Next Article in Journal
Hybrid PET Track-Etched Membranes Grafted by Well-Defined Poly(2-(dimethylamino)ethyl methacrylate) Brushes and Loaded with Silver Nanoparticles for the Removal of As(III)
Previous Article in Journal
A Novel ZnO Nanoparticles Enhanced Surfactant Based Viscoelastic Fluid Systems for Fracturing under High Temperature and High Shear Rate Conditions: Synthesis, Rheometric Analysis, and Fluid Model Derivation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Emerging Food Packaging Applications of Cellulose Nanocomposites: A Review

1
School of Food and Pharmacy, Zhejiang Ocean University, Zhoushan 316022, China
2
School of Engineering Sciences in Chemistry, Biotechnology and Health, KTH Royal Institute of Technology, 10044 Stockholm, Sweden
3
Fujian Provincial Key Laboratory of Plasma and Magnetic Resonance, Institute of Electromagnetics and Acoustics, Xiamen University, Xiamen 361005, China
4
School of Chemical and Biomolecular Engineering, The University of Sydney, Sydney, NSW 2006, Australia
*
Authors to whom correspondence should be addressed.
Polymers 2022, 14(19), 4025; https://doi.org/10.3390/polym14194025
Submission received: 13 July 2022 / Revised: 26 August 2022 / Accepted: 31 August 2022 / Published: 26 September 2022
(This article belongs to the Section Polymer Physics and Theory)

Abstract

:
Cellulose is the most abundant biopolymer on Earth, which is synthesized by plants, bacteria, and animals, with source-dependent properties. Cellulose containing β-1,4-linked D-glucoses further assembles into hierarchical structures in microfibrils, which can be processed to nanocellulose with length or width in the nanoscale after a variety of pretreatments including enzymatic hydrolysis, TEMPO-oxidation, and carboxymethylation. Nanocellulose can be mainly categorized into cellulose nanocrystal (CNC) produced by acid hydrolysis, cellulose nanofibrils (CNF) prepared by refining, homogenization, microfluidization, sonification, ball milling, and the aqueous counter collision (ACC) method, and bacterial cellulose (BC) biosynthesized by the Acetobacter species. Due to nontoxicity, good biodegradability and biocompatibility, high aspect ratio, low thermal expansion coefficient, excellent mechanical strength, and unique optical properties, nanocellulose is utilized to develop various cellulose nanocomposites through solution casting, Layer-by-Layer (LBL) assembly, extrusion, coating, gel-forming, spray drying, electrostatic spinning, adsorption, nanoemulsion, and other techniques, and has been widely used as food packaging material with excellent barrier and mechanical properties, antibacterial activity, and stimuli-responsive performance to improve the food quality and shelf life. Under the driving force of the increasing green food packaging market, nanocellulose production has gradually developed from lab-scale to pilot- or even industrial-scale, mainly in Europe, Africa, and Asia, though developing cost-effective preparation techniques and precisely tuning the physicochemical properties are key to the commercialization. We expect this review to summarise the recent literature in the nanocellulose-based food packaging field and provide the readers with the state-of-the-art of this research area.

Graphical Abstract

1. Introduction

Food packaging is vital to the food industry, which could protect the food from physical damage, chemical hazards, and biological contamination [1]. Synthetic polymer-based plastics [2], paper and paperboard [3], glass [4], and metal [5] are the most widely used packaging materials, and the choice of the packaging material is food-type-dependent. Nowadays, plastics derived from fossil resources are the most common food packaging materials. By far the most plastic, almost 40%, is used for packaging. Annually, approximately 500 billion plastic bags are used worldwide. After use, plastic products are generally incinerated or landfilled, causing serious environmental pollution. In addition, recycling and processing plastic packaging waste is estimated to cause an annual loss of USD 80–120 billion globally [6]. Today, with the promotion of “sustainable development”, new functional food packaging materials that are green, environmentally friendly, biodegradable, renewable, and sustainable have become the cutting-edge research direction in the food packaging field.
Cellulose is the most abundant biopolymer on Earth. It can be synthesized by trees [7], plants [8], sea animals (tunicates) [9], algae [10], and certain cellulose-secreting bacteria [11], and the properties of the cellulose are also found to be source-dependent [12]. However, cellulose derived from trees is still the most commonly investigated due to its high abundance and easy availability. In fact, paper made from woody cellulose has been used as packaging material for decades, though its low wet strength and poor barrier properties inhibit its further applications as packaging material to replace plastics [13]. More recently, cellulose fibers have been processed to nanoparticles through many different methods, such as acid hydrolysis [14], homogenization [15], grinding [16], blending [17], etc., which are normally terminated as nanocellulose [18]. Though the nomenclature of nanocellulose is still not yet standardized, it is commonly categorized into three types in terms of the preparation method and the morphological profile: CNC (cellulose nanocrystals), CNF (cellulose nanofibrils), and BC (bacterial cellulose) [19]. CNC with a rod-like shape with a width of 3–5 nm and a length between 50–500 nm is typically produced by mineral acid hydrolysis of lignocellulosic materials [20]. Another type of nanocellulose is CNF, with a width of 4–20 nm and a length between 500 nm and a few micrometers. CNF is typically produced via a mechanical treatment of lignocellulosic materials in a high-pressure homogenizer or a microfluidizer, achieving defibrillation and breakage of fibrils. Normally, a complementary pretreatment such as enzymatic treatment [21], carboxymethylation [22], or 2,2,6,6-Tetramethylpiperidine-1-oxyl (TEMPO) oxidation [23] is employed to facilitate the defibrillation, thus lowering the energy consumption [24]. BC is biosynthesized by specific bacteria, with many characteristics differing from CNC and CNF. BC is composed of ribbon-like fibers, with a width in the range of 0.01–0.10 μm and a length between one hundred nanometers to the micron level, and the fibers interweave to form a 3D network [25]. For all these three categories of nanocellulose, they generally have a high degree of polymerization (DP), high crystallinity, good thermal stability, and excellent mechanical properties, making nanocellulose a most promising candidate to develop bio-based nanocomposites for green packaging applications.
In this review, the sources, extraction, and properties of nanocellulose are summarized, and the potential of nanocellulose as the matrix, nanofiller, or coating materials to prepare advanced cellulose nanocomposites for food packaging is discussed: (i) the source-specific physicochemical properties of nanocellulose prepared by various cellulose sources, (ii) the inter-relation between preparation methods and the properties of the obtained nanocellulose, (iii) the fabrication strategies of cellulose nanocomposites aiming for potential food packaging applications, (iv) the performance of the thus-prepared cellulose nanocomposites as packaging materials, and (v) the state-of-the-art in the commercialization of nanocellulose on the market. We also discuss the opportunities and challenges of developing cellulose nanocomposites-based food packaging materials.

2. Source and Structure of Cellulose and Its Derived Nanocellulose

As one of the most widely distributed and abundantly available biopolymer materials in nature, cellulose has been widely used throughout human beings’ history. Cellulose is a linear natural biopolymer composed of β-D-glucopyranose, which has a chemical structure with repeating units of cellulose disaccharide (Figure 1) [26]. The repeating unit is then linked by a β-1,4-glycosidic bond by forming an acetal functional group between hydroxyl groups at the C4 position of the glucose unit and a hydroxyl group at the C1 position of the adjacent unit. The structural formula of cellulose is (C6H10O5)n, where n is the degree of polymerization (DP). There are hydroxyl groups (-OH) at the C2, C3, and C6 positions of each glucose unit [27]. The presence of these -OH groups render the cellulose a high chemical reactivity, which allows a series of chemical and physical modification of cellulose, so that the modified cellulose has the expected properties to meet the different needs for use.
The large number of -OH groups in the molecular structure of cellulose make it easy to form intramolecular and intermolecular hydrogen bonds, leading to the aggregation of cellulose molecular chains and the formation of crystalline supramolecular structure. The supramolecular structure of cellulose is divided into crystalline and amorphous regions [28]. In the crystalline region, many intermolecular hydrogen bonds exist, and the molecular chains are orderly packed. However, molecular interaction is weaker in the amorphous region than in the crystalline region, and the molecular chain arrangement is disordered with a loose structure. The XRD analysis of the crystalline structure of cellulose shows that the intensity of the X-ray diffraction peak in the crystalline region is high, while no specific diffraction peak is observed in the amorphous region [29].
In nature, cellulose can be found from many different sources, though the most common ones include wood, plants (cotton, wheat straw, sugarcane bagasse, ramie, hemp, flax, sugar beet pulp, etc.), tunicate, algae, and bacteria [30]. The synthetic mechanism and the properties of the cellulose are source-specific, shown in Figure 2, and the typical microscopic images of cellulose from different sources are presented in Figure 3.

2.1. Cellulose Source and Structure

2.1.1. Wood

Trees are the major source of cellulose in nature, in which the cellulose is synthesized through photosynthesis by cells (Figure 2a) [31]. As shown in Figure 3a, the woody cellulose fibers in nature are ribbon-like. Traditionally, the wood is subjected to a pulping process to obtain pure cellulose, which is then further processed into other materials for food, biomedical, environmental mediation, electronic devices, energy conversion, and other applications. In general, woody cellulose-based materials are easily available, renewable, biocompatible, and biodegradable, thus making them ideal raw materials for many advanced biomaterials. Woody cellulose fiber shows excellent physical properties and chemical stability and is a new material for a sustainable society and industrial ecological development [32].

2.1.2. Plant

Cellulose is widely present in most plants, having an annual production of 1011–1012 tons through photosynthesis, similar to trees. It can be obtained from many sources, mainly cotton, wheat straw, sugarcane bagasse, ramie, hemp, flax, etc. [33]. Plant fibers mainly comprise polymers such as cellulose, hemicellulose, lignin, wax, and pectin. Cellulose is the main component of plant fibers, composed of spirally entangled cellulose microfibrils bound together by an amorphous lignin matrix, for example, cotton cellulose, in Figure 3b. Lignin helps the plants protect against biological attack and acts as a reinforcing agent to strengthen them against gravity and wind, and hemicellulose acts as an adhesive between cellulose and lignin [34]. Plant species, climate, maturity, and soil conditions would affect the physical and chemical properties of plant fibers. Plant fibers are renewable, degradable, low cost, and widely available compared to resources such as oil, natural gas, and coal. Using plant fibers instead of nonrenewable resources is significant in alleviating the energy crisis and environmental pollution problems.
Figure 2. Biosynthesis of (a) wood cellulose, reprinted with permission from [35], 2016, ACS; (b) tunicate cellulose, reprinted with permission from [36], 2007, Springer; and (c) BC, reprinted with permission from [37], 2019, Frontiers Media S.A.
Figure 2. Biosynthesis of (a) wood cellulose, reprinted with permission from [35], 2016, ACS; (b) tunicate cellulose, reprinted with permission from [36], 2007, Springer; and (c) BC, reprinted with permission from [37], 2019, Frontiers Media S.A.
Polymers 14 04025 g002

2.1.3. Tunicate

Tunicates are the only animals that can produce cellulose in nature. The tunicate cellulose is generated through a cellulose synthase complex (Figure 2b), and the produced tunicate cellulose is characterized by a large diameter, high crystallinity, and great molecular weight (Figure 3c). Tunicate cellulose is embedded in a protein matrix to form a leathery mantle that can protect the animals from physical damage and predators. It has been reported that there are over 2300 species of tunicates worldwide, and tunicate cellulose has become a new type of biopolymer for many different applications [26].

2.1.4. Algae

Algae can also produce cellulose microfibers in their cell walls, such as green algae, red algae, and yellow-green algae [38]. It has been found that cellulose synthesis occurs at the plasma membrane-bound cellulose synthase, except for some algae that produce cellulosic scales in the Golgi apparatus. There are also great differences in microfiber structure between different algae, probably due to different biosynthesis processes (Figure 3d).

2.1.5. Bacteria

BC is a natural nanostructured polymeric material mainly produced by bacteria [39]. Like tunicate, the bacteria produce cellulose via cellulose synthase complexes (Figure 2c). BC is ribbon-like, with a width in the range of 0.01–0.10 μm, which is 2–3 orders of magnitude smaller than the diameter of plant cellulose (generally 10 μm), and the fiber length ranges from a few hundred nanometers to the micron level, and the fibers cross each other to form a mesh-like structure (Figure 3e,f). BC differs from plant cellulose in that it is not a structural component of the cell wall and therefore does not contain impurities such as hemicellulose and lignin but is a product of microbial metabolism [40,41]. BC not only has the properties of plant cellulose, but also has other, more outstanding advantages, such as high purity, high degree of polymerization, great crystallinity, high hydrophilicity, high permeability and air permeability, excellent Young’s modulus, good biocompatibility, etc. [42]. Under specific culture conditions, BC can be prepared by static fermentation, dynamic fermentation, and fermentation in special molds, showing different structural and performance characteristics.
Figure 3. SEM images of cellulose obtained from different sources: (a) softwood, reprinted with permission from [43], 2012, RSC; (b) cotton, reprinted with permission from [44], 2013, Elsevier; (c) tunicate, reprinted with permission from [45], 2015, Elsevier; (d) algae, reprinted with permission from [46], 2015, Springer; BC, (e) reprinted with permission from [47], 2016, RSC, and (f), reprinted with permission from [48], 2022, Elsevier.
Figure 3. SEM images of cellulose obtained from different sources: (a) softwood, reprinted with permission from [43], 2012, RSC; (b) cotton, reprinted with permission from [44], 2013, Elsevier; (c) tunicate, reprinted with permission from [45], 2015, Elsevier; (d) algae, reprinted with permission from [46], 2015, Springer; BC, (e) reprinted with permission from [47], 2016, RSC, and (f), reprinted with permission from [48], 2022, Elsevier.
Polymers 14 04025 g003

2.2. Nanocellulose Obtained from Different Sources

As mentioned above, cellulose can be extracted from many different sources, and the cellulose properties are source-dependent. After further processing to nanocellulose, the differences are still present in the final products. In Table 1, we summarized the typical geometrical characteristics and crystallinity index of nanocellulose originating from different cellulose sources. A more detailed discussion of the properties of various nanocellulose derived from different sources can be found in Section 3.

3. Preparation and Characteristics of Nanocellulose

Nanocellulose can be categorized into CNC, CNF, and BC, wherein CNC and CNF are nanocellulose prepared from biomass, and BC is nanocellulose prepared by microorganisms. The preparation of nanocellulose from biomass is mainly achieved by breaking down the cellulose fibers into nanoscale particles, while BC is a microbially produced cellulose on the nanoscale. Some aerobic nonpathogenic bacteria (for example, Taonella mepensis, Aerobacter, Azotobacter, Rhizobium) can produce BC in the form of exopolysaccharides at the gas–liquid interface, and BC combines the properties of both cellulose and nanomaterials [52]. The pretreatment and extraction techniques used to prepare nanocellulose must be carried out under specific conditions to improve the yield and obtain the desired specifications and properties [53]. Methods for producing nanocellulose vary by source and end application purpose. Table 2 summarizes the nanocelluloses prepared by different methods and their properties.

3.1. Pretreatment of Cellulose Fibers for Nanocellulose Production from Biomass

Appropriate pretreatment methods can change the properties of cellulose and endow the final nanocellulose with better properties. The pretreatment for the production of biomass nanocellulose mainly includes TEMPO oxidation, carboxymethylation, and enzymatic hydrolysis (Figure 4) [54]. After the pretreatment, the cellulose is either modified by functional groups or slightly broken down, thus facilitating the following extraction process by reducing energy consumption and uniforming the particle size.
Table 2. Preparation and properties of nanocellulose produced by different methods.
Table 2. Preparation and properties of nanocellulose produced by different methods.
Extraction MethodCelluloseDiameter (nm)Length (nm)Crystallinity Index (%)Degradation Onset Temperature (°C)Tensile Strength (MPa)References
H2SO4 hydrolysisMicrocrystalline cellulose10.8 ± 2.4111.2 ± 25.6/270114[55]
HCl hydrolysisBleached kraft pulp28.548188.2310/[56]
H2SO4 hydrolysis and sonicatedIndustrial pepper waste (Piper nigrum L.)33.4 ± 11.7/69.9300.6/[57]
HCl hydrolysis and sonication 50.7 ± 9.6/73.7291.5
H3PO4 hydrolysis, sonicated 67.8 ± 3.1/75.8298.3
Oxalic acid hydrolysis and sonicated 21.7 ± 4.9/77.8311.2
Citric acid hydrolysis and sonicated23.2 ± 0.6258.8 ± 58.476.4310.1
Acetic acid hydrolysis and sonicated 48.7 ± 9.4343.7 ± 2.378.3308.2
Alkaline treatment and blendingOil palm empty fruit bunch89///33.0[58]
Disc grinderRaw wood5 ± 3/67235233[59]
Ball millRaw sisal12.35/53.6%/92.73[60]
Ultrafine grinderUnbleached Eucalyptus kraft pulp38 ± 163000///[61]
High pressure homogenizationCellulose powder (cotton linters)46.4 ± 7.5417.7 ± 37.6/270114[55]
Homogenizer and sonication BC (K. oboediens R37-9)6.06 ± 0.96815 ± 0.9575.64/142[62]
High pressure homogenizationBleached softwood kraft pulp board18.85 ± 4.51/67.3/177.99[63]
GrinderHardwood bleached kraft pulp5.5 ± 1.6////[64]

3.1.1. TEMPO Oxidation

2,2,6,6-Tetramethylpiperidin-1-oxyl (TEMPO) is a water-soluble piperidine nitroxide radical. TEMPO oxidation method uses sodium hypochlorite (NaClO) as the main oxidant and TEMPO and sodium bromide (NaBr) as the catalyst, thus forming TEMPO/NaBr/NaClO as a new selective oxidation system, which can be used for the oxidation of cellulose (Figure 4a). Under specific conditions (pH 9~11), the primary hydroxyl group at the C6 position of cellulose can be selectively oxidized to the carboxyl group without changing the fiber morphology and crystallinity, thus improving the colloidal stability of the final CNF [65]. This suggests that carboxylate groups formed by TEMPO oxidation are selectively introduced into the surface of cellulose nanofibers, rather than the inner cellulose crystallites [66]. TEMPO-oxidized CNF has long and highly flexible network structures that are more uniform and better dispersed in the aqueous phase (Figure 5b,g) [67]. Kaffashsaie et al. used wood as a raw material to prepare nanocellulose through TEMPO oxidation. The dispersion is stable after even more than 72 h, which was better than the nanocellulose prepared directly by mechanical refining. This is due to the electrostatic repulsion of carboxylate ions among nanocellulose [59]. TEMPO oxidation is normally carried out in a mild environment, and nanocellulose can be obtained by simple mechanical treatment of cellulose after the reaction is completed, which greatly avoids the disadvantage of consuming more energy by simply using mechanical treatment to obtain nanocellulose. Unfortunately, the thermal stability of nanocellulose will be reduced due to the introduction of -COOH groups on C6 of cellulose nanofibers [68].
Figure 4. Different pretreatment methods for nanocellulose preparation: (a) TEMPO oxidation, reprinted with permission from [69], 2019, Academic Press Inc and [70], 2013, Elsevier; (b) carboxymethylation, reprinted with permission from [71], 2022, Elsevier; (c) enzymatic hydrolysis, reprinted with permission from [72], 2021, Elsevier.
Figure 4. Different pretreatment methods for nanocellulose preparation: (a) TEMPO oxidation, reprinted with permission from [69], 2019, Academic Press Inc and [70], 2013, Elsevier; (b) carboxymethylation, reprinted with permission from [71], 2022, Elsevier; (c) enzymatic hydrolysis, reprinted with permission from [72], 2021, Elsevier.
Polymers 14 04025 g004

3.1.2. Carboxymethylation

Carboxymethylation is considered to be another commonly used pretreatment method. Firstly, the cellulose was swollen in a mixture of aqueous sodium hydroxide, and then an organic solvent (usually, alcohols, such as isopropanol) was added under vigorous stirring. After that, chloroacetic acid (CAA) or sodium chloroacetate (SCA) is slowly added for carboxymethylation. The carboxymethylation reaction is based on the Williamson ether synthesis mechanism, in which the 1° and/or 2° hydroxyl (-OH) of cellulose can be etherified with carboxymethyl. The principle of carboxymethylation is to form more nucleophilic alkoxide groups after deprotonating the hydroxyl group (-OH) of polysaccharides in an alkaline solution. Then, -CH2COONa was introduced on cellulose by SN2 reaction between CAA or SCA and cellulose alkoxide (Figure 4b) [73]. Furthermore, by breaking the hydrogen bonds within the cellulose structure, the presence of bases can facilitate or initiate the reaction process and bring about uniform chemical changes in the single cellulose chains (Figure 5i) [74]. The occurrence of side reactions can be avoided under specific conditions by tuning chemical reagent concentration, reaction time, and temperature, thereby improving carboxylated nanocellulose properties and yields [75]. The degree of substitution (DS) of carboxylated nanocellulose is a key factor affecting its properties. The maximum substitution degree of carboxylated nanocellulose is 3, but when the substitution degree of carboxylated nanocellulose is too small (<0.4), carboxylated nanocellulose is not dispersible. In fact, DS increases significantly by increasing reaction temperature and time; however, the time has a greater effect on DS than temperature [76]. Sarmina et al. analyzed the effects of NaOH concentration, monochloroacetic acid (MCA) concentration, time, temperature, and cellulose particle size on DS, optimized the carboxymethylation reaction of cellulose, and prepared nanocellulose with DS up to 2.41 from corn husk [77].

3.1.3. Enzymatic Hydrolysis

Enzymatic hydrolysis is a biological treatment method that involves the digestion or modification of cellulose fibers with enzymes to promote further shortening and defibrillation of the cellulose (Figure 4c). Due to the good selectivity of the enzyme, the extracted nanocellulose does not change much after the enzymatic hydrolysis (Figure 5h) [68]. The main enzymes used for this purpose are cellulases, which break the β-1,4 glycosidic bonds in the cellulose molecule. More specially, the cellulase can only destroy the amorphous regions within the cellulose, thereby reducing the energy requirements for mechanical processing [78]. Dias et al. found that the content of fine particles after enzyme pretreatment increased from 55 ± 1% after laccase treatment to 62 ± 1% after cellulase treatment. These results showed that cellulase could expand the fiber by attacking the surface and interior of the fiber, making more water molecules enter the fiber [61]. However, different cellulases would produce different nanocelluloses. For example, endoglucanases can produce a mixture of CNF and CNC, while exoglucanases can produce nanocellulose complex structures [79]. The enzymatic hydrolysis method employs milder conditions, a promising green and sustainable method, although the reaction time is usually longer. Moreover, the high cost of enzymes is still a persistent issue, but reducing the cost by reusing enzymes by developing enzyme immobilization techniques can potentially solve this problem [80].
Figure 5. Nanocellulose prepared from different methods: (a) H2SO4, reprinted with permission from [81], 2021, Springer; (b) TEMPO, reprinted with permission from [82], 2015, nature; (c) H3PO4, reprinted with permission from [83], 2019, Springer; (d) HCl, reprinted with permission from [84], 2013, RSC; (e) maleic acid, reprinted with permission from [85], 2017, Wiley; (f) oxalic acid, reprinted with permission from [86], 2019, Elsevier; (g) TEMPO-oxidized CNF, reprinted with permission from [87], 2021, Springer; (h) enzymatic hydrolysis, reprinted with permission from [88], 2017, Springer. (i) carboxymethylation, reprinted with permission from [71], 2022, Elsevier.
Figure 5. Nanocellulose prepared from different methods: (a) H2SO4, reprinted with permission from [81], 2021, Springer; (b) TEMPO, reprinted with permission from [82], 2015, nature; (c) H3PO4, reprinted with permission from [83], 2019, Springer; (d) HCl, reprinted with permission from [84], 2013, RSC; (e) maleic acid, reprinted with permission from [85], 2017, Wiley; (f) oxalic acid, reprinted with permission from [86], 2019, Elsevier; (g) TEMPO-oxidized CNF, reprinted with permission from [87], 2021, Springer; (h) enzymatic hydrolysis, reprinted with permission from [88], 2017, Springer. (i) carboxymethylation, reprinted with permission from [71], 2022, Elsevier.
Polymers 14 04025 g005

3.2. Extraction Method of Nanocellulose from Biomass

After decades of intensive investigations, many different nanocellulose extraction methods have been developed and improved. We summarized the recent progress in the method development. In general, the nanocellulose preparation methods can be normally categorized into acid hydrolysis and mechanical treatment, or a combination thereof, and biosynthesis.

3.2.1. Acid Hydrolysis

Acid hydrolysis is the most commonly used method to isolate nanocellulose, in which strong acids easily hydrolyze the amorphous regions of whole cellulose fibers to produce CNC with reduced size. The obtained CNC has a similar morphology to the original cellulose and has a higher degree of crystallinity. Negatively charged nanocelluloses with abundant functional groups can also be produced by acid hydrolysis; for example, the nanocellulose with sulfate groups can be stably dispersed after sulfuric acid hydrolysis aqueous solutions due to surface charge repulsion [89]. The acids used for acid hydrolysis can be divided into inorganic acids (such as sulfuric acid, hydrochloric acid, phosphoric acid, etc.) (Figure 6a) and organic acids (such as oxalic acid, citric acid, maleic acid, etc.) (Figure 6b). After the reaction, the obtained nanocellulose is washed with deionized water, centrifuged to remove residual acid, and then suspended in distilled water. Sometimes, the acid hydrolysis is always followed by a mechanical process (such as high-pressure homogenization, microfluidization, ultrasound, etc.) to disperse the CNC into a uniform and stable suspension. The properties of the CNC suspension are closely related to the hydrolysis conditions (acid type, concentration, reaction temperature, and time) and the mechanical treatment conditions [90]. The main disadvantage of this process is generating a lot of acid-containing wastewater, which must be treated until discharged into the environment.

Inorganic Acid Hydrolysis

By using H2SO4, uniform and short CNC with narrow polydispersity and high crystallinity of over 90% can be produced, which also shows good colloidal stability (Figure 5a). In addition, during the sulfuric acid hydrolysis, sulfate groups are generated on the surface of nanocellulose, thereby forming a negatively charged layer on the CNC, which further improves the dispersibility of CNC. However, residual sulfate groups (-SO42−) can induce dehydration, resulting in lower thermal stability. In addition, when higher hydrolysis temperatures above 45 °C are applied, H2SO4 hydrolysis not only removes the amorphous region but also partially dissolves the crystalline region, resulting in a size reduction of CNC [91]. Due to the small size of the CNC, this provides a large surface area for heat treatment, which can also lead to a reduction in thermal stability [92]. Worse, H2SO4 hydrolysis sometimes causes the degradation of cellulose due to its strong oxidizing property, which negatively impacts the large-scale production and application of CNC.
Hydrochloric acid (HCl) hydrolysis is another method used to prepare CNC. Unlike H2SO4 hydrolysis, HCl hydrolysis provides unmodified and uncharged CNC, which can form more intermolecular hydrogen bonds (Figure 5d) [93]. CNC prepared by HCl hydrolysis can form stable Pickering emulsions, whereas H2SO4-hydrolyzed CNC does not possess this property [94]. Since the surface of CNC prepared by HCl hydrolysis is not charged, it is easy to flocculate in water and does not produce a stable suspension, but its thermal stability is higher than that prepared by H2SO4 hydrolysis [95]. Shang et al. used cetyltrimethylammonium bromide (CTAB) to modify the nanocellulose prepared by HCl hydrolysis and found that the stable CNC dispersion was formed by using low-concentration CTAB, and no aggregation occurred after 7 days of storage [56].
Hydrolysis of cellulose by phosphoric acid (H3PO4) produces CNC with phosphoric acid half-ester groups, so it is slightly charged [96]. This colloid’s stability is much better than that produced by hydrochloric acid hydrolysis, which does not introduce surface charges [20]. The CNC produced by phosphoric acid hydrolysis has a more inhomogeneous profile compared with those prepared by sulfuric acid hydrolysis, so easier flocculation can be expected, but it shows better thermal stability (Figure 5c) [97].
Figure 6. Different preparation methods of nanocellulose (a) inorganic acid hydrolysis, reprinted with permission from [98], 2022, Elsevier; (b) organic acid hydrolysis, reprinted with permission from [99], 2021, Elsevier; (c) mechanical treatment, reprinted with permission from [71], 2022, Elsevier and [100], 2021, Elsevier; (d) microbial synthesis, reprinted with permission from [101], 2022, Elsevier and [37], 2019, Frontiers Media S.A.
Figure 6. Different preparation methods of nanocellulose (a) inorganic acid hydrolysis, reprinted with permission from [98], 2022, Elsevier; (b) organic acid hydrolysis, reprinted with permission from [99], 2021, Elsevier; (c) mechanical treatment, reprinted with permission from [71], 2022, Elsevier and [100], 2021, Elsevier; (d) microbial synthesis, reprinted with permission from [101], 2022, Elsevier and [37], 2019, Frontiers Media S.A.
Polymers 14 04025 g006

Organic Acid Hydrolysis

Organic acids are greener alternatives to inorganic acids for CNC preparation because they are less acidic and corrosive. Since most organic acids have lower boiling points than inorganic acids, they are easier to be recycled. Similar to H2SO4 and H3PO4 hydrolysis, the organic acid may also modify the CNC surface, especially when strong inorganic acids (for example, H2SO4) are present as a mixture. Wang et al. used bleached eucalyptus kraft pulp as raw material and applied a small amount of sulfuric acid (5–10 wt%) to effectively improve the hydrolysis efficiency of formic acid. Under the best reaction conditions, the maximum yield of CNC was 70.65% [102]. The organic acids usually introduce hydrophobic moieties on the surface of CNC, facilitating CNC dispersion in nonpolar organic solvents (Figure 5e,f). The introduction of negatively charged carboxyl groups, for example, by oxalic acid hydrolysis, can improve the stability of CNC suspensions [99].

3.2.2. Mechanical Process

Cellulose fibers can be mechanically processed by different mechanical methods to extract nanocellulose, and the most studied techniques include ultrasonication, ball milling, microfluidization, and high-pressure homogenization (Figure 6c) [53]. Compared with acid hydrolysis, the mechanical process is simple, efficient, and does not require chemical solvents. However, the main disadvantages of these processes are high energy consumption and easy blockage of equipment, though the abovementioned pretreatments can alleviate these problems. Dias et al. prepared CNF using an ultrafine mill after laccase pretreatment. It has been found that compared with CNF prepared directly by mechanical defibrillation without pretreatment, the cellulose fibers delaminated faster and made the microfibrils easier to be dispersed, resulting in a 42% reduction in energy consumption by lowering energy consumption from 10.5 kWh/kg to 6.1 kWh/kg [61].
The extraction of nanocellulose by ultrasonication is a mechanical method based on ultrasonic waves. During ultrasonication, liquid molecules can absorb ultrasonic energy and create a cavitation effect, leading to the formation, expansion, and explosion of microscopic gases. However, this processing generates a lot of heat; therefore, it is usually performed in cooling equipment to control the temperature under the acceptable level [103]. The use of ultrasonic treatment after acid hydrolysis can improve CNC yield and properties. Ultrasonic treatment can also be used in the pretreatment stage of the nanocellulose production process. After sonication, lignocellulosic feedstocks undergo various changes, such as the destruction of cell wall structure, the increase in specific surface area, and the decrease in the DP of the cellulose component [104].
Another mechanical process is homogenization. In this process, the cellulose suspension (about 2%) passes through the tiny gap between the homogenization valve and the impingement ring; the generated high pressure could produce several destructive forces (including cavitation, turbulence, and shear effects). These destructive forces can significantly disrupt the cellulose hydrogen bond network structure, reducing the size of the fibers to the nanometer range, which in turn affects the nanocellulose properties [105].
Microfluidization is another most common mechanical treatment technique, and it works best for the delamination and fibrillation of cellulose fibers [106]. A microfluidizer consisting of a booster pump accelerates the fiber suspension (0.5–2% by mass) under high pressure, going through an ultrathin chamber with a specific geometry (such as a Z-shape or a Y-shape), and the fibers pass through the channel walls. During this process, the strong shear and shock effects fibrillate cellulose to nanoscale dimensions, and these nanoparticles have a uniform diameter of less than 100 nm. Wang et al. used the blended software kraft pulp as the raw material and applied a high-shear fluid microfluidizer to prepare the nanocellulose. The width and length distributions are about 10~40 nm and 600~1000 nm, respectively. The high aspect ratio makes it a potential candidate as a reinforcing material [102].
Ball milling is also a useful technique to prepare nanocellulose, which is a fast, easy-to-operate, and cost-effective method, showing great potential for industrial applications. During this treatment, the cellulose suspension is kept in a hollow cylindrical container partially filled with spherical tools (made of zirconia, ceramic or metal, etc.). Cellulosic fibers are broken down by the energy released in ball-to-ball collisions and the high grinding energy from ball-to-wall friction as the container rotates [103]. When nanocellulose is produced using this process, wet milling is required to maintain the fibrous state and prevent serious defibrillation to an amorphous state [18]. In addition, the shape of the balls, the weight ratio of the balls to the cellulose, the treatment time, and the moisture content are the influencing factors for the preparation of nanocellulose by ball milling [105].
The aqueous counter collision (ACC) method was first developed by Kondo, in which equivalent aqueous suspensions of cellulose are ejected from dual nozzles under high pressure of 200 MPa and collide at high speed to destroy the weak hydrogen bonds in cellulose, thus producing nanocellulose without chemical modification. The number of collisions and the collision pressure are the influencing factors to determine the performance of the obtained nanofibers [107]. More interestingly, compared with other CNFs, CNF prepared by ACC (ACC-CNF) has a relatively hydrophobic surface [108], since ACC could increase the Iα Phase transition to a more stable Iβ Phase, while keeping its crystalline structure unchanged [107]. Yokota et al. further modified the ACC-CNF through acetylation without changing the original nanofiber morphology, which has proven to improve the dispersibility of the nanocellulose in water and stabilize plastic resin particles in water [108]. Ishida et al. prepared a Pickering emulsion by adsorbing ACC-CNF onto the O/W interface, which was then used as a reaction platform for chemical modification to introduced acetyl groups due to the increased surface [109].

3.3. Preparation of BC

The performance of BC depends on the composition of the culture medium and the used bacteria species (Table 3). BC is a natural nanostructured polymeric material mainly produced by bacteria. Similar to tunicate, the bacteria produce cellulose via cellulose synthase complexes (Figure 6d). The proper removal of impurities from the BC matrix helps to improve its mechanical properties. However, the production of BC by microbial fermentation requires using nutrient-rich mediums, in which the HS medium is the most-employed one. Recent work on an improved HS-based medium has focused on alternative sources that can provide a carbon source. Huang et al. used molasses instead of glucose as the carbon source in HS medium, and the produced BC shows improved stability, which is not easy to undergo oxidation reaction and physical modification. The BC yield of 6.67 g is obtained at 1.1 g/L molasses, equivalent to 5 g/L glucose [110]. Nascimento et al. cast BC-derived CNC to produce films, and the results showed that 5% concentration could result in a film with significantly improved tensile strength (from 36.9 to 46.5 MPa), increased elongation at break (from 8.06 to 13.5%), and lower water vapor permeability (17%). In addition, the water resistance of the film was improved, while showing great biocompatibility with Caco-2 cells [111]. Nam et al. used BC as a raw material to prepare CNC-containing aldehyde groups at the surface, which was then combined with silk sericin through physical enhancement and chemical crosslinking to prepare composite films. Compared to the control silk sericin film, the prepared composite film showed higher mechanical properties, better UV blocking, and waterproof and antioxidation properties [112].

4. Fabrication Strategies of Cellulose Nanocomposites for Food Packaging

Available studies have shown that nanocellulose is renewable and environmentally friendly, making it a promising candidate for preparing biomaterials for food packaging. To meet the market requirements for food packaging materials, such as strength, water, and oil resistance, nanocellulose-based composites have been intensively explored [115]. The widely used production methods of nanocellulose-based composites for potential food packaging applications are summarized in Table 4.

4.1. Solution Casting

One of the easiest ways to make nanocellulose-based packaging materials is solution casting. In this method, two or more solutions are mixed and cast onto a Teflon dish; the following evaporation of the solvent, normally water, will induce the nanocellulose to self-assemble, and a film is then obtained (Figure 7a). Krstić et al. prepared nanocellulose films to encapsulate insoluble drugs by mixing with polyethylene oxide (PEO) polymers through solution-casting. The increased porosity and uniformly sized particles improved the wettability and increased the surface area of the film, making it satisfactory for the controlled release of the drugs [126]. Oldoni et al. extracted nanocellulose from mango pulps, which was further cast in solution to produce a thin, light-weight, highly water vapor permeable and ductile film [127]. By taking advantage of the good film-forming properties of the seaweed-derived alginate and carrageenan, Rajeswari et al. blended nanocellulose with alginate and carrageenan to produce a film with strong mechanical toughness [128].

4.2. Layer-by-Layer Assembly

The Layer-by-Layer (LBL) assembly technique was originally proposed by Decher et al. (1997). The mixed solution is organized layer-by-layer through this method to form a thin film. This common technique allows for the facial preparation of films and coatings, which has the advantages of little space requirement and easy operation, thus allowing a wide range of applications (Figure 7b). The LBL technique could produce a wide range of films with different properties because it could take advantage of different polymer characteristics. Koca et al. employed the LBL technique to modify food packaging materials with lysozyme, nanocellulose, and Arabic gum, showing good barrier properties [129]. Yu et al. conducted various studies on the use of LBL technology to prepare composite films for potential food packaging applications, in which nanocellulose, nanosilver, carrageenan, and chitosan were used to achieve the great antimicrobial activity, good breathability, and excellent water barrier properties. The obtained film is nontoxic and effective in inhibiting common pathogenic bacteria [130].

4.3. Extrusion

The fabrication of nanocellulose films by extrusion is a more productive and economical method and is mostly used in industrial production. With this technique, the packaging films can be produced on a large scale. Normally, the polymer is extruded through the extruder equipped with a die made of metal in the desired pattern to form the desired shape (Figure 7c). For example, extrusion can be used to produce tubular nanocellulose films [131]. Karkhanis et al. have prepared PLA/CNC nanocomposite films by extrusion. At 25 °C and relative humidity of 50% or more, this laminated film can extend the shelf life of biscuits by more than 40% compared to conventional PLA films [132]. In another study, PLA/CNC nanocomposite films made by extrusion showed good barrier properties against water vapor and oxygen [133].

4.4. Direct Coating Method

The coating method involves the application of edible and environmentally friendly natural polymers, which are either dipped or coated directly onto the food surface. These coating substances are commonly based on nanocellulose and other naturally extracted bioactive compounds. Studies have found that the films made of these compounds have good antibacterial properties and better performance in preventing moisture loss and controlling the food oxidation rate (Figure 7d). Using such biopolymers, such as nanocellulose, as a coating material can effectively control the loss rate of its nutrients in the environment. Quintana et al. prepared a chitosan-CNC edible coating material composed of essential oils and bioactive plant extracts and applied it directly to strawberries. The results showed that this coating could effectively inhibit microbial growth and delay fruit spoilage [134].

4.5. Hydrogels

Hydrogels are known to be hydrophilic materials in which the polymer molecules are bonded to each other by chemical and physical interactions. Nanocellulose hydrogel has various applications in food packaging (Figure 7e). For example, it has been found to increase the wetness of the food and improve the aesthetics and biodegradability of the food packaging materials. Lu et al. prepared a hydrogel-based freshness indicator using bagasse nanocellulose, which could accurately indicate the degree of discoloration of chicken [135]. Moradi et al. used hydrogel technology to produce a BC-based biofilm that can be used as a new pH-sensing indicator for evaluating the freshness of fish [136]. Pourjavaher et al. developed a food-grade film based on BC nanofibers and extracts of red kale (Brassicaceae), an effective pH indicator at the food contact surface [137].

4.6. Spray Drying Method

As the name suggests, the spray drying method sprays the polymer onto the surface of the food by pushing the polymer solution through a nozzle. The sprayed amount of the polymers could be adjusted by controlling the desired shape of the spray nozzle and the concentration of the solution. After the spray drying, a thin film is formed on the food surface for protection (Figure 7f). Throughout this process, the liquid is first conveyed to a peristaltic pump. Due to the force of the compressed air, the liquid is atomized into smaller droplets and sprayed out. In turn, these small droplets are evaporated immediately by the hot air to form dry particles [119]. When this spray drying method is used based on the solution containing CNC and polylactic acid, the air permeability of the food was found to be significantly lowered. Rojas-Lema et al. blended the copper(II) sulfate-decorated chitosan particles and nanocellulose, spray drying was used to coat these substances onto fruit surfaces to extend shelf life [138].

4.7. Electrostatic Spinning Technology

Electrostatic spinning technology is a simple, flexible, and operable emerging technology. It is mainly used to create membranes using the electrostatic force-enabled spinning technology. This process facilitates the manufacture of biopolymers, which could include both natural and synthetic polymers, and a wide variety of fiber-shaped polymeric materials can be formed (Figure 7g). Pasaoglu et al. employed electrostatic spinning technology to produce nanocellulose membranes with greater environmental friendliness, sufficient surface functionality, good mechanical properties, and excellent shape stability [139]. Maria Leena et al. used electrostatic spinning to prepare corn protein nanofibers containing nanocellulose and resveratrol, which were effective in controlling water loss of food and maintaining the freshness of apple slices for six hours [140].

4.8. Micronanotechnology and Nanoemulsions

Micro- and nanoencapsulation are techniques for encapsulating various bioactive compounds in capsules to protect them from harsh external environments. It is also considered a protective layer for various components (Figure 7h). The functional properties of microcapsules are related to encapsulation efficiency, size, morphology, stability, and release characteristics. This technology involves the formation of polymeric films into small continuous droplets capable of covering solids or liquids. Nanoencapsulation refers to the encapsulation of bioactive substances built on the nano- and microscale. Nano- and microencapsulation is receiving increasing attention as an effective method to improve the bioavailability of bioactive substances because most bioactive agents are unstable in normal environments and deteriorate in contact with water and air [141]. Many available studies have found that using micro- and nanoencapsulation techniques can effectively preserve food products. Salvia-Trujillo et al. had developed a biodegradable nanocellulose-sodium alginate-based film containing lemongrass essential oil, which showed significant antibacterial activity and great browning reaction-inhibiting effects when they were tested to pack fresh-cut Fuji apples, having an extended shelf life of up to two weeks [142]. This environmentally friendly nanocomposite can be used to extend the shelf life of perishable foods.
Nanoemulsions are prepared by adding certain emulsifiers in order to stabilize the oil and water phases, thus producing transparent dispersions with lower viscosity and more stable thermodynamics. As the name suggests, nanoemulsions are emulsion systems with particle sizes between 5 and 100 nanometers and are used specifically for industrial applications, including food, pharmaceuticals, nutritional products, oil extraction, and environmental and agricultural products. Molet-Rodríguez et al. prepared a nanoemulsion containing orange essential oil and nanocellulose, which was employed to stabilize apple juice and showed great bactericidal properties [143]. In general, compared to conventional emulsions, nanoemulsions are normally more effective in inhibiting bacteria.
Figure 7. Fabrication strategies of nanocellulose-based composites: (a) solution casting, reprinted with permission from [144], 2017, Universidade Federal do Rio de Janeiro—UFRJ; (b) Layer-by-layer assembly (LBL), reprinted with permission from [54], 2019, Elsevier; (c) extrusion method, reprinted with permission from [145], 2022, Elsevier; (d) direct coating method, reprinted with permission from [146], 2021, Elsevier; (e) hydrogels, reprinted with permission from [147], 2022, Elsevier; (f) spray drying method, reprinted with permission from [148], 2022, Elsevier; (g) electrostatic spinning, reprinted with permission from [149], 2022, Elsevier; (h) micro-nanotechnology and nanoemulsions, reprinted with permission from [150], 2021, Elsevier; (i) adsorption, reprinted with permission from [151], 2022, Elsevier.
Figure 7. Fabrication strategies of nanocellulose-based composites: (a) solution casting, reprinted with permission from [144], 2017, Universidade Federal do Rio de Janeiro—UFRJ; (b) Layer-by-layer assembly (LBL), reprinted with permission from [54], 2019, Elsevier; (c) extrusion method, reprinted with permission from [145], 2022, Elsevier; (d) direct coating method, reprinted with permission from [146], 2021, Elsevier; (e) hydrogels, reprinted with permission from [147], 2022, Elsevier; (f) spray drying method, reprinted with permission from [148], 2022, Elsevier; (g) electrostatic spinning, reprinted with permission from [149], 2022, Elsevier; (h) micro-nanotechnology and nanoemulsions, reprinted with permission from [150], 2021, Elsevier; (i) adsorption, reprinted with permission from [151], 2022, Elsevier.
Polymers 14 04025 g007

4.9. Adsorption

Adsorption is the process in which a solid material uses its own properties to hold a gas or liquid as an encapsulated film on its surface. Adsorbents incorporated with various polymers, such as nanocellulose and its derivatives, play an important role in improving packaging systems (Figure 7i). Jorge Padrão developed BC films adsorbed with bovine Lactoferrin (BLF) at the surface, which showed a strong bactericidal effect against E. coli and Staphylococcus aureus [124].

5. Performance of Nanocellulose-Based Composites as Food Packaging Materials

Food packaging materials are an extremely important part of the food processing industry and have always been the research focus in the food field. As ideal food packaging materials, they should protect commodities, maintain food quality stability, increase commercial food value, promote sales, and facilitate storage and logistics [152]. Non-biodegradable polymers derived from fossil fuels are the most used materials in food packaging. With increased focus on global environmental issues, the development of biodegradable polymer materials has gained great interest. In recent decades, nanocellulose has been predominantly employed to create biocomposites because of its green source, high specific surface area, high crystallinity, and nontoxic and biodegradable qualities. Many research works have proven that combining nanocellulose and other substances may provide beneficial functionalities, such as barrier characteristics, mechanical properties, antibacterial properties, etc. [153]. The application and advantages of cellulose nanocomposites in the food packaging are summarized in Table 5.

5.1. Barrier and Mechanical Properties

For ideal food packaging materials, barrier and mechanical properties are critical to realistic applications. On the one hand, good barrier properties could protect the food from gas and moisture and slow the oxidation reaction and spoilage; on the other hand, the excellent mechanical properties would avoid physical and chemical damage during transportation and sales.
Shi et al. successfully developed cellulose-based food wrapping paper with high barrier and antibacterial properties by constantly depositing multilayer films on the surface of the paper using chitosan (CS) and carboxymethylated nanocellulose. The obtained multilayer coating not only enhanced the paper’s resistance to grease, oil, water, air, and water vapor, but also improved the paper’s mechanical strength. The modified wrapping paper exhibited no visible cytotoxicity and had an antibacterial rate of 95.8% against E. coli and 98.9% against Staphylococcus aureus [154].
CNC and garlic extract (GE) from garlic peel were blended with chitosan to prepare biocomposite films. UV barrier, thermal and mechanical properties, biodegradability, and antibacterial activities were tested on the films. CNC enhanced tensile strength, Young’s modulus, and elongation, compared to chitosan films, but decreased film transparency. On the other hand, the combination of CNC and GE slightly lowered the mechanical properties. The inclusion of CNC reduced the transparency of the film marginally, whereas the addition of GE dramatically increased the UV barrier properties. The integration of CNC and GE did not influence the thermal stability of the chitosan films. The chitosan composite films’ degradability rate was greater than that of neat chitosan films. The antibacterial characteristics of films were investigated against E. coli, Streptomyces griseorubens, Streptomyces alboviridis, and Staphylococcus aureus, which found that GE in composite films significantly inhibited bacterial growth. Due to the improved physical properties and better antibacterial activity, chitosan films containing both CNC and GE from garlic peel showed potential as active food packaging materials [155].
Translucent films were made from faba bean protein isolate (FBP) using the solution casting method, reinforced with varied CNC content (1, 3, 5, and 7 wt%) prepared by acid hydrolysis of pinecones, while glycerol was used as a plasticizer. The FTIR and SEM data confirmed that intramolecular interactions between CNC and proteins could induce a more compact and uniform film. These interactions had a favorable impact on mechanical strength, as seen by greater tensile strength and Young’s modulus compared to control films, though much stiffer films were expected as the CNC content increased. The addition of CNC increased the thermal stability of the FBP films by raising the typical onset degradation temperature. Furthermore, the linkages formed between CNC and proteins reduced the water affinity of the films, resulting in a decrease in moisture content and water solubility as well as an increase in water contact angle, resulting in more hydrophobic films as the CNC content in the matrix increased [163].
CNC modified by TEMPO oxidation (TM-CNC) was used to improve the performance of canola protein-based films. Varied weight ratios of modified (TM-CNC) and unmodified nanocrystalline cellulose (U-CNC) were tested. 19.61% of initial -OH groups were transformed to -COOH groups by TEMPO oxidation. The addition of U-NCC and TM-NCC boosted tensile strength substantially, with the greatest value of 8.36 MPa for 5% TM-NCC, compared to 3.43 ± 0.66 MPa for control films. In contrast to the control, both U-NCC and TM-NCC improved the water barrier and thermal characteristics of the films [164].
Cellulose nanofibrils-CNF with less than 1% of lignin and lignocellulose nanofibrils-LCNF with 16% of lignin were blended in various ratios to prepare composite films. The inclusion of LCNF in the formulations increased the films’ antioxidant and UV-blocking characteristics, as well as their mechanical and barrier properties. The addition of 25% LCNF to CNF films improved mechanical properties (94% increase in tensile stress and 414% increase in tensile strain at break) while lowering the water vapor transmission rate by 16% and oxygen transmission rate by 53%. The presence of nanocelluloses with varied chemical compositions and morphologies have contributed to the improved performance. Moreover, the presence of lignin in LCNF helped to increase interfacial adhesion between CNF and avoid the formation of accessible pathways for gas molecules [165].

5.2. Antibacterial Property

Food rich in nutrients and water is prone to microbial spoilage, thus leading to great economic loss. In order to prevent that, food packaging materials with satisfactory antibacterial properties are always necessary.
Thongsrikhem et al. used cinnamaldehyde as a crosslinking agent and an antibacterial ingredient to make a gelatin-BC nanocomposite membrane (GCB). Heat treatment at 120 °C for 3 h increased the reaction of the amine group with the aldehyde group of cinnamaldehyde via Schiff base and Michael addition, lowering the GCB film’s water solubility. The addition of BC to gelatin increased the composite film’s tensile strength and decreased its water vapor permeability. The GCB film was nontoxic to L929 cells and possessed significant antibacterial action against E. coli and S. aureus [156].
He et al. developed a coating material containing CMC and CNC with immobilized AgNPs (CNC@AgNPs) in varying proportions. Compared to uncoated paper, CMC/CNC@AgNPs showed better tensile strength, lower water vapor and air permeability, and greater antibacterial activity against E. coli and S. aureus. Furthermore, due to the immobilization effect of AgNPs on CNC, the release rate of AgNPs from the coated paper was greatly decreased. When strawberries were packaged using CMC/CNC@AgNPs-coated paper in ambient conditions, strawberries preserved better shape than unpackaged strawberries, and the shelf life was extended to seven days [157].
An antibacterial composite film based on sodium alginate (SA)/CNF containing peanut red skin extract (PSE) was created by crosslinking with Ca2+. The results showed that the SA/CNF/Ca2+/PSE (SCCP) film had high mechanical strength, great water resistance, and outstanding UV barrier properties. The films had a higher radical scavenging activity in the ABTS assay than the DPPH assay, especially in the presence of 10% ethanol; the maximal ABTS scavenging activity was 99.28%. When the film was applied to pack fruits, the weight loss of fruits was lower for the SCCP film than for the control group. Furthermore, both gram-negative and gram-positive bacteria were successfully prohibited [158].

5.3. Intelligent Packaging

Compared to traditional packaging materials, intelligent packaging has attracted great interest in recent years, and the research mainly focuses on environmentally sensitive materials to indicate the food quality changes.
Moradi et al. developed a unique intelligent colorimetric marker by employing anthocyanins from black carrots and BC nanofibers to check the freshness of rainbow trout and carp slices. Anthocyanin-BC composite film could detect pH change in the packaging materials as storage time increases, and the indicating label changed color correspondingly (Figure 8a). By comparing the label color to the standard color, consumers may visually determine the freshness of fish. The indicator is simple to make, inexpensive, ecologically acceptable materials, and easy to use, thus showing great potential [136].
Shi et al. prepared an intelligent pH-sensitive membrane by combining cyanidin-3-glucoside (C3G) and BC, tested as a tilapia freshness indicator. The results revealed that when BC’s crystallinity increased, the mechanical characteristics of C3G films increased dramatically. BC-C3G films were enhanced in terms of crystallinity and transmittance. Naked eyes can clearly see the color changes of BC-C3G films throughout the freshness monitoring process (Figure 8b), and it has a dependable color response (ΔE) and high sensitivity to TVB-N and TAC changes [166].
Figure 8. Application of nanocellulose indicator for food freshness monitoring: (a) fish, reprinted with permission from [136], 2019, Elsevier; (b) tilapia fillets, reprinted with permission from [167], 2022, Elsevier; (c) chicken breast, reprinted with permission from [135], 2020, Elsevier.
Figure 8. Application of nanocellulose indicator for food freshness monitoring: (a) fish, reprinted with permission from [136], 2019, Elsevier; (b) tilapia fillets, reprinted with permission from [167], 2022, Elsevier; (c) chicken breast, reprinted with permission from [135], 2020, Elsevier.
Polymers 14 04025 g008
As a colorimetric freshness indicator for detecting the freshness of chicken breast, a sugarcane bagasse nanocellulose-based hydrogel was produced. In this process, the nanocellulose was cross-linked by Zn2+ to obtain a robust self-standing hydrogel. The pH-responsive dyes (bromothymol blue/methyl red) were incorporated into the hydrogel, which changed color depending on the freshness of the chicken sample. Since CO2 levels rose with chicken deterioration due to microorganisms’ growth, the indicator hydrogel’s optical color changed from green to red on the third day (Figure 8c), indicating that the bacterial counts (CFU/g) had exceeded the acceptable limit for human intake. This innovative colorimetric freshness indicator produced with a nanocellulose hydrogel responds quickly to chicken deterioration and is intended to make bagasse nanocellulose more useful as a value-added material in smart packaging [135].

5.4. Preservation

The shelf life of food is important since the food products need a few days or even several months until they are delivered to the customers. In order to preserve the food better and extend the shelf life of the products, some kinds of effective packaging are vital.
CNF composite films containing glucose-derived carbon dots (GCD) and N-functionalized GCD (NGCD) were prepared by Ezati et al. The results showed that GCD and NGCD could effectively block the ultraviolet radiation and increase the water vapor permeability of the membrane without affecting the mechanical properties. At the same time, the composite films were resistant to oxidation, with a 99% ABTS and 80–85% DPPH free radical scavenging rate. When these films were applied to citrus or strawberry, they could effectively inhibit the growth of fungi and prolong their shelf life by 2–10 days. In addition, they reduced the weight loss caused by transportation and storage, thus ensuring the freshness of food and improving the economic benefits [166].
A new technique, coaxial 3D printing, was used to create cellulose nanofibers (CNF)-based labels with dual functionalities of fruit freshness preservation and visual monitoring. The shell of fibers was created with CNF-based ink containing blueberry anthocyanin, which had excellent formability and printing fidelity, as well as sensitive visual pH-responsiveness for freshness monitoring. Chitosan containing 1-methylcyclopropene (1-MCP) was injected into hollow microchannels of fibers, where 1-MCP was trapped by the electrostatic effect of chitosan and CNF. The 3D printed labels extended the shelf life of litchis by 6 days while also sensibly indicating variations in freshness, as proven by Headspace-Gas Chromatography-Ion Mobility Spectrometry [168].

6. Industrialization of Nanocellulose Production Worldwide

In recent years, the production of nanocellulose from biomass resources has become a hot topic, and intensive investigations have been carried out worldwide. Under this driving force, nanocellulose production has gradually developed from lab-scale to pilot- or even industrial-scale. In Table 6, we summarized the representative companies and institutes producing nanocellulose and pushing the products into markets. It can be seen that the pilot and industrial production lines of nanocellulose are currently mainly located in developed countries, such as the United States, Canada, Japan, Sweden, Finland, etc. As a representative enterprise in the preparation of CNC, CellForce in Canada developed a pilot production line based on sulfuric acid hydrolysis to prepare CNC in 2012, with a production capacity of 1 ton per day. In 2012, the U.S. Forest Service started up the first nanocellulose production plant in the United States in Wisconsin, mainly based on the sulfuric acid hydrolysis method to prepare CNC (10 kg/d) and the grinding method to prepare CNF (1000 kg/d) [169]. In 2014, the University of Maine built up a CNF pilot production line based on the mechanical refining method, with a production capacity of 1 ton per day. In April 2015, American Process realized the commercial production of nanocellulose based on the AVAP method with a production capacity of 1 ton per day, and the nanocellulose products show controllable morphology and surface hydrophilicity and hydrophobicity. It produced lignin-containing nanocellulose, which can realize the reinforcement and filling of plastics [170]. Founded in 2016, Cellulose Lab is a Canadian company providing many different types of nanocellulose products, including CNC, CNF, and BC, and it has become one of the top suppliers of cellulose nanomaterials in the world [171]. Founded in 2015, FineCell in Sweden has invented an oxalic acid-based technology to prepare a nanocellulose product in dry powder, making it easier to be used by mixing with plastics [172]. In 2019, Ocean TuniCell AS in Norway successfully produced tunicate nanocellulose on a large-scale based on microfluidization combined with different pretreatment technologies, such as TEMPO-mediated oxidation, carboxymethylation, and enzymatic treatment. Right now, it has commercialized this unique animal nanocellulose by developing them in to TUNICELL 3D-bioinks [173].
Over the last decades, the industrialization of nanocellulose showed the greatest progress in Japan. Especially after the development of TEMPO oxidation method by Professor Akira Isogai from the University of Tokyo, many Japanese companies have put a lot of effort into commercializing this functionalized nanocellulose [174]. For example, Nippon Paper has begun to build production lines based on this method, with a designed capacity of 500 tons CNF/year [175]. In January 2017, Oji Paper announced that they have started producing CNF with high viscosity and thixotropy based on phosphoric acid pretreatment, followed by mechanical treatment, with a production capacity of 40 tons of CNF per year. Regarding the application of nanocellulose, Oji Paper and Mitsubishi Chemical jointly launched a commercialized nanocellulose (CNF) film in 2013, which can be used to manufacture large-scale displays and solar cells. Oji Paper has also taken advantage of the dense characteristics of CNF and cooperated with Nikko Chemicals to develop its application in cosmetics. Mitsubishi Pencil used the thixotropic properties of CNF as a tackifier for ink, and successfully developed a gel ink ballpoint pen with good thixotropy. Its ink viscosity during writing was reduced by about 50% compared with traditional products, and this product was successfully released in the Japanese domestic market in May 2017 [176]. In addition, nanocellulose also shows great potential in reinforced composite materials. For example, Professor Hiroyuki Kono from Kyoto University has developed a method to prepare a nanocellulose-reinforced resin, and its purpose is to use nanocellulose to reinforce resin materials and use them in automobiles. Since nanocellulose is a light-weight material, its addition could reduce the weight of the car and thus reduce fuel consumption, further promoting the emission reduction of carbon dioxide [177].
The industrialization of nanocellulose also started quite early in Nordic countries such as Sweden and Finland. For example, INNVENTIA is dedicated to promoting the commercialization of CNF and introduced the nanocellulose into mobile phones in February 2011. A pilot-scale production line with a production capacity of 100 kg per day CNF was also manufactured by INNVENTIA [178]. StoraEnso in Imatra and UPM in Lappeenranta are two major Finnish companies who are engaged in the research of microfibrillated cellulose (MFC). In addition, VTT and Aalto University have developed continuous film preparation based on CNF-based plastic materials for food packaging [179].
Although the abovementioned companies have successfully industrialized the nanocellulose production to some extent, these production lines are still mainly based on the sulfuric acid hydrolysis method, TEMPO oxidation method, and mechanical method, and there are still many persistent issues, such as the difficult recovery of inorganic acid, large water consumption, expensive catalyst, or high energy consumption. Recent studies proposed that the AVAP process is sustainable and the chemicals can be recovered; however, due to the use of sulfur dioxide, the entire system needs very high airtightness [180]. Therefore, green, efficient, and sustainable methods for preparing nanocellulose on an industrial scale still require further efforts by researchers.

7. Conclusions and Outlook

With the global shortage of petrochemical resources, climate warming, and environmental pollution, people pay more and more attention to how to reduce energy consumption, reasonably allocate nonrenewable resources, expand the use of renewable resources, and take the “green” road in line with the “concept of ecological civilization in the new era”. Nanocellulose and its derived nanocomposite have become a research hotspot in food packaging because of its many excellent properties, including high strength, large specific surface area, excellent barrier property, and good biocompatibility, safety, nontoxicity, and degradability. In food packaging materials, nanocellulose-based nanocomposites can be used as fresh-keeping and antibacterial packaging materials, smart packaging materials, and high-barrier packaging materials, showing the high application potential of nanocellulose-based composites. Therefore, as a kind of renewable and environmentally friendly packaging material, nanocellulose-based composites improve the safety and quality of food and are one of the important directions to realize the development of an “environment-friendly industry” in the food industry in the future.
This review describes the inter-relation of cellulose chemical structure and its source, and their different physicochemical properties are discussed. Though cellulose extracted from plants is the most investigated, the cellulose purified from bacteria and animals with unique structural characteristics has raised increasing interest. To date, enzymatic hydrolysis, TEMPO-oxidation, and carboxymethylation are widely utilized pretreatments to help defibrillation of cellulose during processing into nanocellulose, which not only helps reduce the energy consumption but also provides extra functional groups for the final products. Acid hydrolysis, including both mineral and organic acids, could remove amorphous regions, resulting in cellulose nanocrystal (CNC), though the highly corrosive conditions and the low yield of CNC are persistent issues. Through mechanical treatments, such as refining, homogenization, microfluidization, sonification, ball milling, and the aqueous counter collision (ACC) method, cellulose nanofibrils (CNF) could be produced, but the extremely high energy input prohibits the commercialization of these techniques. Bacterial cellulose (BC), a unique bacteria-derived nanocellulose, has recently gained numerous research interests. Its higher aspect ratio and larger diameter make it a promising material for food packaging applications. In order to facilitate the application of nanocellulose in food packaging, it has always been processed to different forms of materials, such as film, gel, coating membrane, and emulsions, by various fabrication technologies, including solution casting, Layer-by-Layer (LBL) assembly, extrusion, coating, gel-forming, spray drying, electrostatic spinning, adsorption, etc. Thanks to the nontoxicity, good biodegradability and biocompatibility, high aspect ratio, low thermal expansion coefficient, excellent mechanical strength, and unique optical properties, nanocellulose-based food packaging materials have been widely applied to pack fruits, meat products, instant foods, dairy products, and beverages. Since nanocellulose and the functional fillers incorporated into the cellulose-based nanocomposites impart the materials’ excellent barrier and mechanical properties, antibacterial activity, and stimuli-responsive performance, they have greatly improved the quality stability and shelf life of foods.
So as to provide sufficient nanocellulose products for green food packaging, many companies in Europe, Africa, and Asia have been pushing the lab-scale production of nanocellulose for a pilot- or even industrial-scale production. However, there are still several persistent issues that should be overcome in the near future to realize the successful commercialization of nanocellulose-based composites as food packaging materials.
  • Expansion of the cellulose sources to other biomass besides the traditional raw materials, such as tunicate and BC, for higher quality nanocellulose to develop many advanced applications. Until now, the nanocellulose on the market is dominantly produced from wood and other plant-based sources. Though several works of research focused on the nanocellulose preparation from tunicate and BC and demonstrated their better performance than the woody nanocellulose, further investigation on the preparation–characterization–performance correlation of this specific nanocellulose is necessary.
  • Development of facial, cost-effective, efficient, and environment-friendly nanocellulose extraction method. Though several novel extraction methods have been developed, sulfuric acid hydrolysis and mechanical refining are still the most widely used ones. However, the harsh acid hydrolysis, high water consumption, huge amount of polluted wastewater, intense energy consumption, and low yield greatly prohibited industrially feasible nanocellulose production. Therefore, more efforts should be put into developing new nanocellulose preparation methods, such as organic acid-based methods, which already showed potential to be a green approach to preparing functionalized nanocellulose.
  • Development of new cellulose nanocomposite fabrication approaches. In the lab, the solution casting method is still widely used for research purposes, which is unsuitable for industrial production. In the pilot scale, extrusion is used, though it is not a perfect method since nanocellulose is always dispersed in water, which negatively affects the extrusion performance. Therefore, developing a scalable strategy to prepare nanocellulose-based composites for food packaging materials is vital.
  • Improvement of the performance of nanocellulose-based composites as packaging materials. As discussed above, the ideal food packaging materials require UV-proof, gas and vapor barrier properties, excellent mechanical force, and good hydrophobicity. Especially for the last one, new strategies need to be developed to alter the hygroscopic nature of nanocellulose and enhance the wet strength, thus making its applications more practical in daily life. For example, esterification as a pretreatment or coating with natural wax seems suitable to fulfil this purpose.
  • Development of nanocellulose-based intelligent packaging materials. Currently, achieving the cellulose nanocomposites’ responsive properties is mainly realized by incorporating various organic and inorganic fillers. However, the release and migration of functional fillers and their potential health risks have not been comprehensively evaluated. Future studies should not only focus on the safety issue of the nanocellulose itself but also on the functional fillers used.
  • Design of the food-specific, nanocellulose-based packaging materials. Though many research works generally focused on the preparation and properties of packaging materials, they paid little attention to the interaction between the materials and the food, and even ignored various aspects which would influence the application of the materials. For example, the influence of the environmental conditions on the quality change of both food products and the packaging materials should be investigated to prove the feasibility and suitability of the packaging materials for the specific food.

Author Contributions

Conceptualization and design: Y.Z. (Yadong Zhao) and X.Z.; writing—original draft preparation, J.L., F.Z., Y.Z. (Yaqi Zhong) and Y.Z. (Yadong Zhao); writing—review and editing, P.G. and R.Z.; supervision, Y.Z. (Yadong Zhao), F.T., R.Z. and P.J.C.; funding acquisition, Y.Z. (Yadong Zhao) and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

Yadong Zhao is grateful for the financial support from the Zhejiang Ocean University (11135091221), Science and Technology Planning Project of Zhoushan of China (No. 2022C41001). Xianhui Zhang thanks the financial support from National Natural Science Foundation of China (Grant No. 51877184). Fang Tian is financially supported by Yuyao Agricultural, Rural, and Social Development Science and Technology Project: Research and Demonstration Application of Green Processing Technology of Kiwifruit (20201YYS030002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Yadong Zhao is grateful for the financial support from the Zhejiang Ocean University (11135091221), Science and Technology Planning Project of Zhoushan of China (No. 2022C41001). Xianhui Zhang thanks the financial support from National Natural Science Foundation of China (Grant No. 51877184).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Han, J.W.; Ruiz-Garcia, L.; Qian, J.P.; Yang, X.T. Food packaging: A comprehensive review and future trends. Compr. Rev. Food Sci. Food Saf. 2018, 17, 860–877. [Google Scholar] [CrossRef] [PubMed]
  2. Alojaly, H.; Benyounis, K.Y. Packaging With Plastics and Polymeric Materials. In Encyclopedia of Materials: Plastics and Polymers; Hashmi, M.S.J., Ed.; Elsevier: Oxford, UK, 2022; pp. 485–501. [Google Scholar]
  3. Rovera, C.; Türe, H.; Hedenqvist, M.S.; Farris, S. Water vapor barrier properties of wheat gluten/silica hybrid coatings on paperboard for food packaging applications. Food Packag. Shelf Life 2020, 26, 100561. [Google Scholar] [CrossRef]
  4. Foraboschi, P. Analytical modeling to predict thermal shock failure and maximum temperature gradients of a glass panel. Mater. Des. 2017, 134, 301–319. [Google Scholar] [CrossRef]
  5. Videira-Quintela, D.; Guillén, F.; Martin, O.; Montalvo, G. Antibacterial LDPE films for food packaging application filled with metal-fumed silica dual-side fillers. Food Packag. Shelf Life 2021, 31, 100772. [Google Scholar] [CrossRef]
  6. Suleman, R.; Amjad, A.; Ismail, A.; Javed, S.; Ghafoor, U.; Fahad, S. Impact of plastic bags usage in food commodities: An irreversible loss to environment. Environ. Sci. Pollut. Res. 2022, 29, 49483–49489. [Google Scholar] [CrossRef] [PubMed]
  7. Tedjani, C.F.; Ben Mya, O.; Rebiai, A. Isolation and characterization of cellulose from date palm tree spathe sheath. Sustain. Chem. Pharm. 2020, 17, 100307. [Google Scholar] [CrossRef]
  8. Gabriel, T.; Belete, A.; Syrowatka, F.; Neubert, R.H.; Gebre-Mariam, T. Extraction and characterization of celluloses from various plant byproducts. Int. J. Biol. Macromol. 2020, 158, 1248–1258. [Google Scholar] [CrossRef]
  9. Chanthathamrongsiri, N.; Petchsomrit, A.; Leelakanok, N.; Siranonthana, N.; Sirirak, T. The comparison of the properties of nanocellulose isolated from colonial and solitary marine tunicates. Heliyon 2021, 7, e07819. [Google Scholar] [CrossRef]
  10. Han, J.S.; Kim, S.Y.; Seo, Y.B. Disk-shaped cellulose fibers from red algae, Eucheuma cottonii and its use for high oxygen barrier. Int. J. Biol. Macromol. 2022, 210, 752–758. [Google Scholar] [CrossRef]
  11. Rastogi, A.; Sahoo, S.; Bandyopadhyay, T.K.; Mukherjee, R.; Banerjee, R. Detailed morphological and kinetic studies of cellulose biosynthesis from Leifsonia soli. Polymer 2022, 242, 124568. [Google Scholar] [CrossRef]
  12. Jonoobi, M.; Oladi, R.; Davoudpour, Y.; Oksman, K.; Dufresne, A.; Hamzeh, Y.; Davoodi, R. Different preparation methods and properties of nanostructured cellulose from various natural resources and residues: A review. Cellulose 2015, 22, 935–969. [Google Scholar] [CrossRef]
  13. Liu, Y.; Ahmed, S.; Sameen, D.E.; Wang, Y.; Lu, R.; Dai, J.; Li, S.; Qin, W. A review of cellulose and its derivatives in biopolymer-based for food packaging application. Trends Food Sci. Technol. 2021, 112, 532–546. [Google Scholar] [CrossRef]
  14. Spiliopoulos, P.; Spirk, S.; Pääkkönen, T.; Viljanen, M.; Svedström, K.; Pitkänen, L.; Awais, M.; Kontturi, E. Visualizing Degradation of Cellulose Nanofibers by Acid Hydrolysis. Biomacromolecules 2021, 22, 1399–1405. [Google Scholar] [CrossRef] [PubMed]
  15. Serial, M.; Velichko, E.; Nikolaeva, T.; Adel, R.D.; Terenzi, C.; Bouwman, W.; van Duynhoven, J. High-pressure homogenized citrus fiber cellulose dispersions: Structural characterization and flow behavior. Food Struct. 2021, 30, 100237. [Google Scholar] [CrossRef]
  16. Shamshina, J.L.; Acharya, S.; Rumi, S.S.; Liyanage, S.; Parajuli, P.; Abidi, N. Cryogenic grinding of cotton fiber cellulose: The effect on physicochemical properties. Carbohydr. Polym. 2022, 289, 119408. [Google Scholar] [CrossRef]
  17. Qian, M.; Lei, H.; Villota, E.; Zhao, Y.; Wang, C.; Huo, E.; Zhang, Q.; Mateo, W.; Lin, X. High yield production of nanocrystalline cellulose by microwave-assisted dilute-acid pretreatment combined with enzymatic hydrolysis. Chem. Eng. Process. Process Intensif. 2020, 160, 108292. [Google Scholar] [CrossRef]
  18. Phanthong, P.; Reubroycharoen, P.; Hao, X.; Xu, G.; Abudula, A.; Guan, G. Nanocellulose: Extraction and application. Carbon Resour. Convers. 2018, 1, 32–43. [Google Scholar] [CrossRef]
  19. Blanco, A.; Monte, M.C.; Campano, C.; Balea, A.; Merayo, N.; Negro, C. Nanocellulose for industrial use: Cellulose nanofibers (CNF), cellulose nanocrystals (CNC), and bacterial cellulose (BC). In Handbook of Nanomaterials for Industrial Applications; Elsevier: Amsterdam, The Netherlands, 2018; pp. 74–126. [Google Scholar]
  20. Vanderfleet, O.M.; Cranston, E.D. Production routes to tailor the performance of cellulose nanocrystals. Nat. Rev. Mater. 2020, 6, 124–144. [Google Scholar] [CrossRef]
  21. Ceccherini, S.; Ståhl, M.; Sawada, D.; Hummel, M.; Maloney, T.C. Effect of Enzymatic Depolymerization of Cellulose and Hemicelluloses on the Direct Dissolution of Prehydrolysis Kraft Dissolving Pulp. Biomacromolecules 2021, 22, 4805–4813. [Google Scholar] [CrossRef]
  22. Casaburi, A.; Rojo, Ú.M.; Cerrutti, P.; Vázquez, A.; Foresti, M.L. Carboxymethyl cellulose with tailored degree of substitution obtained from bacterial cellulose. Food Hydrocoll. 2018, 75, 147–156. [Google Scholar] [CrossRef]
  23. Baron, R.I.; Coseri, S. Preparation of water-soluble cellulose derivatives using TEMPO radical-mediated oxidation at extended reaction time. React. Funct. Polym. 2020, 157, 104768. [Google Scholar] [CrossRef]
  24. Boufi, S.; González, I.; Delgado-Aguilar, M.; Tarrès, Q.; Pèlach, M.; Mutjé, P. Nanofibrillated cellulose as an additive in papermaking process: A review. Carbohydr. Polym. 2016, 154, 151–166. [Google Scholar] [CrossRef] [PubMed]
  25. Parte, F.G.B.; Santoso, S.P.; Chou, C.-C.; Verma, V.; Wang, H.-T.; Ismadji, S.; Cheng, K.-C. Current progress on the production, modification, and applications of bacterial cellulose. Crit. Rev. Biotechnol. 2020, 40, 397–414. [Google Scholar] [CrossRef] [PubMed]
  26. Moon, R.J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J. Cellulose nanomaterials review: Structure, properties and nanocomposites. Chem. Soc. Rev. 2011, 40, 3941–3994. [Google Scholar] [CrossRef] [PubMed]
  27. Heinze, T. Cellulose: Structure and properties. In Cellulose Chemistry and Properties: Fibers, Nanocelluloses and Advanced Materials; Springer: Cham, Switzerland, 2015; pp. 1–52. [Google Scholar]
  28. Mazeau, K.; Heux, L. Molecular Dynamics Simulations of Bulk Native Crystalline and Amorphous Structures of Cellulose. J. Phys. Chem. B 2003, 107, 2394–2403. [Google Scholar] [CrossRef]
  29. Nam, S.; French, A.D.; Condon, B.D.; Concha, M. Segal crystallinity index revisited by the simulation of X-ray diffraction patterns of cotton cellulose Iβ and cellulose II. Carbohydr. Polym. 2016, 135, 1–9. [Google Scholar] [CrossRef]
  30. Jarvis, M. Cellulose stacks up. Nature 2003, 426, 611–612. [Google Scholar] [CrossRef]
  31. Islam, M.S.; Chen, L.; Sisler, J.; Tam, K.C. Cellulose nanocrystal (CNC)-inorganic hybrid systems: Synthesis, properties and applications. J. Mater. Chem. B 2018, 6, 864–883. [Google Scholar] [CrossRef]
  32. Unbehaun, H.; Dittler, B.; Kühne, G.; Wagenführ, A. Investigation into the biotechnological modification of wood and its application in the wood-based material industry. Acta Biotechnol. 2000, 20, 305–312. [Google Scholar] [CrossRef]
  33. Nechyporchuk, O.; Belgacem, M.N.; Bras, J. Production of cellulose nanofibrils: A review of recent advances. Ind. Crop. Prod. 2016, 93, 2–25. [Google Scholar] [CrossRef]
  34. Kalia, S.; Kaith, B.; Kaur, I. Pretreatments of natural fibers and their application as reinforcing material in polymer composites-A review. Polym. Eng. Sci. 2009, 49, 1253–1272. [Google Scholar] [CrossRef]
  35. Rajala, S.; Siponkoski, T.; Sarlin, E.; Mettänen, M.; Vuoriluoto, M.; Pammo, A.; Juuti, J.; Rojas, O.J.; Franssila, S.; Tuukkanen, S. Cellulose Nanofibril Film as a Piezoelectric Sensor Material. ACS Appl. Mater. Interfaces 2016, 8, 15607–15614. [Google Scholar] [CrossRef] [PubMed]
  36. Brown, R.M.; Saxena, I.M. Cellulose: Molecular and Structural Biology; Springer: Dordrecht, The Netherlands, 2007. [Google Scholar] [CrossRef]
  37. Azeredo, H.M.C.; Barud, H.; Farinas, C.S.; Vasconcellos, V.M.; Claro, A.M. Bacterial Cellulose as a Raw Material for Food and Food Packaging Applications. Front. Sustain. Food Syst. 2019, 3, 106530. [Google Scholar] [CrossRef]
  38. Tsekos, I. The sites of cellulose synthesis in algae: Diversity and evolution of cellulose-synthesizing enzyme complexes. J. Phycol. 1999, 35, 635–655. [Google Scholar] [CrossRef]
  39. Hu, S.-Q.; Gao, Y.-G.; Tajima, K.; Sunagawa, N.; Zhou, Y.; Kawano, S.; Fujiwara, T.; Yoda, T.; Shimura, D.; Satoh, Y.; et al. Structure of bacterial cellulose synthase subunit D octamer with four inner passageways. Proc. Natl. Acad. Sci. USA 2010, 107, 17957–17961. [Google Scholar] [CrossRef]
  40. Sani, A.; Dahman, Y. Improvements in the production of bacterial synthesized biocellulose nanofibres using different culture methods. J. Chem. Technol. Biotechnol. 2009, 85, 151–164. [Google Scholar] [CrossRef]
  41. Mohammadkazemi, F.; Azin, M.; Ashori, A. Production of bacterial cellulose using different carbon sources and culture media. Carbohydr. Polym. 2015, 117, 518–523. [Google Scholar] [CrossRef]
  42. Cheng, Z.; Yang, R.; Liu, X.; Liu, X.; Chen, H. Green synthesis of bacterial cellulose via acetic acid pre-hydrolysis liquor of agricultural corn stalk used as carbon source. Bioresour. Technol. 2017, 234, 8–14. [Google Scholar] [CrossRef]
  43. Zhang, W.; Zhang, Y.; Lu, C.; Deng, Y. Aerogels from crosslinked cellulose nano/micro-fibrils and their fast shape recovery property in water. J. Mater. Chem. 2012, 22, 11642–11650. [Google Scholar] [CrossRef]
  44. Morais, J.P.S.; Rosa, M.F.; Filho, M.M.S.; Nascimento, L.D.; Nascimento, D.M.; Cassales, A.R. Extraction and characterization of nanocellulose structures from raw cotton linter. Carbohydr. Polym. 2013, 91, 229–235. [Google Scholar] [CrossRef]
  45. Zhao, Y.; Zhang, Y.; Lindström, M.E.; Li, J. Tunicate cellulose nanocrystals: Preparation, neat films and nanocomposite films with glucomannans. Carbohydr. Polym. 2015, 117, 286–296. [Google Scholar] [CrossRef] [PubMed]
  46. Xiang, Z.; Gao, W.; Chen, L.; Lan, W.; Zhu, J.Y.; Runge, T. A comparison of cellulose nanofibrils produced from Cladophora glomerata algae and bleached eucalyptus pulp. Cellulose 2016, 23, 493–503. [Google Scholar] [CrossRef]
  47. Das, A.A.K.; Bovill, J.; Ayesh, M.; Stoyanov, S.D.; Paunov, V.N. Fabrication of living soft matter by symbiotic growth of unicellular microorganisms. J. Mater. Chem. B 2016, 4, 3685–3694. [Google Scholar] [CrossRef] [PubMed]
  48. Ajdary, R.; Abidnejad, R.; Lehtonen, J.; Kuula, J.; Raussi-Lehto, E.; Kankuri, E.; Tardy, B.; Rojas, O.J. Bacterial nanocellulose enables auxetic supporting implants. Carbohydr. Polym. 2022, 284, 119198. [Google Scholar] [CrossRef] [PubMed]
  49. Zhao, Y.; Li, J. Unique and outstanding quantum dots (QD)/tunicate cellulose nanofibrils (TCNF) nanohybrid platform material for use as 1D ink and 2D film. Carbohydr. Polym. 2020, 242, 116396. [Google Scholar] [CrossRef]
  50. Kassab, Z.; Ben Youcef, H.; Hannache, H.; El Achaby, M. Isolation of Cellulose Nanocrystals from Various Lignocellulosic Materials: Physico-chemical characterization and Application in Polymer Composites Development. Mater. Today Proc. 2019, 13, 964–973. [Google Scholar] [CrossRef]
  51. Li, J.; Zhang, X.; Zhao, Y.; Ma, M.; Song, Y.; Zheng, B.; Zhou, R.; Ostrikov, K. Nisin electroadsorption-enabled multifunctional bacterial cellulose membranes for highly efficient removal of organic and microbial pollutants in water. Chem. Eng. J. 2022, 440, 135922. [Google Scholar] [CrossRef]
  52. Almeida, T.; Silvestre, A.; Vilela, C.; Freire, C. Bacterial Nanocellulose toward Green Cosmetics: Recent Progresses and Challenges. Int. J. Mol. Sci. 2021, 22, 2836. [Google Scholar] [CrossRef]
  53. Noremylia, M.; Hassan, M.Z.; Ismail, Z. Recent advancement in isolation, processing, characterization and applications of emerging nanocellulose: A review. Int. J. Biol. Macromol. 2022, 206, 954–976. [Google Scholar] [CrossRef]
  54. Sharma, A.; Thakur, M.; Bhattacharya, M.; Mandal, T.; Goswami, S. Commercial application of cellulose nano-composites—A review. Biotechnol. Rep. 2019, 21, e00316. [Google Scholar] [CrossRef]
  55. Kwon, G.; Lee, K.; Kim, D.; Jeon, Y.; Kim, U.-J.; You, J. Cellulose nanocrystal-coated TEMPO-oxidized cellulose nanofiber films for high performance all-cellulose nanocomposites. J. Hazard. Mater. 2020, 398, 123100. [Google Scholar] [CrossRef] [PubMed]
  56. Shang, Z.; An, X.; Seta, F.T.; Ma, M.; Shen, M.; Dai, L.; Liu, H.; Ni, Y. Improving dispersion stability of hydrochloric acid hydrolyzed cellulose nano-crystals. Carbohydr. Polym. 2019, 222, 115037. [Google Scholar] [CrossRef] [PubMed]
  57. Holilah, H.; Bahruji, H.; Ediati, R.; Asranudin, A.; Jalil, A.A.; Piluharto, B.; Nugraha, R.E.; Prasetyoko, D. Uniform rod and spherical nanocrystalline celluloses from hydrolysis of industrial pepper waste (Piper nigrum L.) using organic acid and inorganic acid. Int. J. Biol. Macromol. 2022, 204, 593–605. [Google Scholar] [CrossRef] [PubMed]
  58. Sukmawan, R.; Kusmono; Rahmanta, A.P.; Saputri, L.H. The effect of repeated alkali pretreatments on the morphological characteristics of cellulose from oil palm empty fruit bunch fiber-reinforced epoxy adhesive composite. Int. J. Adhes. Adhes. 2022, 114, 103095. [Google Scholar] [CrossRef]
  59. Kaffashsaie, E.; Yousefi, H.; Nishino, T.; Matsumoto, T.; Mashkour, M.; Madhoushi, M.; Kawaguchi, H. Direct conversion of raw wood to TEMPO-oxidized cellulose nanofibers. Carbohydr. Polym. 2021, 262, 117938. [Google Scholar] [CrossRef]
  60. Yu, W.; Yi, Y.; Wang, H.; Yang, Y.; Zeng, L.; Tan, Z. Light-colored cellulose nanofibrils produced from raw sisal fibers without costly bleaching. Ind. Crop. Prod. 2021, 172, 114009. [Google Scholar] [CrossRef]
  61. Dias, M.C.; Belgacem, M.N.; de Resende, J.V.; Martins, M.A.; Damásio, R.A.P.; Tonoli, G.H.D.; Ferreira, S.R. Eco-friendly laccase and cellulase enzymes pretreatment for optimized production of high content lignin-cellulose nanofibrils. Int. J. Biol. Macromol. 2022, 209, 413–425. [Google Scholar] [CrossRef]
  62. Chitbanyong, K.; Pisutpiched, S.; Khantayanuwong, S.; Theeragool, G.; Puangsin, B. TEMPO-oxidized cellulose nanofibril film from nano-structured bacterial cellulose derived from the recently developed thermotolerant Komagataeibacter xylinus C30 and Komagataeibacter oboediens R37–9 strains. Int. J. Biol. Macromol. 2020, 163, 1908–1914. [Google Scholar] [CrossRef]
  63. Xu, Y.; Yang, S.; Zhao, P.; Wu, M.; Song, X.; Ragauskas, A.J. Effect of endoglucanase and high-pressure homogenization post-treatments on mechanically grinded cellulose nanofibrils and their film performance. Carbohydr. Polym. 2020, 253, 117253. [Google Scholar] [CrossRef]
  64. Song, H.Y.; Park, S.Y.; Kim, S.; Youn, H.J.; Hyun, K. Linear and nonlinear oscillatory rheology of chemically pretreated and non-pretreated cellulose nanofiber suspensions. Carbohydr. Polym. 2021, 275, 118765. [Google Scholar] [CrossRef]
  65. Saito, T.; Isogai, A. TEMPO-Mediated Oxidation of Native Cellulose. The Effect of Oxidation Conditions on Chemical and Crystal Structures of the Water-Insoluble Fractions. Biomacromolecules 2004, 5, 1983–1989. [Google Scholar] [CrossRef]
  66. Zheng, D.; Zhang, Y.; Guo, Y.; Yue, J. Isolation and Characterization of Nanocellulose with a Novel Shape from Walnut (Juglans regia L.) Shell Agricultural Waste. Polymers 2019, 11, 1130. [Google Scholar] [CrossRef] [PubMed]
  67. Shen, X.-J.; Huang, P.-L.; Chen, J.-H.; Wu, Y.-Y.; Liu, Q.-Y.; Sun, R.-C. Comparison of acid-hydrolyzed and TEMPO-oxidized nanocellulose for reinforcing alginate fibers. BioResources 2017, 12, 8180–8198. [Google Scholar] [CrossRef]
  68. Xu, R.; Du, H.; Wang, H.; Zhang, M.; Wu, M.; Liu, C.; Yu, G.; Zhang, X.; Si, C.; Choi, S.-E.; et al. Valorization of Enzymatic Hydrolysis Residues from Corncob into Lignin-Containing Cellulose Nanofibrils and Lignin Nanoparticles. Front. Bioeng. Biotechnol. 2021, 9, 677963. [Google Scholar] [CrossRef] [PubMed]
  69. Guan, Q.; Chen, J.; Chen, D.; Chai, X.; He, L.; Peng, L.; Zhang, J.; Li, J. A new sight on the catalytic oxidation kinetic behaviors of bamboo cellulose fibers under TEMPO-oxidized system: The fate of carboxyl groups in treated pulps. J. Catal. 2019, 370, 304–309. [Google Scholar] [CrossRef]
  70. Dufresne, A. Nanocellulose: A new ageless bionanomaterial. Mater. Today 2013, 16, 220–227. [Google Scholar] [CrossRef]
  71. Zhou, Z.; Xia, K.; Liu, T.; Guo, H.; Liu, X.; Zhang, X. Preparation of carboxymethyl cellulose nanofibers and their application in warp size of textile. Int. J. Biol. Macromol. 2022, 207, 40–47. [Google Scholar] [CrossRef]
  72. Phirom-On, K.; Apiraksakorn, J. Development of cellulose-based prebiotic fiber from banana peel by enzymatic hydrolysis. Food Biosci. 2021, 41, 101083. [Google Scholar] [CrossRef]
  73. Xie, L.; Shen, M.; Wang, Z.; Xie, J. Structure, function and food applications of carboxymethylated polysaccharides: A comprehensive review. Trends Food Sci. Technol. 2021, 118, 539–557. [Google Scholar] [CrossRef]
  74. Pinto, E.; Aggrey, W.N.; Boakye, P.; Amenuvor, G.; Sokama-Neuyam, Y.A.; Fokuo, M.K.; Karimaie, H.; Sarkodie, K.; Adenutsi, C.D.; Erzuah, S.; et al. Cellulose processing from biomass and its derivatization into carboxymethylcellulose: A review. Sci. Afr. 2021, 15, e01078. [Google Scholar] [CrossRef]
  75. Heinze, T.; El Seoud, O.A.; Koschella, A. Etherification of Cellulose. In Cellulose Derivatives; Springer: Berlin/Heidelberg, Germany, 2018; pp. 429–477. [Google Scholar] [CrossRef]
  76. Toğrul, H.; Arslan, N. Production of carboxymethyl cellulose from sugar beet pulp cellulose and rheological behaviour of carboxymethyl cellulose. Carbohydr. Polym. 2003, 54, 73–82. [Google Scholar] [CrossRef]
  77. Yeasmin, M.S.; Mondal, M.I.H. Synthesis of highly substituted carboxymethyl cellulose depending on cellulose particle size. Int. J. Biol. Macromol. 2015, 80, 725–731. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, Y.; Sun, Q.; Pan, Y.; Wei, S.; Xia, Q.; Liu, S.; Ji, H.; Deng, C.; Hao, J. Investigation on the correlation between changes in water and texture properties during the processing of surimi from golden pompano (Trachinotus ovatus). J. Food Sci. 2021, 86, 376–384. [Google Scholar] [CrossRef]
  79. Dinesh, G.; Kandasubramanian, B. Fabrication of transparent paper devices from nanocellulose fiber. Mater. Chem. Phys. 2022, 281, 125707. [Google Scholar] [CrossRef]
  80. Arantes, V.; Dias, I.K.R.; Berto, G.L.; Pereira, B.; Marotti, B.S.; Nogueira, C.F.O. The current status of the enzyme-mediated isolation and functionalization of nanocelluloses: Production, properties, techno-economics, and opportunities. Cellulose 2020, 27, 10571–10630. [Google Scholar] [CrossRef]
  81. Chen, M.; Parot, J.; Hackley, V.A.; Zou, S.; Johnston, L.J. AFM characterization of cellulose nanocrystal height and width using internal calibration standards. Cellulose 2021, 28, 1933–1946. [Google Scholar] [CrossRef]
  82. Usov, I.; Nyström, G.; Adamcik, J.; Handschin, S.; Schütz, C.; Fall, A.; Bergström, L.; Mezzenga, R. Understanding nanocellulose chirality and structure–properties relationship at the single fibril level. Nat. Commun. 2015, 6, 7564. [Google Scholar] [CrossRef]
  83. Vanderfleet, O.M.; Reid, M.S.; Bras, J.; Heux, L.; Godoy-Vargas, J.; Panga, M.K.R.; Cranston, E.D. Insight into thermal stability of cellulose nanocrystals from new hydrolysis methods with acid blends. Cellulose 2018, 26, 507–528. [Google Scholar] [CrossRef]
  84. Yu, H.; Qin, Z.; Liang, B.; Liu, N.; Zhou, Z.; Chen, L. Facile extraction of thermally stable cellulose nanocrystals with a high yield of 93% through hydrochloric acid hydrolysis under hydrothermal conditions. J. Mater. Chem. A 2013, 1, 3938–3944. [Google Scholar] [CrossRef]
  85. Wang, R.; Chen, L.; Zhu, J.; Yang, R. Tailored and Integrated Production of Carboxylated Cellulose Nanocrystals (CNC) with Nanofibrils (CNF) through Maleic Acid Hydrolysis. ChemNanoMat 2017, 3, 328–335. [Google Scholar] [CrossRef]
  86. Henschen, J.; Li, D.; Ek, M. Preparation of cellulose nanomaterials via cellulose oxalates. Carbohydr. Polym. 2019, 213, 208–216. [Google Scholar] [CrossRef] [PubMed]
  87. Ono, Y.; Nakamura, Y.; Zhou, Y.; Horikawa, Y.; Isogai, A. Linear and branched structures present in high-molar-mass fractions in holocelluloses prepared from chara, haircap moss, adiantum, ginkgo, Japanese cedar, and eucalyptus. Cellulose 2021, 28, 3935–3949. [Google Scholar] [CrossRef]
  88. Song, J.; Yang, F.; Zhang, Y.; Hu, F.; Wu, S.; Jin, Y.; Guo, J.; Rojas, O.J. Interactions between fungal cellulases and films of nanofibrillar cellulose determined by a quartz crystal microbalance with dissipation monitoring (QCM-D). Cellulose 2017, 24, 1947–1956. [Google Scholar] [CrossRef]
  89. Marakana, P.G.; Dey, A.; Saini, B. Isolation of nanocellulose from lignocellulosic biomass: Synthesis, characterization, modification, and potential applications. J. Environ. Chem. Eng. 2021, 9, 106606. [Google Scholar] [CrossRef]
  90. Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: A new family of nature-based materials. Angew. Chem. Int. Ed. 2011, 50, 5438–5466. [Google Scholar] [CrossRef]
  91. Guo, Y.; Zhang, Y.; Zheng, D.; Li, M.; Yue, J. Isolation and characterization of nanocellulose crystals via acid hydrolysis from agricultural waste-tea stalk. Int. J. Biol. Macromol. 2020, 163, 927–933. [Google Scholar] [CrossRef]
  92. Huang, L.; Zhang, X.; Xu, M.; Chen, J.; Shi, Y.; Huang, C.; Wang, S.; An, S.; Li, C. Preparation and mechanical properties of modified nanocellulose/PLA composites from cassava residue. AIP Adv. 2018, 8, 025116. [Google Scholar] [CrossRef]
  93. Lin, K.-H.; Enomae, T.; Chang, F.-C. Cellulose Nanocrystal Isolation from Hardwood Pulp using Various Hydrolysis Conditions. Molecules 2019, 24, 3724. [Google Scholar] [CrossRef] [Green Version]
  94. Hastuti, N.; Kanomata, K.; Kitaoka, T. Hydrochloric Acid Hydrolysis of Pulps from Oil Palm Empty Fruit Bunches to Produce Cellulose Nanocrystals. J. Polym. Environ. 2018, 26, 3698–3709. [Google Scholar] [CrossRef]
  95. Xie, H.; Du, H.; Yang, X.; Si, C. Recent Strategies in Preparation of Cellulose Nanocrystals and Cellulose Nanofibrils Derived from Raw Cellulose Materials. Int. J. Polym. Sci. 2018, 2018, 7923068. [Google Scholar] [CrossRef]
  96. Le Gars, M.; Douard, L.; Belgacem, N.; Bras, J. Cellulose Nanocrystals: From Classical Hydrolysis to the Use of Deep Eutectic Solvents. In Smart Nanosystems for Biomedicine, Optoelectronics and Catalysis; IntechOpen: London, UK, 2020. [Google Scholar]
  97. Septevani, A.A.; Rifathin, A.; Sari, A.A.; Sampora, Y.; Ariani, G.N.; Sudiyarmanto; Sondari, D. Oil palm empty fruit bunch-based nanocellulose as a super-adsorbent for water remediation. Carbohydr. Polym. 2019, 229, 115433. [Google Scholar] [CrossRef] [PubMed]
  98. Almashhadani, A.Q.; Leh, C.P.; Chan, S.-Y.; Lee, C.Y.; Goh, C.F. Nanocrystalline cellulose isolation via acid hydrolysis from non-woody biomass: Importance of hydrolysis parameters. Carbohydr. Polym. 2022, 286, 119285. [Google Scholar] [CrossRef] [PubMed]
  99. Jiang, J.; Zhu, Y.; Jiang, F. Sustainable isolation of nanocellulose from cellulose and lignocellulosic feedstocks: Recent progress and perspectives. Carbohydr. Polym. 2021, 267, 118188. [Google Scholar] [CrossRef] [PubMed]
  100. Chiulan, I.; Panaitescu, D.M.; Radu, E.-R.; Vizireanu, S.; Sătulu, V.; Biţă, B.; Gabor, R.A.; Nicolae, C.A.; Raduly, M.; Rădiţoiu, V. Influence of TEMPO oxidation on the properties of ethylene glycol methyl ether acrylate grafted cellulose sponges. Carbohydr. Polym. 2021, 272, 118458. [Google Scholar] [CrossRef]
  101. Chen, C.; Ding, W.; Zhang, H.; Zhang, L.; Huang, Y.; Fan, M.; Yang, J.; Sun, D. Bacterial cellulose-based biomaterials: From fabrication to application. Carbohydr. Polym. 2021, 278, 118995. [Google Scholar] [CrossRef]
  102. Wang, H.; Du, H.; Liu, K.; Liu, H.; Xu, T.; Zhang, S.; Chen, X.; Zhang, R.; Li, H.; Xie, H.; et al. Sustainable preparation of bifunctional cellulose nanocrystals via mixed H2SO4/formic acid hydrolysis. Carbohydr. Polym. 2021, 266, 118107. [Google Scholar] [CrossRef]
  103. Pradhan, D.; Jaiswal, A.K.; Jaiswal, S. Emerging technologies for the production of nanocellulose from lignocellulosic biomass. Carbohydr. Polym. 2022, 285, 119258. [Google Scholar] [CrossRef]
  104. Teo, H.L.; Wahab, R.A. Towards an eco-friendly deconstruction of agro-industrial biomass and preparation of renewable cellulose nanomaterials: A review. Int. J. Biol. Macromol. 2020, 161, 1414–1430. [Google Scholar] [CrossRef]
  105. Nasir, M.; Hashim, R.; Sulaiman, O.; Asim, M. Nanocellulose Preparation methods and applications. In Cellulose-Reinforced Nanofibre Composites; WP Woodhead Publishing: Sawston, UK, 2017. [Google Scholar]
  106. Naderi, A.; Lindström, T.; Sundström, J. Carboxymethylated nanofibrillated cellulose: Rheological studies. Cellulose 2014, 21, 1561–1571. [Google Scholar] [CrossRef]
  107. Kondo, T.; Kose, R.; Naito, H.; Kasai, W. Aqueous counter collision using paired water jets as a novel means of preparing bio-nanofibers. Carbohydr. Polym. 2014, 112, 284–290. [Google Scholar] [CrossRef]
  108. Yokota, S.; Tagawa, S.; Kondo, T. Facile surface modification of amphiphilic cellulose nanofibrils prepared by aqueous counter collision. Carbohydr. Polym. 2020, 255, 117342. [Google Scholar] [CrossRef]
  109. Ishida, K.; Yokota, S.; Kondo, T. Localized surface acetylation of aqueous counter collision cellulose nanofibrils using a Pickering emulsion as an interfacial reaction platform. Carbohydr. Polym. 2021, 261, 117845. [Google Scholar] [CrossRef] [PubMed]
  110. Huang, W.-M.; Chen, J.-H.; Nagarajan, D.; Lee, C.-K.; Varjani, S.; Lee, D.-J.; Chang, J.-S. Immobilization of Chlorella sorokiniana AK-1 in bacterial cellulose by co-culture and its application in wastewater treatment. J. Taiwan Inst. Chem. Eng. 2022; in press. [Google Scholar] [CrossRef]
  111. Nascimento, E.S.; Barros, M.O.; Cerqueira, M.A.; Lima, H.L.; Borges, M.D.F.; Pastrana, L.M.; Gama, F.M.; Rosa, M.F.; Azeredo, H.M.; Gonçalves, C. All-cellulose nanocomposite films based on bacterial cellulose nanofibrils and nanocrystals. Food Packag. Shelf Life 2021, 29, 100715. [Google Scholar] [CrossRef]
  112. Nam, J.; Hyun, Y.; Oh, S.; Park, J.; Jin, H.-J.; Kwak, H.W. Effect of cross-linkable bacterial cellulose nanocrystals on the physicochemical properties of silk sericin films. Polym. Test. 2021, 97, 107161. [Google Scholar] [CrossRef]
  113. Vedove, T.M.; Maniglia, B.C.; Tadini, C.C. Production of sustainable smart packaging based on cassava starch and anthocyanin by an extrusion process. J. Food Eng. 2020, 289, 110274. [Google Scholar] [CrossRef]
  114. Lin, S.-P.; Huang, S.-H.; Ting, Y.; Hsu, H.-Y.; Cheng, K.-C. Evaluation of detoxified sugarcane bagasse hydrolysate by atmospheric cold plasma for bacterial cellulose production. Int. J. Biol. Macromol. 2022, 204, 136–143. [Google Scholar] [CrossRef]
  115. Cazón, P.; Vázquez, M. Bacterial cellulose as a biodegradable food packaging material: A review. Food Hydrocoll. 2020, 113, 106530. [Google Scholar] [CrossRef]
  116. Yadav, M.; Behera, K.; Chang, Y.-H.; Chiu, F.-C. Cellulose Nanocrystal Reinforced Chitosan Based UV Barrier Composite Films for Sustainable Packaging. Polymers 2020, 12, 202. [Google Scholar] [CrossRef]
  117. Shen, S.; Wang, N.; Jia, J.; Song, D.; Zuo, T.; Liu, K.; Che, Q. Constructing the basal nanofibers suit of layer-by-layer self-assembly membranes as anion exchange membranes. J. Mol. Liq. 2022, 350, 118536. [Google Scholar] [CrossRef]
  118. Dhar, P.; Gaur, S.S.; Soundararajan, N.; Gupta, A.; Bhasney, S.M.; Milli, M.; Kumar, A.; Katiyar, V. Reactive Extrusion of Polylactic Acid/Cellulose Nanocrystal Films for Food Packaging Applications: Influence of Filler Type on Thermomechanical, Rheological, and Barrier Properties. Ind. Eng. Chem. Res. 2017, 56, 4718–4735. [Google Scholar] [CrossRef]
  119. Wang, W.; Du, G.; Li, C.; Zhang, H.; Long, Y.; Ni, Y. Preparation of cellulose nanocrystals from asparagus (Asparagus officinalis L.) and their applications to palm oil/water Pickering emulsion. Carbohydr. Polym. 2016, 151, 1–8. [Google Scholar] [CrossRef] [PubMed]
  120. Zhang, X.; Liu, Y.; Wang, Y.; Luo, X.; Li, Y.; Li, B.; Wang, J.; Liu, S. Surface modification of cellulose nanofibrils with protein nanoparticles for enhancing the stabilization of O/W pickering emulsions. Food Hydrocoll. 2019, 97, 105180. [Google Scholar] [CrossRef]
  121. Niu, X.; Liu, Y.; Song, Y.; Han, J.; Pan, H. Rosin modified cellulose nanofiber as a reinforcing and co-antimicrobial agents in polylactic acid/chitosan composite film for food packaging. Carbohydr. Polym. 2018, 183, 102–109. [Google Scholar] [CrossRef]
  122. Liu, Y.; Wang, S.; Lan, W. Fabrication of antibacterial chitosan-PVA blended film using electrospray technique for food packaging applications. Int. J. Biol. Macromol. 2018, 107, 848–854. [Google Scholar] [CrossRef]
  123. Zhang, Z.; Wu, Y.; Wang, Z.; Zou, X.; Zhao, Y.; Sun, L. Fabrication of silver nanoparticles embedded into polyvinyl alcohol (Ag/PVA) composite nanofibrous films through electrospinning for antibacterial and surface-enhanced Raman scattering (SERS) activities. Mater. Sci. Eng. C 2016, 69, 462–469. [Google Scholar] [CrossRef]
  124. Padrão, J.; Gonçalves, S.; Silva, J.P.; Sencadas, V.; Lanceros-Méndez, S.; Pinheiro, A.C.; Vicente, A.A.; Rodrigues, L.R.; Dourado, F. Bacterial cellulose-lactoferrin as an antimicrobial edible packaging. Food Hydrocoll. 2016, 58, 126–140. [Google Scholar] [CrossRef]
  125. Patel, A.S.; Lakshmibalasubramaniam, S.; Nayak, B.; Tripp, C.; Kar, A.; Sappati, P.K. Improved stability of phycobiliprotein within liposome stabilized by polyethylene glycol adsorbed cellulose nanocrystals. Int. J. Biol. Macromol. 2020, 163, 209–218. [Google Scholar] [CrossRef] [PubMed]
  126. Krstić, M.; Radojević, M.; Stojanović, D.; Radojević, V.; Uskoković, P.; Ibrić, S. Formulation and characterization of nanofibers and films with carvedilol prepared by electrospinning and solution casting method. Eur. J. Pharm. Sci. 2017, 101, 160–166. [Google Scholar] [CrossRef]
  127. Oldoni, F.C.; Bernardo, M.P.; Filho, J.G.O.; de Aguiar, A.C.; Moreira, F.K.; Mattoso, L.H.; Colnago, L.A.; Ferreira, M.D. Valorization of mangoes with internal breakdown through the production of edible films by continuous solution casting. LWT 2021, 145, 111339. [Google Scholar] [CrossRef]
  128. Rajeswari, A.; Christy, E.J.S.; Swathi, E.; Pius, A. Fabrication of improved cellulose acetate-based biodegradable films for food packaging applications. Environ. Chem. Ecotoxicol. 2020, 2, 107–114. [Google Scholar] [CrossRef]
  129. Koca, N.; Bayramoğlu, B. Layer-by-layer assembly of lysozyme with iota-carrageenan and gum Arabic for surface modification of food packaging materials with improved barrier properties. Colloids Surf. A Physicochem. Eng. Asp. 2022, 639, 128391. [Google Scholar] [CrossRef]
  130. Yu, Z.; Wang, W.; Kong, F.; Lin, M.; Mustapha, A. Cellulose nanofibril/silver nanoparticle composite as an active food packaging system and its toxicity to human colon cells. Int. J. Biol. Macromol. 2019, 129, 887–894. [Google Scholar] [CrossRef]
  131. Thomas, G.; Sathish Kumar, K.; Panda, S.K.; Bindu, J. Nanocellulose Incorporated Polylactic Acid Films for Chilled Preservation of Indian Anchovy (Stolephorus indicus) (van Hasselt, 1823); Society of Fisheries Technologists: Kerala, India, 2021. [Google Scholar]
  132. Karkhanis, S.S.; Stark, N.M.; Sabo, R.C.; Matuana, L.M. Potential of extrusion-blown poly(lactic acid)/cellulose nanocrystals nanocomposite films for improving the shelf-life of a dry food product. Food Packag. Shelf Life 2021, 29, 100689. [Google Scholar] [CrossRef]
  133. Karkhanis, S.S.; Stark, N.M.; Sabo, R.C.; Matuana, L.M. Water vapor and oxygen barrier properties of extrusion-blown poly(lactic acid)/cellulose nanocrystals nanocomposite films. Compos. Part A Appl. Sci. Manuf. 2018, 114, 204–211. [Google Scholar] [CrossRef]
  134. Quintana, S.E.; Llalla, O.; García-Risco, M.R.; Fornari, T. Comparison between essential oils and supercritical extracts into chitosan-based edible coatings on strawberry quality during cold storage. J. Supercrit. Fluids 2021, 171, 105198. [Google Scholar] [CrossRef]
  135. Lu, P.; Yang, Y.; Liu, R.; Liu, X.; Ma, J.; Wu, M.; Wang, S. Preparation of sugarcane bagasse nanocellulose hydrogel as a colourimetric freshness indicator for intelligent food packaging. Carbohydr. Polym. 2020, 249, 116831. [Google Scholar] [CrossRef]
  136. Moradi, M.; Tajik, H.; Almasi, H.; Forough, M.; Ezati, P. A novel pH-sensing indicator based on bacterial cellulose nanofibers and black carrot anthocyanins for monitoring fish freshness. Carbohydr. Polym. 2019, 222, 115030. [Google Scholar] [CrossRef]
  137. Pourjavaher, S.; Almasi, H.; Meshkini, S.; Pirsa, S.; Parandi, E. Development of a colorimetric pH indicator based on bacterial cellulose nanofibers and red cabbage (Brassica oleraceae) extract. Carbohydr. Polym. 2017, 156, 193–201. [Google Scholar] [CrossRef]
  138. Rojas-Lema, S.; Terol, J.; Fages, E.; Balart, R.; Quiles-Carrillo, L.; Prieto, C.; Torres-Giner, S. Microencapsulation of Copper(II) Sulfate in Ionically Cross-Linked Chitosan by Spray Drying for the Development of Irreversible Moisture Indicators in Paper Packaging. Polymers 2020, 12, 2039. [Google Scholar] [CrossRef]
  139. Pasaoglu, M.E.; Koyuncu, I. Substitution of petroleum-based polymeric materials used in the electrospinning process with nanocellulose: A review and future outlook. Chemosphere 2020, 269, 128710. [Google Scholar] [CrossRef] [PubMed]
  140. Leena, M.M.; Yoha, K.; Moses, J.; Anandharamakrishnan, C. Edible coating with resveratrol loaded electrospun zein nanofibers with enhanced bioaccessibility. Food Biosci. 2020, 36, 100669. [Google Scholar] [CrossRef]
  141. Ye, Q.; Georges, N.; Selomulya, C. Microencapsulation of active ingredients in functional foods: From research stage to commercial food products. Trends Food Sci. Technol. 2018, 78, 167–179. [Google Scholar] [CrossRef]
  142. Salvia-Trujillo, L.; Rojas-Graü, M.A.; Soliva-Fortuny, R.; Martín-Belloso, O. Use of antimicrobial nanoemulsions as edible coatings: Impact on safety and quality attributes of fresh-cut Fuji apples. Postharvest Biol. Technol. 2015, 105, 8–16. [Google Scholar] [CrossRef]
  143. Molet-Rodríguez, A.; Turmo-Ibarz, A.; Salvia-Trujillo, L.; Martín-Belloso, O. Incorporation of antimicrobial nanoemulsions into complex foods: A case study in an apple juice-based beverage. LWT 2021, 141, 110926. [Google Scholar] [CrossRef]
  144. Benini, K.C.C.D.C.; Cioffi, M.O.H.; Voorwald, H.J.C. PHBV/cellulose nanofibrils composites obtained by solution casting and electrospinning process. In Proceedings of the 3rd Brazilian Conference on Composite Materials, Gramado, Rio Grande do Sul, Brazil, 28–31 August 2016. [Google Scholar] [CrossRef]
  145. Trigui, K.; Magnin, A.; Putaux, J.-L.; Boufi, S. Twin-screw extrusion for the production of nanocellulose-PVA gels with a high solid content. Carbohydr. Polym. 2022, 286, 119308. [Google Scholar] [CrossRef]
  146. Nguyen, H.-L.; Tran, T.H.; Hao, L.T.; Jeon, H.; Koo, J.M.; Shin, G.; Hwang, D.S.; Hwang, S.Y.; Park, J.; Oh, D.X. Biorenewable, transparent, and oxygen/moisture barrier nanocellulose/nanochitin-based coating on polypropylene for food packaging applications. Carbohydr. Polym. 2021, 271, 118421. [Google Scholar] [CrossRef]
  147. Lin, Z.; Huang, R.; Wu, J.; Penkova, A.; Qi, W.; He, Z.; Su, R. Injectable self-healing nanocellulose hydrogels crosslinked by aluminum: Cellulose nanocrystals vs. cellulose nanofibrils. Chin. J. Chem. Eng. 2022 in press. [CrossRef]
  148. Yi, K.; Fu, S.; Huang, Y. Nanocellulose-based superhydrophobic coating with acid resistance and fluorescence. Prog. Org. Coat. 2022, 168, 106911. [Google Scholar] [CrossRef]
  149. Angel, N.; Li, S.; Yan, F.; Kong, L. Recent advances in electrospinning of nanofibers from bio-based carbohydrate polymers and their applications. Trends Food Sci. Technol. 2022, 120, 308–324. [Google Scholar] [CrossRef]
  150. Souza, A.G.; Ferreira, R.R.; Paula, L.C.; Mitra, S.K.; Rosa, D.S. Starch-based films enriched with nanocellulose-stabilized Pickering emulsions containing different essential oils for possible applications in food packaging. Food Packag. Shelf Life 2020, 27, 100615. [Google Scholar] [CrossRef]
  151. Zeng, H.; Hao, H.; Wang, X.; Shao, Z. Chitosan-based composite film adsorbents reinforced with nanocellulose for removal of Cu(II) ion from wastewater: Preparation, characterization, and adsorption mechanism. Int. J. Biol. Macromol. 2022, 213, 369–380. [Google Scholar] [CrossRef] [PubMed]
  152. Na, Z.; Xi, C.; Xiaoyun, X.; Siyi, P.; Lueng, W. Development situation and prospects of food lightweight packaging materia. Sci. Technol. Foood Ind. 2014, 35, 363–367. [Google Scholar]
  153. Bharimalla, A.K.; Deshmukh, S.P.; Vigneshwaran, N.; Patil, P.G.; Prasad, V. Nanocellulose Based Polymer Composites for Applications in Food Packaging:Future Prospects and Challenges. Polym. Plast. Technol. Eng. 2017, 56, 805–823. [Google Scholar] [CrossRef]
  154. Shi, H.; Wu, L.; Luo, Y.; Yu, F.; Li, H. A facile method to prepare cellulose fiber-based food packaging papers with improved mechanical strength, enhanced barrier, and antibacterial properties. Food Biosci. 2022, 48, 101729. [Google Scholar] [CrossRef]
  155. Salim, M.H.; Kassab, Z.; Abdellaoui, Y.; García-Cruz, A.; Soumare, A.; Ablouh, E.-H.; El Achaby, M. Exploration of multifunctional properties of garlic skin derived cellulose nanocrystals and extracts incorporated chitosan biocomposite films for active packaging application. Int. J. Biol. Macromol. 2022, 210, 639–653. [Google Scholar] [CrossRef]
  156. Thongsrikhem, N.; Taokaew, S.; Sriariyanun, M.; Kirdponpattara, S. Antibacterial activity in gelatin-bacterial cellulose composite film by thermally crosslinking with cinnamaldehyde towards food packaging application. Food Packag. Shelf Life 2021, 31, 100766. [Google Scholar] [CrossRef]
  157. He, Y.; Li, H.; Fei, X.; Peng, L. Carboxymethyl cellulose/cellulose nanocrystals immobilized silver nanoparticles as an effective coating to improve barrier and antibacterial properties of paper for food packaging applications. Carbohydr. Polym. 2020, 252, 117156. [Google Scholar] [CrossRef]
  158. Dai, Q.; Huang, X.; Jia, R.; Fang, Y.; Qin, Z. Development of antibacterial film based on alginate fiber, and peanut red skin extract for food packaging. J. Food Eng. 2022, 330, 111106. [Google Scholar] [CrossRef]
  159. Mohammadalinejhad, S.; Almasi, H.; Moradi, M. Immobilization of Echium amoenum anthocyanins into bacterial cellulose film: A novel colorimetric pH indicator for freshness/spoilage monitoring of shrimp. Food Control 2020, 113, 107169. [Google Scholar] [CrossRef]
  160. Zhang, D.-Y.; Yang, J.-X.; Liu, E.-J.; Hu, R.-Z.; Yao, X.-H.; Chen, T.; Zhao, W.-G.; Liu, L.; Fu, Y.-J. Soft and elastic silver nanoparticle-cellulose sponge as fresh-keeping packaging to protect strawberries from physical damage and microbial invasion. Int. J. Biol. Macromol. 2022, 211, 470–480. [Google Scholar] [CrossRef]
  161. Thivya, P.; Bhosale, Y.; Anandakumar, S.; Hema, V.; Sinija, V. Development of active packaging film from sodium alginate/carboxymethyl cellulose containing shallot waste extracts for anti-browning of fresh-cut produce. Int. J. Biol. Macromol. 2021, 188, 790–799. [Google Scholar] [CrossRef] [PubMed]
  162. Jiang, L.; Wang, F.; Xie, X.; Xie, C.; Li, A.; Xia, N.; Gong, X.; Zhang, H. Development and characterization of chitosan/guar gum active packaging containing walnut green husk extract and its application on fresh-cut apple preservation. Int. J. Biol. Macromol. 2022, 209, 1307–1318. [Google Scholar] [CrossRef] [PubMed]
  163. Rojas-Lema, S.; Nilsson, K.; Trifol, J.; Langton, M.; Gomez-Caturla, J.; Balart, R.; Garcia-Garcia, D.; Moriana, R. Faba bean protein films reinforced with cellulose nanocrystals as edible food packaging material. Food Hydrocoll. 2021, 121, 107019. [Google Scholar] [CrossRef]
  164. Dissanayake, T.; Chang, B.P.; Mekonnen, T.H.; Ranadheera, C.S.; Narvaez-Bravo, C.; Bandara, N. Reinforcing canola protein matrix with chemically tailored nanocrystalline cellulose improves the functionality of canola protein-based packaging materials. Food Chem. 2022, 383, 132618. [Google Scholar] [CrossRef] [PubMed]
  165. Trifol, J.; Moriana, R. Barrier packaging solutions from residual biomass: Synergetic properties of CNF and LCNF in films. Ind. Crop. Prod. 2022, 177, 114493. [Google Scholar] [CrossRef]
  166. Shi, C.; Ji, Z.; Zhang, J.; Jia, Z.; Yang, X. Preparation and characterization of intelligent packaging film for visual inspection of tilapia fillets freshness using cyanidin and bacterial cellulose. Int. J. Biol. Macromol. 2022, 205, 357–365. [Google Scholar] [CrossRef]
  167. Ezati, P.; Rhim, J.-W.; Molaei, R.; Priyadarshi, R.; Han, S. Cellulose nanofiber-based coating film integrated with nitrogen-functionalized carbon dots for active packaging applications of fresh fruit. Postharvest Biol. Technol. 2022, 186, 111845. [Google Scholar] [CrossRef]
  168. Zhou, W.; Wu, Z.; Xie, F.; Tang, S.; Fang, J.; Wang, X. 3D printed nanocellulose-based label for fruit freshness keeping and visual monitoring. Carbohydr. Polym. 2021, 273, 118545. [Google Scholar] [CrossRef]
  169. Haishun, D.; Chao, L.; Miaomiao, Z.; Qingshan, K.; Bin, L.; Mo, X. Preparation and industrialization status of nanocellulose. Prog. Chem. 2018, 30, 448. [Google Scholar]
  170. Nelson, K.; Retsina, T.; Iakovlev, M.; Van Heiningen, A.; Deng, Y.; Shatkin, J.A.; Mulyadi, A. American Process: Production of Low Cost Nanocellulose for Renewable, Advanced Materials Applications. In Artificial Intelligence for Materials Science; Springer: Berlin/Heidelberg, Germany, 2016; pp. 267–302. [Google Scholar]
  171. Bondancia, T.J.; de Aguiar, J.; Batista, G.; Cruz, A.J.G.; Marconcini, J.M.; Mattoso, L.H.C.; Farinas, C.S. Production of Nanocellulose Using Citric Acid in a Biorefinery Concept: Effect of the Hydrolysis Reaction Time and Techno-Economic Analysis. Ind. Eng. Chem. Res. 2020, 59, 11505–11516. [Google Scholar] [CrossRef]
  172. Ioelovich, M.J. Microcellulose Vs Nanocellulose–A Review. World J. Adv. Eng. Technol. Sci. 2022, 5, 001–015. [Google Scholar] [CrossRef]
  173. Amoroso, M.; Apelgren, P.; Säljö, K.; Montelius, M.; Orrhult, L.S.; Engström, M.; Gatenholm, P.; Kölby, L. Functional and morphological studies of in vivo vascularization of 3D-bioprinted human fat grafts. Bioprinting 2021, 23, e00162. [Google Scholar] [CrossRef]
  174. Saito, T.; Kimura, S.; Nishiyama, Y.; Isogai, A. Cellulose Nanofibers Prepared by TEMPO-Mediated Oxidation of Native Cellulose. Biomacromolecules 2007, 8, 2485–2491. [Google Scholar] [CrossRef]
  175. Isogai, A. Emerging Nanocellulose Technologies: Recent Developments. Adv. Mater. 2020, 33, 630. [Google Scholar] [CrossRef]
  176. Lyne, B. Market prospects for nanocellulose. In The Royal Institute of Technology; Alberta Biomaterials Development Centre: Edmunton, AB, Canada, 2013. [Google Scholar]
  177. Isogai, A. Cellulose Nanofibers: Recent Progress and Future Prospects. J. Fiber Sci. Technol. 2020, 76, 310–326. [Google Scholar] [CrossRef]
  178. Tom, L.; Christian, A.; Magnus, G.; Torgny, P. The Emergence of Practical Nanocellulose Applications For A More Sustainable Paper/Board Industry. Q. J. Indian Pulp Pap. Technol. Assoc. 2015, 1, 53–61. [Google Scholar]
  179. Härkönen, M.; Tammelin, T.; Lille, M.; Qvintus, P.; Laine, J.; Koskinen, T.M. The Nanocellulose Challenge. In Proceedings of the 2009 Wood and Fiber Product Seminar, Espoo, Finland, 22–23 September 2009; p. 59. [Google Scholar]
  180. Kyle, S.; Jessop, Z.M.; Al-Sabah, A.; Hawkins, K.; Lewis, A.; Maffeis, T.; Charbonneau, C.; Gazze, A.; Francis, L.W.; Iakovlev, M.; et al. Characterization of pulp derived nanocellulose hydrogels using AVAP® technology. Carbohydr. Polym. 2018, 198, 270–280. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustration of cellulose molecule, reprinted with permission from [26], 2011, RSC.
Figure 1. Schematic illustration of cellulose molecule, reprinted with permission from [26], 2011, RSC.
Polymers 14 04025 g001
Table 1. Examples of the geometrical dimensions and crystallinity index of nanocellulose from various sources.
Table 1. Examples of the geometrical dimensions and crystallinity index of nanocellulose from various sources.
SourceLength (nm)Width (nm)Crystallinity Index (%)References
Softwood483 ± 2324.1 ± 1.2/[49]
Cotton linter1771290.45[44]
Algae315 ± 309 ± 381[50]
Bacteria/64.6 ± 15.390.3[51]
Tunicate2100 ± 7008.7 ± 2.4/[49]
Table 3. Methods for the preparation of BC.
Table 3. Methods for the preparation of BC.
Culture MediumBacteriaConditionsCharacteristicsReferences
Wine pomaceKomagataeibacter rhaeticus K1530 °C, 10 daysYield 1.95 ± 0.22 g/L, nanocellulose concentration 91.67 ± 2.76%, crystallinity index 90.61%, diameter range 30–130 nm[113]
HSTaonella mepensis30 °C, 7 daysYield 2.472 g/L, crystallinity index 90.3%, average width 64.6 ± 15.3 nm[51]
STTaonella mepensis30 °C, 7 daysYield 1.784 g/L, crystallinity index 82.8%, average width 53.3 ± 20.1 nm[51]
Sugarcane bagasseKomagateibacter xylinus28 °C, 9 daysFiber diameter 47 ± 10 nm, crystallinity index 79%, water content 99.43 ± 0.03%[114]
Table 4. Nanocellulose-based composites for food packaging applications.
Table 4. Nanocellulose-based composites for food packaging applications.
Cellulose ComponentOther ComponentsMethodPerformanceReferences
CNCChitosanSolution castingThe tensile strength and Young’s modulus of the film have been increased by 39% and 78%, respectively. Water solubility has been reduced by 26.5–35.7%, with good UV resistance and water repellency.[116]
CNFPolyurethane(PU), quaternized chitosan (QCS) and negatively charged phosphotungstic acidLayer-by-Layer assemblyThe conductivity of hydroxide reached 14.3 mS/cm at 80 °C and lasted for more than one month.[117]
CNCPolylactic acid (PLA)ExtrusionImproved processability, melt strength, and rheological properties. Good performance in storing oil-based and dairy products can prolong their shelf life.[118]
CNCPalm oil/waterEmulsionsChecking whether food is spoiled or not.[119]
BCProtein nanoparticlesHydrogelGood wettability, interfacial adsorption capacity, and higher antioxidant property.[120]
CNFPolylactic acid (PLA), Chitosan, rosinSpray dryingGreat antibacterial effect and increased elasticity and water vapor permeability.[121]
CNCChitosan, polyvinyl alcoholElectrospinningPreventing the growth of pathogenic bacteria [122]
CNFPolyvinyl alcohol (PVA), Silver NanoparticlesElectrospinningGood antibacterial activity against Staphylococcus aureus, E. coli, and P. aeruginosa.[123]
BCBovine Lactoferrin (BLF)AdsorptionStrong bactericidal effect on E. coli and Staphylococcus aureus.[124]
CNCPolyethylene glycol, algal bile proteinAdsorptionImproved protein stability.[125]
Table 5. Application and performance of nanocellulose-based food packaging materials.
Table 5. Application and performance of nanocellulose-based food packaging materials.
CompositionPerformanceReferences
CS/carboxymethylated nanocelluloseIncreased resistance to grease, oil, water, air, and water vapor. Good mechanical characteristics and enhanced antibacterial activity against both E. coli and Staphylococcus aureus.[154]
Corn starch/BCImproved barrier to water vapor and oxygen.[155]
Gelatin/BC/cinnamaldehydeIncreased tensile strength and lower water vapor permeability. Inhibited against Staphylococcus aureus and E. coli germs.[156]
CMC/CNC/AgNPsExcellent mechanical strength, water vapor and air barrier characteristics, and antibacterial activities.[157]
SA/CNF/Ca2+/PSEHigh strength, good water resistance, excellent ultraviolet barrier performance, and significant antibacterial effects against both gram-negative and gram-positive bacteria.[158]
Anthocyanins/BCConsumers can judge the freshness of fish by comparing the label color with the standard color.[136]
EAE/BCThe fabricated BC-EAE indicator responded to pH by changing color from red to yellow over the pH range of 2–12.[159]
AgNPs@CS-1:1The storage time of strawberries packaged by AgNPs@CS-1:1 was extended to 12 days without microbial invasion.[160]
SA/Carboxymethylated nanocellulose/SOWEsGreat effect on controlling browning index in fresh-cut apple and potato over the storage of 12 days and 5 days.[161]
CS/guar gum/walnut green husk extractGood performance in reducing firmness, weight loss, total soluble solids, and inhibiting browning and microbial growth of fresh-cut apples.[162]
Table 6. Global nanocellulose production facilities with their products and capacities.
Table 6. Global nanocellulose production facilities with their products and capacities.
CompanyCountryMethodProductsProduction Capacity
INNVENTIASwedenEnzymatic and microfluidizerCNF100 kg/d
Nippon paperJapanTEMPO oxidation and mechanical defibrillationCNF150 kg/d
Stora EnsoSwedenEnzymatic and mechanical defibrillationMFCn.a.
CellForceCanadaSulfuric acid hydrolysisCNC1000 kg/d
U.S. Forest ServiceU.S.Sulfuric acid hydrolysis and drinderCNC and CNFCNC (10 kg/d) and CNF (1000 kg/d)
Cellulose LabCanadan.a.CNC, CNF and BCn.a.
American Process U.S.AVAPCNF1000 kg/d
University of MaineU.S.Mechanical refining methodCNF1000 kg/d
VTTFinlandEnzymatic pretreated with Masuko grinderCNF15 kg/d
FineCellSwedenOxalic acid hydrolysisCNFn.a.
Ocean TuniCell ASNorwayEnzymatic, TEMPO oxidation, carboxylmethylation, and microfluidizerTunicate CNFn.a.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, J.; Zhang, F.; Zhong, Y.; Zhao, Y.; Gao, P.; Tian, F.; Zhang, X.; Zhou, R.; Cullen, P.J. Emerging Food Packaging Applications of Cellulose Nanocomposites: A Review. Polymers 2022, 14, 4025. https://doi.org/10.3390/polym14194025

AMA Style

Li J, Zhang F, Zhong Y, Zhao Y, Gao P, Tian F, Zhang X, Zhou R, Cullen PJ. Emerging Food Packaging Applications of Cellulose Nanocomposites: A Review. Polymers. 2022; 14(19):4025. https://doi.org/10.3390/polym14194025

Chicago/Turabian Style

Li, Jingwen, Feifan Zhang, Yaqi Zhong, Yadong Zhao, Pingping Gao, Fang Tian, Xianhui Zhang, Rusen Zhou, and Patrick J. Cullen. 2022. "Emerging Food Packaging Applications of Cellulose Nanocomposites: A Review" Polymers 14, no. 19: 4025. https://doi.org/10.3390/polym14194025

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