen, X.; Tang, X. Structure, physical and antioxidant properties of chitosan-gum arabic edible films incorporated with cinnamon essential oil. Int. J. Biol. Macromol. 2019, 134, 230–236. [Google Scholar] [CrossRef] [PubMed]
  267. Fasihi, H.; Noshirvani, N.; Hashemi, M.; Fazilati, M.; Salavati, H.; Coma, V. Antioxidant and antimicrobial properties of carbohydrate-based films enriched with cinnamon essential oil by Pickering emulsion method. Food Packag. Shelf Life 2019, 19, 147–154. [Google Scholar] [CrossRef]
  268. Wu, J.; Sun, X.; Guo, X.; Ge, S.; Zhang, Q. Physicochemical properties, antimicrobial activity and oil release of fish gelatin films incorporated with cinnamon essential oil. Aquac. Fish. 2017, 2, 185–192. [Google Scholar] [CrossRef]
  269. Aisha, I.; Abdullahi, Y. Development of Whey Protein Concentrate Edible Membrane with Cinnamon Essential Oil. J. Adv. Biol. Biotechnol. 2017, 11, 1–14. [Google Scholar] [CrossRef]
  270. Yang, S.-Y.; Lee, K.-Y.; Beak, S.-E.; Kim, H.; Song, K.B. Antimicrobial activity of gelatin films based on duck feet containing cinnamon leaf oil and their applications in packaging of cherry tomatoes. Food Sci. Biotechnol. 2017, 26, 1429–1435. [Google Scholar] [CrossRef] [PubMed]
  271. Nisar, T.; Wang, Z.-C.; Yang, X.; Tian, Y.; Iqbal, M.; Guo, Y. Characterization of citrus pectin films integrated with clove bud essential oil: Physical, thermal, barrier, antioxidant and antibacterial properties. Int. J. Biol. Macromol. 2018, 106, 670–680. [Google Scholar] [CrossRef]
  272. De Figueiredo Sousa, H.A.; de Oliveira Filho, J.G.; Egea, M.B.; Da Rosa Silva, E.; Macagnan, D.; Pires, M.; Peixoto, J. Active film incorporated with clove essential oil on storage of banana varieties. Nutr. Food Sci. 2019, 49, 911–924. [Google Scholar] [CrossRef]
  273. Sharma, S.; Barkauskaite, S.; Jaiswal, S.; Duffy, B.; Jaiswal, A.K. Development of Essential Oil Incorporated Active Film Based on Biodegradable Blends of Poly (Lactide)/Poly (Butylene Adipate-co-Terephthalate) for Food Packaging Application. J Package Technol. Res. 2020, 4, 235–245. [Google Scholar] [CrossRef]
  274. Azadbakht, E.; Maghsoudlou, Y.; Khomiri, M.; Kashiri, M. Development and structural characterization of chitosan films containing Eucalyptus globulus essential oil: Potential as an antimicrobial carrier for packaging of sliced sausage. Food Packag. Shelf Life 2018, 17, 65–72. [Google Scholar] [CrossRef]
  275. Bonilla, J.; Poloni, T.; Lourenço, R.V.; Sobral, P.J. Antioxidant potential of eugenol and ginger essential oils with gelatin/chitosan films. Food Biosci. 2018, 23, 107–114. [Google Scholar] [CrossRef]
  276. Jamróz, E.; Juszczak, L.; Kucharek, M. Investigation of the physical properties, antioxidant and antimicrobial activity of ternary potato starch-furcellaran-gelatin films incorporated with lavender essential oil. Int. J. Biol. Macromol. 2018, 114, 1094–1101. [Google Scholar] [CrossRef]
  277. Song, X.; Zuo, G.; Chen, F. Effect of essential oil and surfactant on the physical and antimicrobial properties of corn and wheat starch films. Int. J. Biol. Macromol. 2018, 107, 1302–1309. [Google Scholar] [CrossRef]
  278. Akyuz, L.; Kaya, M.; Ilk, S.; Cakmak, Y.S.; Salaberria, A.M.; Labidi, J.; Yılmaz, B.A.; Sargin, I. Effect of different animal fat and plant oil additives on physicochemical, mechanical, antimicrobial and antioxidant properties of chitosan films. Int. J. Biol. Macromol. 2018, 111, 475–484. [Google Scholar] [CrossRef]
  279. Oliveira, S.P.L.F.; Bertan, L.C.; Rensis, C.M.V.B.d.; Bilck, A.P.; Vianna, P.C.B. Whey protein-based films incorporated with oregano essential oil. Polímeros 2017, 27, 158–164. [Google Scholar] [CrossRef] [Green Version]
  280. Sedlaříková, J.; Doležalová, M.; Egner, P.; Pavlačková, J.; Krejčí, J.; Rudolf, O.; Peer, P. Effect of oregano and marjoram essential oils on the physical and antimicrobial properties of chitosan based systems. Int. J. Polym. Sci. 2017, 2017, 2593863. [Google Scholar] [CrossRef]
  281. Ranjbar, M.; Azizi, M.H.; Hashtjin, A.M. Evaluation of physico-mechanical and antimicrobial properties of gelatincarboxymethyl cellulose film containing essential oil of bane (Pistacia atlantica). Nutr. Food Sci. Res. 2017, 4, 11–17. [Google Scholar]
  282. Akhter, R.; Masoodi, F.A.; Wani, T.A.; Rather, S.A. Functional characterization of biopolymer based composite film: Incorporation of natural essential oils and antimicrobial agents. Int. J. Biol. Macromol. 2019, 137, 1245–1255. [Google Scholar] [CrossRef]
  283. Mahcene, Z.; Khelil, A.; Hasni, S.; Akman, P.K.; Bozkurt, F.; Birech, K.; Goudjil, M.B.; Tornuk, F. Development and characterization of sodium alginate based active edible films incorporated with essential oils of some medicinal plants. Int. J. Biol. Macromol. 2020, 145, 124–132. [Google Scholar] [CrossRef]
  284. Hasheminya, S.-M.; Mokarram, R.R.; Ghanbarzadeh, B.; Hamishekar, H.; Kafil, H.S.; Dehghannya, J. Development and characterization of biocomposite films made from kefiran, carboxymethyl cellulose and Satureja Khuzestanica essential oil. Food Chem. 2019, 289, 443–452. [Google Scholar] [CrossRef]
  285. Kouravand, F.; Jooyandeh, H.; Barzegar, H.; Hojjati, M. Characterization of cross-linked whey protein isolate-based films containing Satureja Khuzistanica Jamzad essential oil. J. Food Process. Preserv. 2018, 42. [Google Scholar] [CrossRef]
  286. Abdollahi, M.; Damirchi, S.; Shafafi, M.; Rezaei, M.; Ariaii, P. Carboxymethyl cellulose-agar biocomposite film activated with summer savory essential oil as an antimicrobial agent. Int. J. Biol. Macromol. 2019, 126, 561–568. [Google Scholar] [CrossRef]
  287. Sharma, S.; Barkauskaite, S.; Duffy, B.; Jaiswal, A.K.; Jaiswal, S. Characterization and antimicrobial activity of biodegradable active packaging enriched with clove and thyme essential oil for food packaging application. Foods 2020, 9. [Google Scholar] [CrossRef]
  288. Zeid, A.; Karabagias, I.K.; Nassif, M.; Kontominas, M.G. Preparation and evaluation of antioxidant packaging films made of polylactic acid containing thyme, rosemary, and oregano essential oils. J. Food Process. Preserv. 2019, 83, 1299. [Google Scholar] [CrossRef]
  289. Karacabey, E.; Bayindirli, L.; Artik, N.; Mazza, G. Modeling solid-liquid extraction kinetics of trans -resveratrol and trans -ε-viniferin from grape cane. J. Food Process Eng. 2013, 36, 103–112. [Google Scholar] [CrossRef]
  290. Azwanida, N.N. A review on the extraction methods use in medicinal plants, principle, strength and limitation. Med. Aromat. Plants 2015, 4. [Google Scholar] [CrossRef]
  291. Aleksic Sabo, V.; Knezevic, P. Antimicrobial activity of Eucalyptus camaldulensis Dehn. plant extracts and essential oils: A review. Ind. Crop. Prod. 2019, 132, 413–429. [Google Scholar] [CrossRef] [PubMed]
  292. Havelt, T.; Brettschneider, S.; Do, X.T.; Korte, I.; Kreyenschmidt, J.; Schmitz, M. Sustainable extraction and characterisation of bioactive compounds from horse chestnut seed coats for the development of bio-based additives. Resources 2019, 8, 114. [Google Scholar] [CrossRef] [Green Version]
  293. Havelt, T.; Schmitz, M. Identifizierung und charakterisierung bioaktiver inhaltsstoffe in thymian: 8. tagung arznei- und gewürzpflanzenforschung, Bonn. Jul. Kühn Arch. 2018, 112–114. [Google Scholar] [CrossRef]
  294. Havelt, T.; Frase, J.N.; Pude, R.; Schmitz, M. Characterisation of bioactive ingredients in extracts of fresh and dried coniferous trees for the development of sustainable packaging materials. Processes 2020, 8, 1366. [Google Scholar] [CrossRef]
  295. Kimura, H.; Ogawa, S.; Ishihara, T.; Maruoka, M.; Tokuyama-Nakai, S.; Jisaka, M.; Yokota, K. Antioxidant activities and structural characterization of flavonol O-glycosides from seeds of Japanese horse chestnut (Aesculus turbinata BLUME). Food Chem. 2017, 228, 348–355. [Google Scholar] [CrossRef]
  296. Ay, E.B.; Gul, M.; Acikgoz, M.A.; Yarilgac, T.; Kara, S.M. Assessment of antioxidant activity of giant snowdrop (Galanthus elwesii Hook) extracts with their total phenol and flavonoid contents. Indian J. Pharm. Educ. Res. 2018, 52, 128–132. [Google Scholar] [CrossRef] [Green Version]
  297. Karimi, E.; Mehrabanjoubani, P.; Homayouni-Tabrizi, M.; Abdolzadeh, A.; Soltani, M. Phytochemical evaluation, antioxidant properties and antibacterial activity of Iranian medicinal herb Galanthus transcaucasicus Fomin. J. Food Meas. Charact. 2018, 12, 433–440. [Google Scholar] [CrossRef]
  298. Biswas, M.; Pal, A.; Dey, M.; Dey, A.; Bandyopadhyay, A. Influence of a biobased reagent on properties of industrial resin for printing ink application vis-à-vis comparison with standard commercial resin. Polym. Renew. Resour. 2018, 9, 59–74. [Google Scholar] [CrossRef]
  299. Veiga-Santos, P.; Silva, L.T.; Souza, C.O.d.; Da Silva, J.R.; Albuquerque, E.C.; Druzian, J.I. Coffee-cocoa additives for bio-based antioxidant packaging. Food Packag. Shelf Life 2018, 18, 37–41. [Google Scholar] [CrossRef]
  300. Wang, K.; Lim, P.N.; Tong, S.Y.; Thian, E.S. Development of grapefruit seed extract-loaded poly(ε-caprolactone)/chitosan films for antimicrobial food packaging. Food Packag. Shelf Life 2019, 22, 100396. [Google Scholar] [CrossRef]
  301. Wang, L.; Guo, H.; Wang, J.; Jiang, G.; Du, F.; Liu, X. Effects of Herba Lophatheri extract on the physicochemical properties and biological activities of the chitosan film. Int. J. Biol. Macromol. 2019, 133, 51–57. [Google Scholar] [CrossRef] [PubMed]
  302. Ashrafi, A.; Jokar, M.; Mohammadi Nafchi, A. Preparation and characterization of biocomposite film based on chitosan and kombucha tea as active food packaging. Int. J. Biol. Macromol. 2018, 108, 444–454. [Google Scholar] [CrossRef]
  303. Sogut, E.; Seydim, A.C. The effects of Chitosan and grape seed extract-based edible films on the quality of vacuum packaged chicken breast fillets. Food Packag. Shelf Life 2018, 18, 13–20. [Google Scholar] [CrossRef]
  304. Piñeros-Hernandez, D.; Medina-Jaramillo, C.; López-Córdoba, A.; Goyanes, S. Edible cassava starch films carrying rosemary antioxidant extracts for potential use as active food packaging. Food Hydrocoll. 2017, 63, 488–495. [Google Scholar] [CrossRef]
  305. Mehdizadeh, T.; Langroodi, A.M. Chitosan coatings incorporated with propolis extract and Zataria multiflora Boiss oil for active packaging of chicken breast meat. Int. J. Biol. Macromol. 2019, 141, 401–409. [Google Scholar] [CrossRef]
  306. Langroodi, A.M.; Tajik, H.; Mehdizadeh, T.; Moradi, M.; Moghaddas Kia, E.; Mahmoudian, A. Effects of sumac extract dipping and chitosan coating enriched with Zataria multiflora Boiss oil on the shelf-life of meat in modified atmosphere packaging. LWT Food Sci. Technol. 2018, 98, 372–380. [Google Scholar] [CrossRef]
  307. Rezaeigolestani, M.; Misaghi, A.; Khanjari, A.; Basti, A.A.; Abdulkhani, A.; Fayazfar, S. Antimicrobial evaluation of novel poly-lactic acid based nanocomposites incorporated with bioactive compounds in-vitro and in refrigerated vacuum-packed cooked sausages. Int. J. Food Microbiol. 2017, 260, 1–10. [Google Scholar] [CrossRef] [PubMed]
  308. Rehman, A.; Jafari, S.M.; Aadil, R.M.; Assadpour, E.; Randhawa, M.A.; Mahmood, S. Development of active food packaging via incorporation of biopolymeric nanocarriers containing essential oils. Trends Food Sci. Technol. 2020, 101, 106–121. [Google Scholar] [CrossRef]
  309. Rashed, M.M.; Zhang, C.; Ghaleb, A.D.; Li, J.; Nagi, A.; Majeed, H.; Bakry, A.M.; Haider, J.; Xu, Z.; Tong, Q. Techno-functional properties and sustainable application of nanoparticles-based Lavandula angustifolia essential oil fabricated using unsaturated lipid-carrier and biodegradable wall material. Ind. Crop. Prod. 2019, 136, 66–76. [Google Scholar] [CrossRef]
  310. Wen, P.; Zhu, D.-H.; Feng, K.; Liu, F.-J.; Lou, W.-Y.; Li, N.; Zong, M.-H.; Wu, H. Fabrication of electrospun polylactic acid nanofilm incorporating cinnamon essential oil/β-cyclodextrin inclusion complex for antimicrobial packaging. Food Chem. 2016, 196, 996–1004. [Google Scholar] [CrossRef] [PubMed]
  311. Qiu, C.; Chang, R.; Yang, J.; Ge, S.; Xiong, L.; Zhao, M.; Li, M.; Sun, Q. Preparation and characterization of essential oil-loaded starch nanoparticles formed by short glucan chains. Food Chem. 2016, 221, 1426–1433. [Google Scholar] [CrossRef]
  312. Dos Santos Paglione, I.; Galindo, M.V.; Medeiros, J.A.S.d.; Yamashita, F.; Alvim, I.D.; Ferreira Grosso, C.R.; Sakanaka, L.S.; Shirai, M.A. Comparative study of the properties of soy protein concentrate films containing free and encapsulated oregano essential oil. Food Packag. Shelf Life 2019, 22, 100419. [Google Scholar] [CrossRef]
  313. Mohsenabadi, N.; Rajaei, A.; Tabatabaei, M.; Mohsenifar, A. Physical and antimicrobial properties of starch-carboxy methyl cellulose film containing rosemary essential oils encapsulated in chitosan nanogel. Int. J. Biol. Macromol. 2018, 112, 148–155. [Google Scholar] [CrossRef]
  314. Khoobdel, M.; Ahsaei, S.M.; Farzaneh, M. Insecticidal activity of polycaprolactone nanocapsules loaded with Rosmarinus officinalis essential oil in Tribolium castaneum (Herbst). Entomol. Res. 2017, 47, 175–184. [Google Scholar] [CrossRef]
  315. Werdin González, J.O.; Jesser, E.N.; Yeguerman, C.A.; Ferrero, A.A.; Fernández Band, B. Polymer nanoparticles containing essential oils: New options for mosquito control. Environ. Sci. Pollut. Res. 2017, 24, 17006–17015. [Google Scholar] [CrossRef]
  316. Lin, L.; Mao, X.; Sun, Y.; Rajivgandhi, G.; Cui, H. Antibacterial properties of nanofibers containing chrysanthemum essential oil and their application as beef packaging. Int. J. Food Microbiol. 2019, 292, 21–30. [Google Scholar] [CrossRef]
  317. Maestrello, C.; Tonon, L.; Madrona, G.; Scapim, M.; Bergamasco, R. Production and Characterization of Biodegradable Films Incorporated with Clove Essential Oil/β-cyclodextrin Microcapsules. Chem. Eng. Trans. 2017, 57, 1274–1283. [Google Scholar] [CrossRef]
  318. Farshi, P.; Tabibiazar, M.; Ghorbani, M.; Mohammadifar, M.; Amirkhiz, M.B.; Hamishehkar, H. Whey protein isolate-guar gum stabilized cumin seed oil nanoemulsion. Food Biosci. 2019, 28, 49–56. [Google Scholar] [CrossRef]
  319. Karimirad, R.; Behnamian, M.; Dezhsetan, S. Development and characterization of nano biopolymer containing cumin oil as a new approach to enhance antioxidant properties of button mushroom. Int. J. Biol. Macromol. 2018, 113, 662–668. [Google Scholar] [CrossRef]
  320. Wu, Z.; Zhou, W.; Pang, C.; Deng, W.; Xu, C.; Wang, X. Multifunctional chitosan-based coating with liposomes containing laurel essential oils and nanosilver for pork preservation. Food Chem. 2019, 295, 16–25. [Google Scholar] [CrossRef] [PubMed]
  321. Lin, L.; Gu, Y.; Cui, H. Moringa oil/chitosan nanoparticles embedded gelatin nanofibers for food packaging against Listeria monocytogenes and Staphylococcus aureus on cheese. Food Packag. Shelf Life 2019, 19, 86–93. [Google Scholar] [CrossRef]
  322. Fraj, A.; Jaâfar, F.; Marti, M.; Coderch, L.; Ladhari, N. A comparative study of oregano (Origanum vulgare L.) essential oil-based polycaprolactone nanocapsules/ microspheres: Preparation, physicochemical characterization, and storage stability. Ind. Crop. Prod. 2019, 140, 111669. [Google Scholar] [CrossRef]
  323. Merino, N.; Berdejo, D.; Bento, R.; Salman, H.; Lanz, M.; Maggi, F.; Sánchez-Gómez, S.; García-Gonzalo, D.; Pagán, R. Antimicrobial efficacy of Thymbra capitata (L.) Cav. essential oil loaded in self-assembled zein nanoparticles in combination with heat. Ind. Crop. Prod. 2019, 133, 98–104. [Google Scholar] [CrossRef]
  324. Vafania, B.; Fathi, M.; Soleimanian-Zad, S. Nanoencapsulation of thyme essential oil in chitosan-gelatin nanofibers by nozzle-less electrospinning and their application to reduce nitrite in sausages. Food Bioprod. Process. 2019, 116, 240–248. [Google Scholar] [CrossRef]
  325. Lin, L.; Zhu, Y.; Cui, H. Electrospun thyme essential oil/gelatin nanofibers for active packaging against Campylobacter jejuni in chicken. LWT Food Sci. Technol. 2018, 97, 711–718. [Google Scholar] [CrossRef]
  326. Lee, M.H.; Seo, H.-S.; Park, H.J. Thyme oil encapsulated in halloysite nanotubes for antimicrobial packaging system. J. Food Sci. 2017, 82, 922–932. [Google Scholar] [CrossRef] [PubMed]
  327. Jang, S.-H.; Jang, S.-R.; Lee, G.-M.; Ryu, J.-H.; Park, S.-I.; Park, N.-H. Halloysite Nanocapsules Containing Thyme Essential Oil: Preparation, Characterization, and Application in Packaging Materials. J. Food Sci. 2017, 82, 2113–2120. [Google Scholar] [CrossRef] [PubMed]
  328. Ardekani, N.T.; Khorram, M.; Zomorodian, K.; Yazdanpanah, S.; Veisi, H.; Veisi, H. Evaluation of electrospun poly (vinyl alcohol)-based nanofiber mats incorporated with Zataria multiflora essential oil as potential wound dressing. Int. J. Biol. Macromol. 2019, 125, 743–750. [Google Scholar] [CrossRef] [PubMed]
  329. Bos, H.; Meesters, K.; Conijn, S.; Corré, W.; Patel, M. Sustainability Aspects of Biobased Applications. Available online: http://edepot.wur.nl/170079 (accessed on 25 May 2019).
  330. Schumann, B.; Schmid, M. Packaging concepts for fresh and processed meat—Recent progresses. Innov. Food Sci. Emerg. Technol. 2018, 47, 88–100. [Google Scholar] [CrossRef]
  331. Golembiewski, B.; Sick, N.; Broering, S. The emerging research landscape on bioeconomy: What has been done so far and what is essential from a technology and innovation management perspective? Innov. Food Sci. Emerg. Technol. 2015, 29, 308–317. [Google Scholar] [CrossRef]
  332. Carraresi, L.; Berg, S.; Broering, S. Emerging value chains within the bioeconomy: Structural changes in the case of phosphate recovery. J. Clean. Prod. 2018, 183, 87–101. [Google Scholar] [CrossRef]
  333. Keegan, D.; Kretschmer, B.; Elbersen, B.; Panoutsou, C. Cascading use: A systematic approach to biomass beyond the energy sector. Biofuels Bioprod. Biorefining 2013, 7, 193–206. [Google Scholar] [CrossRef]
  334. Theinsathid, P.; Chandrachai, A.; Suwannathep, S.; Keeratipibul, S. Lead users and early adoptors of bioplastics: A market-led approach to innovative food packaging films. J. Biobased Mater. Bioenergy 2011, 5, 17–29. [Google Scholar] [CrossRef]
  335. Wensing, J.; Caputo, V.; Carraresi, L.; Bröring, S. The effects of green nudges on consumer valuation of bio-based plastic packaging. Ecol. Econ. 2020, 178, 106783. [Google Scholar] [CrossRef]
  336. Rossi, A.M.; Hinrichs, C.C. Hope and skepticism: Farmer and local community views on the socio-economic benefits of agricultural bioenergy. Biomass Bioenergy 2011, 35, 1418–1428. [Google Scholar] [CrossRef]
  337. Wensing, J.; Carraresi, L.; Broering, S. Do pro-environmental values, beliefs and norms drive farmers’ interest in novel practices fostering the Bioeconomy? J. Environ. Manag. 2019, 232, 858–867. [Google Scholar] [CrossRef]
  338. Van Eck, P.S.; Jager, W.; Leeflang, P.S.H. Opinion leaders’ role in innovation diffusion: A simulation study. J. Prod. Innov. Manag. 2011, 28, 187–203. [Google Scholar] [CrossRef]
  339. Xiong, H.; Payne, D.; Kinsella, S. Peer effects in the diffusion of innovations: Theory and simulation. J. Behav. Exp. Econ. 2016, 63, 1–13. [Google Scholar] [CrossRef]
  340. Berg, S.; Cloutier, L.M.; Broering, S. Collective stakeholder representations and perceptions of drivers of novel biomass-based value chains. J. Clean. Prod. 2018, 200, 231–241. [Google Scholar] [CrossRef]
  341. Sleenhoff, S.; Cuppen, E.; Osseweijer, P. Unravelling emotional viewpoints on a bio-based economy using Q methodology. Public Underst. Sci. 2015, 24, 858–877. [Google Scholar] [CrossRef] [PubMed]
  342. Sijtsema, S.J.; Onwezen, M.C.; Reinders, M.J.; Dagevos, H.; Partanen, A.; Meeusen, M. Consumer perception of bio-based products-An exploratory study in 5 European countries. Wagening J. Life Sci. 2016, 77, 61–69. [Google Scholar] [CrossRef]
  343. Stern, T.; Ploll, U.; Spies, R.; Schwarzbauer, P.; Hesser, F.; Ranacher, L. Understanding perceptions of the bioeconomy in Austria—An explorative case study. Sustainability 2018, 10. [Google Scholar] [CrossRef] [Green Version]
  344. Petljak, K.; Naletina, D.; Bilogrević, K. Considering ecologically sustainable packaging during decision-making while buying food products. Ekon. Poljopr. 2019, 66, 107–126. [Google Scholar] [CrossRef] [Green Version]
  345. Koutsimanis, G.; Harte, J.; Almenar, E. Freshness maintenance of cherries ready for consumption using convenient, microperforated, bio-based packaging. J. Sci. Food Agric. 2015, 95, 972–982. [Google Scholar] [CrossRef]
  346. Herbes, C.; Beuthner, C.; Ramme, I. Consumer attitudes towards biobased packaging—A cross-cultural comparative study. J. Clean. Prod. 2018, 194, 203–218. [Google Scholar] [CrossRef]
  347. Scherer, C.; Emberger-Klein, A.; Menrad, K. Biogenic product alternatives for children: Consumer preferences for a set of sand toys made of bio-based plastic. Sustain. Prod. Consum. 2017, 10, 1–14. [Google Scholar] [CrossRef]
  348. Scherer, C.; Emberger-Klein, A.; Menrad, K. Consumer preferences for outdoor sporting equipment made of biobased plastics: Results of a choice-based-conjoint experiment in Germany. J. Clean. Prod. 2018, 203, 1085–1094. [Google Scholar] [CrossRef]
  349. Klein, F.; Emberger-Klein, A.; Menrad, K.; Moehring, W.; Blesin, J.-M. Influencing factors for the purchase intention of consumers choosing bioplastic products in Germany. Sustain. Prod. Consum. 2019, 19, 33–43. [Google Scholar] [CrossRef]
  350. Rumm, S. Verbrauchereinschätzungen zu Biokunststoffen: Eine Analyse vor dem Hintergrund des Heuristic-Systematic Model. Available online: https://mediatum.ub.tum.de/1306582 (accessed on 18 October 2018).
  351. Müller, G.; Hanecker, E.; Blasius, K.; Seidemann, C.; Tempel, L.; Sadocco, P.; Pozo, B.F.; Boulougouris, G.; Lozo, B.; Jamnicki, S.; et al. End-of-life Solutions for Fibre and Bio-based Packaging Materials in Europe. Packag. Technol. Sci. 2014, 27, 1–15. [Google Scholar] [CrossRef]
  352. Yue, C.; Hall, C.R.; Behe, B.K.; Campbell, B.L.; Dennis, J.H.; Lopez, R.G. Are consumers willing to pay more for biodegradable containers than for plastic ones? Evidence from hypothetical conjoint analysis and nonhypothetical experimental auctions. In Proceedings of the 2010 Annual Meeting, Denver, CO, USA, 25–27 July 2010; pp. 757–772. [Google Scholar]
  353. Klaiman, K.; Ortega, D.L.; Garnache, C. Consumer preferences and demand for packaging material and recyclability. Resour. Conserv. Recycl. 2016, 115, 1–8. [Google Scholar] [CrossRef]
Figure 1. Classification of plastics used for packaging applications. Adapted from [95]. Poly(amide) (PA), poly(butylene adipate terephthalate) (PBAT), poly(butylene succinate) (PBS), poly(butylene succinate) (PBS), poly(caprolactone) (PCL), poly(ethylene) (PE), poly(ethylene terephthalate) (PET), poly(hydroxyalkanoate) (PHA), poly(lactic acid) (PLA), poly(propylene) (PP), poly(trimethylene terephthalate) (PTT).
Figure 1. Classification of plastics used for packaging applications. Adapted from [95]. Poly(amide) (PA), poly(butylene adipate terephthalate) (PBAT), poly(butylene succinate) (PBS), poly(butylene succinate) (PBS), poly(caprolactone) (PCL), poly(ethylene) (PE), poly(ethylene terephthalate) (PET), poly(hydroxyalkanoate) (PHA), poly(lactic acid) (PLA), poly(propylene) (PP), poly(trimethylene terephthalate) (PTT).
Applsci 11 05307 g001
Figure 2. PLA structure.
Figure 2. PLA structure.
Applsci 11 05307 g002
Figure 3. Overview of several poly(hydroxyl alkanoate)s (PHA) used for food packaging applications: (a) general structure of PHAs with residues R1/R2 as alkyl chains ranging from 1-13 carbons, (b) poly(hydroxyl butyrate) (PHB), (c) Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), and (d) poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx).
Figure 3. Overview of several poly(hydroxyl alkanoate)s (PHA) used for food packaging applications: (a) general structure of PHAs with residues R1/R2 as alkyl chains ranging from 1-13 carbons, (b) poly(hydroxyl butyrate) (PHB), (c) Poly(3-hydroxybutyrate-co-3-hydroxyvalerate) (PHBV), and (d) poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx).
Applsci 11 05307 g003
Figure 4. Lignin monolignol structures: p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol forming the specific residues p-hydroxylphenyl (H), guaiacyl (G), and syringyl (S) [140].
Figure 4. Lignin monolignol structures: p-coumaryl alcohol, coniferyl alcohol, and sinapyl alcohol forming the specific residues p-hydroxylphenyl (H), guaiacyl (G), and syringyl (S) [140].
Applsci 11 05307 g004
Figure 5. Overview of several polysaccharides used for food packaging (edible) films: (a) starch, (b) cellulose, (c) alginate, and (d) chitosan [188].
Figure 5. Overview of several polysaccharides used for food packaging (edible) films: (a) starch, (b) cellulose, (c) alginate, and (d) chitosan [188].
Applsci 11 05307 g005
Table 1. Barrier and mechanical properties of selected synthetic polymers and biomass-derived chemically synthesized materials.
Table 1. Barrier and mechanical properties of selected synthetic polymers and biomass-derived chemically synthesized materials.
Film CompositionPermeabilityMechanical PropertiesReferences
O2CO2H2O VaporTensile StrengthElongation at Break
[cm3·mm/(m2·d·atm)][g·mm/(m2·d·atm)][MPa][%]
Synthetic Polymers
HDPE44–91
(20 °C, 65% RH)
100
(23 °C, 0% RH)
0.15
(38 °C, 90% RH)
32150[110,111]
LPDE98–183
(23 °C, 50% RH)
NR0.44
(38 °C, 90% RH)
10400[110,111]
PE50–200
(23 °C, 50% or 0% RH)
100–1000
(23-25 °C, unknown RH)
0.5–2
(23 °C, 85% RH)
18350[111,112,113]
PET1–5
(23 °C, 50% or 0% RH)
3–7
(23 °C, 75% RH)
0.5–2
(23 °C, 85% RH)
55300[111,112,113]
PP50–100
(23 °C, 50% or 0% RH)
200–900
(unknown conditions)
0.2–0.4
(23 °C, 85% RH)
2680[111,112,113]
Biomass-derived chemically synthesized materials
PLA3.5–15
(23 °C, 50% or 0% RH)
32.9–72
(23 °C, 0% RH)
1.6–3.6
(38 °C, 85% RH)
50.453[110,114,115]
NR: Data not reported.
Table 2. Barrier and mechanical properties of bio-based polymers produced by microorganisms.
Table 2. Barrier and mechanical properties of bio-based polymers produced by microorganisms.
Film CompositionPermeabilityMechanical PropertiesReferences
O2CO2H2O VaporTensile StrengthElongation at Break
[cm3·mm/(m2·d·atm)][g·mm/(m2·d atm)][MPa][%]
PHB
(P3HB)
2–11.4
(23 °C, 0% RH)
3–28.9
(23 °C, 0% RH)
1–5
(unknown conditions)
35–403–8[110,137]
PHBV
(P(3HB-co-3HV))
4.9–16.7
(25 °C, 0% RH)
146
(25 °C, 0% RH)
1.5
(38 °C, 90% RH)
38NR[110,137]
PHBHHx8.3
(23 °C, 0% RH)
54
(23 °C, 0% RH)
1.42
(23 °C, 0% RH)
20850[132,137]
NR—Data not reported.
Table 3. Barrier and mechanical properties of protein-based polymers.
Table 3. Barrier and mechanical properties of protein-based polymers.
Film CompositionPermeabilityMechanical PropertiesReferences
O2CO2H2O VaporTensile StrengthElongation at Break
[g·mm/(m2·d·atm)] *[MPa][%]
Soy protein isolate film with glycerol (casting method)NRNR1710
(23 °C, 50% RH)
6.97113.94[180]
Whey protein isolate film with glycerol92.448 g/(m2·d)
(25 °C, 90% RH)
NR1680
(25 °C, 75% RH)
313[181]
Pumpkin oil cake protein isolate film with glycerol16.06 cm3/(m2·d·atm)
(23 °C; unknown RH)
21.15 cm3/(m2·d·atm) *
(23 °C; unknown RH)
NR0.86–6.5622.2–196.61[165]
Fish gelatin film with glycerolNRNR1040
(24 °C, 50% RH)
9.0844.93[182]
Beef skin gelatin films with corn oil (extrusion)0.8–4.7 × 10−4 cm3·mm/(m2·d·atm)
(23 °C, 50% RH)
NR4.05–8.61 × 106
(23 °C, 50% RH)
1.43–5.371.68–2.60[183]
* In some data, the units were normalized. NR—Data not reported.
Table 4. Barrier and mechanical properties of polysaccharide-based polymers.
Table 4. Barrier and mechanical properties of polysaccharide-based polymers.
Film CompositionPermeabilityMechanical PropertiesReferences
O2H2O VaporTensile StrengthElongation at Break
[g·mm/(m2·d·atm)] *[MPa][%]
High amylose cornstarch films without plasticizer
(amylose:amylopectin ratio 80:20)
NR1260
(20 °C; 52.9% RH)
34.321.41[189]
Low amylose cornstarch films without plasticizer
(amylose:amylopectin ratio 25:75)
NR1430
(20 °C; 52.9% RH)
44.382.40[189]
Thermoplastic (cassava) starch (extrusion)0.182
(Ambient temp., 0% RH)
36.8
(25 °C, 50% RH)
5.878[190,191]
Methylcellulose mixtures in ethanolNR446–945
(25 °C; 52% RH)
25–3329–14[192]
Hydroxypropyl methylcellulose without plasticizerNR974,000
(23 °C, 50% RH)
61.0429.51[193]
Carboxymethyl cellulose film with glycerol (casting method)NR683
(25 °C, 52.8% RH)
6.10201.73[194]
CMC-film with 50 wt% ethanol organosolv lignin with glycerol (casting method)NR2570
(20 °C, 0% RH)
205.92[149]
Agar/10% lignin composite film with glycerolNR13,400
(25 °C, 50% RH)
51.822.1[195]
(Sodium) alginate film with glycerol (casting method)NR13,600
(25 °C, 50% RH)
41.18.5[196]
Chitosan film with glycerol0.188 × 10−2
(25 °C, <50% RH)
210–3020
(25 °C, 0–100% RH)
8.938.5[197]
* In some data, the units were normalized. NR: Data not reported.
Table 5. Advantages and disadvantages of different bio-based polymers discussed in Section 3.
Table 5. Advantages and disadvantages of different bio-based polymers discussed in Section 3.
MaterialAdvantageDisadvantageReference
PLA-basedRenewable, biodegradable, biocompatible
Usable for mono- and multilayer applications
Desirable mechanical properties (stiffness, tensile strength)
Good gas permeability
Transparent
Expensive (synthesis)
Limited to rigid packaging
[39,83,104,107,108,109]
PHA-basedWater-resistant surfaces by coating
Functionalize grease resistance and sealability
Good thermomechanical properties
Desirable gas permeability and WVP
Not transparent[106,110,129,130]
Lignocellulose biomass and/or lignin-basedAbundance in nature
Antioxidant activity with long-term stability
Improves mechanical, thermal, and barrier properties
Reduction in WVP
Deficient quality (technical lignins)
Copolymerization requires functionalization
[147,149,158,159,160]
Protein-basedBiodegradable
Abundance in nature
Good film-forming properties
Desirable barrier properties
Transparent
Low cost/cost-effective
Low mechanical strength
Lack of heat stability
[83,90,165,166,167,168]
Polysaccharide-basedEnvironmentally compatible, biodegradable
Abundance in nature
Effective oxygen barriers (intermediate to low humidity)
Good mechanical properties
Transparent
Potential for edible packaging
Poor water vapor barrier
Sensitive to moisture
[2,82,106,184,187,198]
Table 6. Spotlight literature for plant essential oils used in the context of active food packaging.
Table 6. Spotlight literature for plant essential oils used in the context of active food packaging.
BiomassPackaging MatrixResultsReference
Apricot kernel EOChitosanPrepared films showed better water resistance and improved antioxidant, antimicrobial, and mechanical properties; fungal growth on packaged bread is inhibited[260]
Banana leaf EOGelatinImproved antimicrobial properties (against E. coli and S. aureus) of gelatin films enriched with banana leaf EO; improvements on mechanical properties observed[261]
Bergamot, lemongrass, rosemary, and clove EOsPLAEnhanced mechanical and antimicrobial properties against E. coli[262]
Cinnamon and ginger EOsCMC and ChitosanDecreased water vapor permeability (particularly for cinnamon EO), antifungal activity against A. niger with a higher efficacy of cinnamon EO[263]
Cinnamon bark EOGelatinAntioxidant and antimicrobial effects observed (against S. typhimurium and L. monocytogenes), water resistance is increased[264]
Cinnamon bark EOPLA and Sea squirt (Halocynthia roretzi) shell proteinEnhanced antioxidant activity and antimicrobial effects against L. monocytogenes, S. aureus, E. coli and S. typhimurium; change of mechanical properties, decrease of water solubility, and water permeability[265]
Cinnamon EOChitosan and Gum arabicBetter water barrier properties with a decrease in mechanical properties; high antioxidant effect when applying appropriate ratios of chitosan, gum Arabic, and cinnamon EO[266]
Cinnamon EOCMC and PVAEnhancement of antioxidant and photostabilizing properties; highly effective against Penicillium digitatum; shelf life of packaged bread was increased[267]
Cinnamon EOGelatinWater vapor permeability and light-absorbing properties of enriched films increase while water content and elongation at break decrease; antifungal and antimicrobial activity against E. coli, S. aureus, A. niger, Rhizopus oryzae, and Paecilomyces varioti observed[268]
Cinnamon EOWhey proteinEffect of enriched films against S. aureus, no effect against E. coli observed[269]
Cinnamon leaf oilGelatinAntimicrobial effect against foodborne pathogens reported (E. coli, S. typhimurium, S. aureus, L. monocytogenes) [270]
Clove bud EOPectinAntioxidant and antimicrobial effects observed (against S. aureus, E. coli, and L. monocytogenes); improved mechanical properties (flexibility, resistance to breakage, water barrier properties, and heat stability)[271]
Clove EOStarchAntifungal activity against Colletotrichum gloeosporioides and Colletotrichum musae, but not against Saccharomyces bourladii; enhanced shelf life of packaged bananas[272]
Eucalyptus and Cinnamon EOsPLA and PBATAntimicrobial activities against E. coli and S. aureus observed for both EOs with cinnamon EO showing a higher antimicrobial effect, increased biofilm inhibition and decreased UV-light transmission[273]
Eucalyptus globulus EOChitosanAntibacterial effects against S. entertidis, E. coli, B. cereus, and S. aureus observed (especially in liquid phase); lower antibacterial effect in vapor phase[274]
Ginger EO and EugenolGelatin and ChitosanAntioxidant effect observed for both Eugenol and ginger EO (depending on film formulation); comparable water vapor permeability with increased elasticity[275]
Lavender EOStarch, Furcellaran, and GelatinEnhanced antioxidant and antimicrobial effects against E. coli and S. aureus; change of mechanical properties with addition of Lavender EO (decrease of tensile strength, water absorption, etc.)[276]
Lemon EOStarchOptical and mechanical properties examined; antimicrobial effects against S. aureus and E. coli[277]
Olive oil, corn oil, sunflower oilChitosanParticularly olive oil enriched films showed better mechanical properties and a high antibacterial activity[278]
Oregano EOWhey proteinHigher amounts of EO resulted in higher water vapor permeability and film flexibility; antimicrobial activity against Penicillium commune[279]
Origanum vulgare, O. majorana EOsChitosanAntimicrobial effect against S. aureus and B. cereus observed with both EOs, significantly enhanced effects for O. vulgare EO[280]
Pistacia atlantica EOCMC and GelatinAntimicrobial effect against E. coli, S. aureus, Clostridium sporogenes, and particularly Salmonella enterica; reduction of e.g., water vapor permeability, film thickness, and tensile strength[281]
Rosemary and mint EOsChitosan, Pectin, and StarchRosemary and mint EOs improved water barrier properties and inhibited Bacillus subtilis, E. coli, and L. monocytogenes; both EOs resulted in enhanced antioxidant effects[282]
Rosmarinus officinalis, Artemisia herba-alba, Ocimum basilicum and Mentha pulegium EOsAlginateStrong antibacterial activity against S. aureus, E. coli, Salmonella enterica, Enterococcus faecium, Klebsiella pneumoniae, and Enterococcus faecalis; physical properties analyzed, antioxidant effect observed[283]
Satureja Khuzestanica EOKefiran and CMCAntimicrobial effects against S. aureus and E. coli, significant antioxidant properties, change in physical properties (e.g., decrease in water vapor permeability)[284]
Satureja Khuzistanica Jamzad EOWhey proteinAntimicrobial effect particularly against S. aureus with Pseudomonas aeruginosa showing the highest resistance of analyzed bacteria; increased elongation at break and water vapor permeability[285]
Summer savory EOCMC and AgarAntimicrobial effects particularly against S. aureus, B. cereus, and L. monocytogenes with lower effects against E. coli; alteration of physical properties (increased water vapor permeability, improved mechanical flexibility)[286]
Thyme and Clove EOsPLA and PBATPositive properties observed for both EOs, but particularly for clove EO films, including UV-blocking and highly antimicrobial effects (inhibition of E. coli, complete killing of S. aureus)[287]
Thyme, rosemary, and oregano EOsPLASignificant antioxidant effect on packaged minced fish with moderate alteration of mechanical properties[288]
Abbreviations: Essential Oil—EO; Escherichia coliE. coli; Staphylococcus aureusS. aureus; Aspergillus nigerA. niger; Salmonella typhimuriumS. typhimurium; Listeria monocytogenesL. monocytogenes; Salmonella entertidisS. entertidis; Bacillus cereusB. cereus.
Table 7. Spotlight literature for plant extracts used in the context of active food packaging.
Table 7. Spotlight literature for plant extracts used in the context of active food packaging.
Extracted biomassPackaging matrixResultsReference
Coffee beans and de-fatted cocoa beansStarchSynergistic antioxidant effect, decreased water vapor permeability, increased shelf life of palm oil[299]
Grapefruit seed extractPCL and ChitosanFilms showed better mechanical properties and inhibited bacterial growth of E. coli and P. aeruginosa for up to 6 days; successful tests with packaged salmon and bread[300]
Herba Lophatheri extractChitosanMoisture and oil resistance are enhanced, both antioxidant and antimicrobial activities observed (E. coli, S. aureus)[301]
Kombucha tea extractChitosanDecreased water vapor permeability and improved antioxidant, photoabsorbing, and antimicrobial effects (against E. coli and S. aureus); 3 days extended shelf life for packaged minced beef[302]
Grape seed extractChitosanEnhanced antioxidant and antimicrobial activity (total mesophilic aerobic bacteria, coliforms, E. coli, L. monocytogenes, S. aureus, and P. aeruginosa; shelf-life extension of refrigerated, vacuum-packed chicken breast fillets) [303]
Rosemary extractStarchSignificant antioxidant effect, increased UV-stability[304]
Propolis extract and Zataria multiflora EOChitosanAntimicrobial effects measured for mesophilic total viable plate counts, lactic acid bacteria, psychotropic bacteria, and Pseudomonas; synergistic effects observed; lower microbial load on packaged chicken[305]
Sumac extract and Zataria multiflora EOChitosanAntioxidant effects and prolonged shelf life on packaged meat observed; antimicrobial activity against different bacteria (e.g., Pseudomonas spp.)[306]
Propolis extract and Zataria multiflora EOPLAIncreased shelf life of packaged sausages, antimicrobial effects against common food pathogens (S. aureus, E. coli, Vibrio parahaemolyticus, L. monocytogenes)[307]
Abbreviations: Escherichia coliE. coli; Pseudomonas aeruginosaP. aeruginosa; Staphylococcus aureusS. aureus; Listeria monocytogenesL. monocytogenes.
Table 8. Spotlight literature for encapsulated plant oils used in the context of active food packaging.
Table 8. Spotlight literature for encapsulated plant oils used in the context of active food packaging.
BiomassPackaging MatrixEncapsulation DetailsResultsReference
Chrysanthemum EO-Chitosan nanofibersAntioxidant and antimicrobial effect against L. monocytogenes observed e.g., on packaged beef, prolongation of shelf life possible[316]
Cinnamon EOPLANanofibersBetter antimicrobial effect against S. aureus and E. coli observed for encapsulated EO; encapsulation process is more suitable formulation method to maintain EO properties; shelf life of packaged pork was prolonged[310]
Clove EOAlginateInclusion complexSuccessful incorporation of clove EO complexes; resulting in less transparent and flexible films, decreased elasticity, increased water vapor permeability[317]
Cumin seed oil-Nanoemulsion (Whey protein, Guar gum)Antimicrobial effect of encapsulated oil against S. aureus, E. coli, and A. flavus[318]
Cuminum cyminum EO-Chitosan nanoparticlesSignificant antioxidant effect in packaged white button mushrooms observed, resulting in presumed shelf-life prolongation[319]
Laurel EO and silver nanoparticlesPELiposomes in ChitosanAntioxidant properties observed during 7 days of storage with only about 30% of EO released from liposomes; antimicrobial effect against S. aureus and E. coli results in 6 days prolonged shelf life of packaged pork[320]
Lavandula angustofolia EO-Nanoemulsion (Whey protein)Encapsulation enhanced thermal stability of EO; antibacterial effect is observed[309]
Menthone, Oregano, Cinnamon, Lavender and Citral EOs-Starch nanoparticlesEnhanced stability of antioxidants against thermal influence after encapsulation; antimicrobial effects against E. coli and S. aureus are prolonged[311]
Moringa oilGelatin nanofibersChitosan nanoparticlesHigh antimicrobial activity of encapsulated Moringa oil against L. monocytogenes and S. aureus for 10 days without affecting the sensory properties of packaged cheese[321]
Oregano EO-PCL nanocapsulesHigh retention of encapsulated Rosemary EO (determined via carvacrol content) observed, suitability for long-term delivery of carvacrol can be assumed[322]
Oregano EOSoy proteinMicroencapsulation by ionic gelationStrong antioxidant and antimicrobial properties against E. coli and S. aureus; enhanced effects and mechanical properties with microencapsulated EO in contrast to free EOs[312]
Rosemary EOStarch and CMCChitosan nanogelFilms with encapsulated EO show higher water vapor permeability, higher transparency, and tensile strength; immediate (free EO) and gradual (encapsulated EO) antimicrobial effects against S. aureus were observed[313]
Thymbra capitata EO-Zein nanoparticlesBoth free and encapsulated EO are effective against E. coli and L. monocytogenes; presumably due to controlled release, encapsulated EO showed lower antimicrobial efficacy compared to free EO[323]
Thyme EO-Nanofibers (Chitosan, Gelatin)Both free and encapsulated thyme EO has antioxidant and antimicrobial effects against Clostridium perfringens; tests show that such nanofibers could be used to substitute nitrite in meat products[324]
Thyme EOGelatinNanofibersAntimicrobial effect against Campylobacter jejuni in packaged chicken observed[325]
Thyme EOInk (for paper packaging)Halloysite nanotubesStrong antibacterial activity against E. coli, mesophilic aerobic bacteria, molds, and yeasts for up to 10 days after encapsulation in Halloysite nanotubes[326,327]
Zataria multiflora EOPVANanofibers (Chitosan, PVA, Gelatin)Encapsulated Zataria multiflora EO completely inhibited growth of S. aureus, P. aeruginosa, and Candida albicans for 24 h; tested material is developed for use as wound dressing[328]
Abbreviations: Essential Oil—EO; Listeria monocytogenesL. monocytogenes; Staphylococcus aureusS. aureus; Escherichia coliE. coli; Aspergillus flavusA. flavus; Pseudomonas aeruginosaP. aeruginosa.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Korte, I.; Kreyenschmidt, J.; Wensing, J.; Bröring, S.; Frase, J.N.; Pude, R.; Konow, C.; Havelt, T.; Rumpf, J.; Schmitz, M.; et al. Can Sustainable Packaging Help to Reduce Food Waste? A Status Quo Focusing Plant-Derived Polymers and Additives. Appl. Sci. 2021, 11, 5307. https://0-doi-org.brum.beds.ac.uk/10.3390/app11115307

AMA Style

Korte I, Kreyenschmidt J, Wensing J, Bröring S, Frase JN, Pude R, Konow C, Havelt T, Rumpf J, Schmitz M, et al. Can Sustainable Packaging Help to Reduce Food Waste? A Status Quo Focusing Plant-Derived Polymers and Additives. Applied Sciences. 2021; 11(11):5307. https://0-doi-org.brum.beds.ac.uk/10.3390/app11115307

Chicago/Turabian Style

Korte, Imke, Judith Kreyenschmidt, Joana Wensing, Stefanie Bröring, Jan Niklas Frase, Ralf Pude, Christopher Konow, Thomas Havelt, Jessica Rumpf, Michaela Schmitz, and et al. 2021. "Can Sustainable Packaging Help to Reduce Food Waste? A Status Quo Focusing Plant-Derived Polymers and Additives" Applied Sciences 11, no. 11: 5307. https://0-doi-org.brum.beds.ac.uk/10.3390/app11115307

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