Next Article in Journal
The Effect of Different Medium Compositions and LAB Strains on Fermentation Volatile Organic Compounds (VOCs) Analysed by Proton Transfer Reaction-Time of Flight-Mass Spectrometry (PTR-ToF-MS)
Previous Article in Journal
Precision Fermentation as an Alternative to Animal Protein, a Review
Previous Article in Special Issue
Bioproduction of 2-Phenylethanol by Yarrowia lipolytica on Sugar Beet Molasses as a Low-Cost Substrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bacterial Nanocellulose Produced by Cost-Effective and Sustainable Methods and Its Applications: A Review

Department of Materials Science and Bioengineering, School of Engineering, Nagaoka University of Technology, Nagaoka 940-2188, Niigata, Japan
Fermentation 2024, 10(6), 316; https://doi.org/10.3390/fermentation10060316
Submission received: 20 March 2024 / Revised: 8 June 2024 / Accepted: 11 June 2024 / Published: 14 June 2024

Abstract

:
This review discusses the recent advancements in cost-effective fermentation methods for producing bacterial nanocellulose (BC) from food and agro-industrial waste. Achieving economical cell culture media is crucial for large-scale BC production, requiring nutrient-rich media at low cost to maximize cellulose yield. Various pretreatment methods, including chemical, physical, and biological approaches, are stated to break down waste into accessible molecules for cellulose-producing bacteria. Additionally, strategies such as dynamic bioreactors and genetic engineering methods are investigated to enhance BC production. This review also focuses on the environmental impact assessment and updated application challenges of BC such as medical applications, energy storage/electronics, filtration membranes, and food packaging. By providing insights from the recent literature findings, this review highlights the innovative potential and challenges in economically and efficiently producing BC from waste streams.

1. Introduction

Utilizing biopolymers to manufacture products offers a sustainable approach that avoids increasing the carbon footprint. By doing so, carbon neutrality is maintained, facilitating the natural carbon cycle [1]. This strategy aligns closely with the objectives of the United Nations’ Sustainable Development Goals (SDGs), which emphasize the importance of transitioning from petroleum-based polymers to bio-based alternatives. These sustainable alternatives hold significant potential to alleviate the accumulation of single-use products in the environment, primarily due to their non-biodegradability and inadequate waste management practices [1,2].
Cellulose, among various biopolymers, emerges as a particularly promising candidate for widespread adoption in various applications such as textiles [3], packaging [4], functional papers [5], and biomedical applications [6]. Its abundance in nature and versatility makes it an attractive option. Through the utilization of cellulose-based materials, dependency on fossil fuels to the synthesis of petroleum-based polymers can be decreased, carbon emissions can be lowered, and environmental deterioration can be alleviated [7]. Recent advancements in engineering have facilitated the cost-effective production and widespread availability of various cellulose-derived products, positioning it as a viable alternative to conventional plastics. Additionally, cellulose-based plastics offer biodegradable alternatives that can help address concerns regarding waste accumulation and its adverse effects on ecosystems [8].
In plants, cellulose is intricately bound with other biopolymers like hemicellulose, lignin, and pectin. These non-cellulosic biopolymers are typically removed through acid/alkali pretreatments and bleaching processes. Additionally, nanocellulose—a highly sought-after material for its unique properties—can be obtained from plants using nanofibrillation techniques. Moreover, cellulose can be synthesized by bacteria, resulting in BC, produced by species like Acetobacter (A.) (also known as Gluconacetobacter (G.) or Komagataeibacter (K.)) [9,10,11,12], Taonella mepensis (T. mepensis) [13], and Lactiplantibacillus plantarum (L. plantarum) [14]. BC, distinguished by its high purity devoid of hemicellulose, lignin, or other natural components, exhibits physicochemical properties that differ from those of plant cellulose. Notably, BC possesses an ultrafine nanoscale mat structure (see Figure 1) with a high degree of polymerization, crystallinity, mechanical strength, water retention capacity, porosity, and surface area [15]. These properties, vital for various applications, can be further tailored through precise fermentation strategies, selection of microbial strains [16], and modifications such as using acid [17] and plasma treatments [18]. Producing cellulose using bacteria eliminates the need for chemical treatments associated with plant cellulose extraction, thereby minimizing the environmental impact [19].
The Hestrin–Schramm (HS) medium stands as the predominant choice for BC production, yet its synthetic nature contributes to high process expenses [21,22]. The affordability of the culture medium is a decisive factor in the production of BC, making the search for cost-effective alternatives a fundamental aspect of cellulose biosynthesis. The attainment of a substantial BC yield is intricately linked to this endeavor [23]. To effectively address the imperative of cost reduction in BC manufacturing, a substantial focus has been placed on using waste sugars sourced from a variety of outlets [24].
The aggregate count of publications pertaining to BC has demonstrated a consistent annual increase over the past decade, as illustrated in Figure 2A. This trend indicates a growing interest and significance in BC research. Moreover, Figure 2B illustrates the extensive spectrum of applications attributed to BC across diverse fields. This depiction reveals the versatile potential and broad applicability of BC in various domains. In recent decades, an increasing number of studies have acknowledged the importance of large-scale production to accelerate the commercialization of BC. Accordingly, strategies have been proposed for utilizing agro-industrial biomass as cheap culture media [25]. The utilization of waste in BC production aligns with the zero-waste concept, contributing to sustainability and mitigating environmental issues associated with waste disposal [26].
Therefore, this review highlights approaches to fermenting waste and other strategies for economic feasibility and environmental sustainability in BC production from agro-industrial waste by summarizing the state of knowledge over the past decade. These findings show the potential for future advancements in utilizing waste streams for sustainable production processes and utilizing BC in various applications.

2. Industrial Wastes Used in Fermentation

Choosing a cost-effective culture medium aligns with the contemporary trend toward a zero-waste economy, emphasizing the importance of repurposing food manufacturing waste [27]. Utilizing waste from the food supply chain to produce chemicals, biopolymers, biofuels, and bioactive compounds reflects a broader transition toward a circular bio-economy, marking a significant shift in industrial practices [28,29]. Organic waste, generated at various stages of agro-industrial processes, is often deemed unsuitable for commercial use due to quality concerns. However, it represents a reservoir rich in sugars such as glucose, fructose, and sucrose, along with nitrogen and vitamins [30]—precisely the components essential for cellulose biosynthesis. Over the last thirty years, molasses and corn steep liquor have emerged as prominent alternative sources of complex carbon and nitrogen for BC production [24,31]. Molasses derived from sugar cane or beets has been found to be effective in BC production as a carbon source [24,31]. In recent years, researchers have increasingly investigated other waste sources for BC production, as shown in Table 1.
The utilization of grapes as an economical carbohydrate source for BC production has been studied [32,33]. Typically, unpressed residues or grape extracts rich in free monosaccharides have been employed as alternative carbon sources in HS media or in alternative media supplemented with corn steep liquor as a nitrogen source [34]. However, the primary waste generated in wineries is grape marc or grape pomace, which, post-fermentation and pressing, contains limited available monosaccharides for fermentation [35]. Conversely, many studies have focused on separating components of the grape marc (such as seeds, stalks, pulp, or skin) and subjecting them to separation processes to utilize mainly the components with the highest monosaccharide content [35]. However, such additional processing steps elevate production costs. The acid hydrolysis process applied to wine bagasse residues offers a straightforward and cost-effective method to degrade cellulose and hemicellulose structures, thereby enhancing the concentration of fermentable monosaccharides and BC production yield [36].
The utilization of potato waste to produce BC at a low cost has been assessed [37]. In this instance, potato tuber juice is used as the culture medium for BC biosynthesis, obtained similarly to industrial starch production methods [38]. The potato juice medium necessitates no pretreatment or supplementation with additional nutrient sources, apart from the inclusion of 1% v/v ethanol. Following the dilution of the potato juice with water at a 1:1 ratio to formulate the medium, the produced BC yield is similar to that achieved using a commercial HS medium. Moreover, the BC derived from the potato juice shows no variance from conventionally produced HS-BC concerning its physical, chemical characteristics, and cytotoxicity, affirming its comparable quality [38]. While the juice can be directly incorporated into culture media, macromolecular waste such as cellulose in potato peels necessitates hydrolysis to obtain smaller molecules like monosaccharides and disaccharides before being introduced into the medium [37]. The culture medium solely comprised of sugars derived from the acid hydrolysis of potato peels has the potential to enhance BC production yield or facilitate the creation of functionalized films with antioxidant properties [37].
Table 1. The utilization of waste-derived nutrient sources in the production of BC, based on recent research findings published from 2019 to 2024.
Table 1. The utilization of waste-derived nutrient sources in the production of BC, based on recent research findings published from 2019 to 2024.
Waste SourceBacteria NameCultivation ConditionBC Yield (g/L)Ref.
Grape pomace hydrolysateK. melomenususStatic culture, 3 days, 30 °C1.2[32]
Potato peelG. xylinusStatic culture, pH 6.0, 4 days, 30 °C1.27–2.61
HS a: 1.21
[37]
Potato juiceK. xylinusStatic culture, pH 6.0, 7 days, 28 °C2–4
HS a: 3–5
[38]
Straw biomass (sugarcane bagasse, bamboo, corncob, wheat straw, and rice straw)A. xylinusStatic culture, 6 days, 30 °C1.9–2.6
HS a: 1.7
[39]
Orange peel wasteG. xylinus
K. sucrofermentants
Static culture, 8 days, 30 °C
Dynamic culture, 30 °C
0.9–6.1
HS a: 0.9
1.9–11.6
[40,41,42,43,44]
Pineapple peel wasteG. xylinusStatic culture, pH 4.5, 14 days, 30 °C3.8
HS a: 2.1
[45,46,47]
Pineapple coreK. xylinusStatic culture, pH 4, 15 days, room temperature2.4–2.5
HS a: 2.5
[48]
Mango wasteK. xylinusStatic culture, 16 days, 30 °C1–6
HS a: 2.0
[49,50]
Pear peel wasteL. plantarumStatic culture, 8 days0.7–3.5
HS a: 1.9
[14]
Watermelon wasteG. xylinusStatic culture, 16 days5.8[51]
Asparagus wasteK. rhaeticusStatic culture, pH 4.5, 25 days, 30 °C1.0–2.5[52]
Sweet lime pulp wasteA. xylinusDynamic culture, pH 5–6, 7 days, 25–35 °C5.2–7.0[53]
Jasminum sambac and Camellia sinensisK. intermediusStatic culture, 7 days, 30 °C3.7–7.1
HS a: 5.6
[54]
Coffee groundK. rhaeticusStatic culture, 7 days, 30 °C0.5–11[55,56]
Liquid wastes from preserved tamarind and preserved mangoA. xylinusStatic culture, 10 days, 30 °C4.7 and 4.5
HS a: 2.5
[57]
Olive oil mill wastewaterK. xylinusStatic culture, pH 4.5, 7 days, 30 °C1–5
HS a: 1
[58]
Sago residueG. xylinusStatic culture, pH 6.0, 14 days, 30 °C1.55
HS a: 1.57
[59,60]
Pecan nutshellG. entaniiStatic culture, pH 3.5, 28 days, 30 °C2.8[61]
Residue of cashew apple juice processingK. xylinusStatic culture, pH 4.3, 12 days, 30 °C1–3
HS a: 0.5–3
[62]
Brewing by-productsK. rhaeticusStatic culture, pH 6, 10 days, 30 °C4.0[63]
Beet molasses, vinasse, and waste beer fermentation brothK. xylinusDynamic culture, pH 5, 7 days, 30 °C8.2
HS a: 1.7
[64]
Mulberry pomace waste extractK. xylinusStatic culture, pH 8, 10 days, 30 °C1.5[65]
Cheese wheyG. xylinusStatic culture, pH 5.5, 14 days, 28 °C1–3.55
HS a: 3.26
[66]
Tofu wastewaterK. xylinusStatic culture, pH 4.5, 15 days, 30 °C10.6[67]
Soybean residueG. xylinusStatic culture, pH 4.5, 15 days, 30 °C1.5–2.8
HS a: 5.1
[68]
Waste and by-product streams from biodiesel and confectionery industriesG. xylinusStatic culture, pH 5.0, 7 days, 30 °C1–7.32[69]
Tobacco waste extractG. xylinusStatic–shaking (150 rpm) cultures, pH 6.5,
7–15 days, 30 °C
0.5–5.2
HS a: 3.26
[70]
Rice-washed waterK. xylinusStatic culture, pH 4.5, 15 days, 30 °C6.57[67]
Kitchen wasteA. xylinum
K. rhaeticus
Static culture, 15 days, 30 °C
Static culture, 10 days, 30 °C
2.0
4.7
[71,72]
Paper waste sludge hydrolysateA. xylinumStatic culture, pH 4.8, 15 days, 30 °C12.0[73]
Beeswax recycling wastewaterK. xylinusStatic culture, pH 6, 14 days, 28 °C2.6
HS a: 2.7
[74]
Spent black liquor from cotton pulpingG. xylinusStatic culture, pH 5.5, 2–3 weeks, 30 °C0.2–3.25
HS a: 4.66
[75]
Textile wasteT. mepensisStatic culture, 14 days, 30 °C1–2.2
HS a: 2.2
[76]
a Yield of BC in conventional HS medium used as a reference.
The obstacle in utilizing lignocellulosic residues lies in their resistant structure, necessitating a pretreatment process to deconstruct their complexity. This pretreatment step aims to enhance the accessibility of cellulose and hemicellulose by disrupting hydrogen bonding within the lignocellulose structure for effective hydrolysis, which releases sugars essential for microbial utilization [77]. The acidic hydrolysis of lignocellulosic and starchy wastes has been found to enhance the release of fermentable sugars, facilitating microbial product production. Different mineral acids such as HCl, HNO3, and H2SO4 employed for the saccharification are studied in the potato peel waste [37]. The BC yield is higher in potato peel waste acid hydrolysate media due to their buffering capacity, particularly in potato peel waste-HNO3-hydrolysate and potato peel waste-H2SO4-hydrolysate (2.2 and 1.8-fold increases compared to the HS medium) [37]. While biomass waste, consisting of cellulose, hemicellulose, and lignin, shows promise as an alternative carbon source, its utilization via acid hydrolysis to convert cellulose into fermentable sugars may introduce fermentation inhibitors like furfural, 5-hydroxymethyl-furfural (HMF), and formic acid into the hydrolysate. These carboxylic acids and furan aldehydes, might exhibit toxicity to the bacteria, potentially impeding their growth and fermentation process [78]. Various studies have recommended the adoption of detoxification procedures prior to fermentation when employing cellulosic acid hydrolysis. Cheng, Yang et al. [79] demonstrated that active carbon adsorption could remove at least 80% of furfural from acid-hydrolyzed corn stalks while retaining 90% of the total sugars. Additionally, they utilized an ion-exchange resin for detoxification and observed that ion exchange was more effective than active carbon in degrading acetic acid, although it resulted in lower removal of furfuran compounds [79]. Despite active carbon’s efficient adsorption of furan inhibitors, it may also eliminate other beneficial compounds [79]. For instance, the active carbon effectively removes inhibitory compounds and also adsorbs total phenolics, indicating its non-specific adsorption properties [80]. This trait may lead to a reduction in nutrients in the hydrolysate, thereby restricting its use in fermentation processes. Lin, Huang et al. [81] studied sulfuric acid hydrolysis of sugarcane bagasse, and atmospheric cold plasma (ACP) to eliminate the toxic inhibitors. Under the optimized ACP conditions (argon ACP at 200 W for 25 min), the degradation rates of formic acid, HMF, and furfural reached 25.2, 78.6, and 100%, respectively. In BC production by K. xyinus, the ACP-treated sugarcane bagasse hydrolysate group demonstrated high BC production (1.68 g/L), but it was still lower than that of the ACP-untreated sugarcane bagasse hydrolysate (1.88 g/L) [81]. This result is similar to ACP-assisted pineapple peel waste hydrolysate detoxification for the production of BC by K. xyinus [82]. The BC produced by K. xylinus in both Ar and air-plasma-treated media exhibits yields comparable to BC produced by various other strains. It is noteworthy that BC production in fermentation using ACP plasma-treated media results in higher yields (3.82 g/L) compared to several reported strains. Specifically, A. xylinum achieves a BC yield of 2.86 g/L using acetic acid pre-hydrolysis liquor of corn stalks [79], G. xylinus yields 2.53 g/L of BC from litchi extract [83], and A. aceti yields only 1.73 g/L of BC after fermentation in optimized HS media [84]. Despite this, the BC yield from K. xylinus in this study remains lower than that reported using citrus peel and pomace media by enzymolysis for K. xylinus (5.7 g/L) [85] and optimized HS media for K. intermedius (3.91 g/L) [86]. These reports suggest that detoxification by plasma treatment may also lead to some loss of crucial nutrients in the hydrolysate, resulting in decreased BC production.
The enzymatic hydrolysis process offers a more environmentally friendly alternative to conventional acid treatment, operating at lower temperatures. For example, amylase, e.g., alpha-amylase and glucoamylase, is used to release the reduced sugar from the biomass [87]. This not only reduces the energy consumption and operational costs in biomass conversion but also minimizes the environmental impact [88]. However, a drawback is the use of commercial enzymes, which might increase production costs. One potential solution is to produce hydrolytic enzymes onsite, like crude glucoamylase [89], cellulase [90], and xylanase (hemicellulase) [91], from waste lignocellulose through fermentation. This approach allows for the onsite generation of enzymes, thus reducing the reliance on commercial enzymes [91]. Those enzymes can be produced by bacteria such as Streptomyces rochei (S. rochei) [92] or fungi like Trichoderma reesei (T. reesei) [14,93]. Enzymatic hydrolysis increases the content of simple sugars with the subsequent hydrolysis time showing minimal effects on sugar release [14]. Waste jasmine flowers are utilized to produce BC with antimicrobial properties through a sequential fermentation process involving the fungus T. reesei and the bacterium T. mepensis. Initially, waste jasmine flowers are employed to generate cellulase and xylanase with T. reesei. The waste jasmine flowers are then enzymatically hydrolyzed, and the resulting hydrolysate is utilized for BC production. The highest BC yield from the supplemented jasmine flower hydrolysate is 2.1 times higher than that achieved with the HS medium [93].
The abundant availability of paper sludge, a primary waste product from paper manufacturing, presents both environmental and economic challenges. Fermentation medium costs represent a significant portion of BC production expenses, making the utilization of paper sludge for BC production an attractive prospect to mitigate the environmental impact and reduce costs [94]. Chemical analysis reveals a high sugar content in paper sludge, primarily consisting of cellulose and hemicellulose, but low levels of simple sugars and glucose in crude paper sludge due to the complex nature of its cellulosic materials. This necessitates pre-hydrolysis such as saccharification by cellulase produced by S. rochei to support BC production [92]. Utilization of starchy kitchen wastes (SKWs) as a cost-effective substrate is used for BC production [71,95]. By employing a Box–Behnken design, the maximum production of reducing sugars from SKWs through enzymatic hydrolysis using amylase produced by Bacillus methylotrophicus SCJ4 is optimized. The study found that optimal conditions yielded 148.0 g/L of reducing sugars and 2.11 g/L of BC [95].
The disposal of waste from the dairy industry is a significant concern due to the high-water usage in factories. Effluents from dairy processing contain lipids, lactose, detergents, and sanitizing agents that require treatment before being discharged into the environment [96,97]. Dairy wastewater typically contains elevated levels of total nitrogen (14–830 mg/L) and total phosphorus (9–280 mg/L). The nitrogen sources also affect BC production [98]. Therefore, utilizing dairy waste offers the dual advantage of wastewater purification and viable BC production [96,99]. Whey, a by-product of cheese or soybean curd manufacturing, contains approximately 55% of the nutrients found in the original milk, including soluble protein, lactose, minerals, and B-group vitamins [66]. The cheese whey can be used to replace half of the expensive yeast extract in the conventional HS medium, facilitating economic BC production [100]. Cultivating G. xylinus using untreated or lactose-pretreated cheese whey results in BC yields of 1 g/L and 3 g/L, respectively, after 14 days of static cultivation [66]. Soybean residue, rich in amino acids comparable to commercial yeast extract, can serve as a sole nitrogen source when supplemented with sucrose in the culture medium [101]. After 7 days of cultivation with a 10% G. xylinus inoculum, BC yields ranging from 1.5 g/L to 2.8 g/L are obtained, depending on the pretreatment method used to increase the water solubility of the soybean residue [68]. However, when the soybean residue is the sole nitrogen source in the HS-based culture medium, BC production decreases to approximately 1 g/L [68].

3. Strategies to Improve BC Yield

3.1. Dynamic Bioreactor

Traditionally, BC is generated through static culture methods, where cellulose is formed as a gelatinous pellicle at the interface between the air and liquid. However, once BC pellicles reach a certain thickness, production reduces due to nutrient depletion in the aerobic zone [77,102] and oxygen deficiency in the lower regions of the pellicle [103]. Prolonged fermentation under these conditions offers minimal contribution to BC production [104]. The optimization of three key factors—incubation temperature, shaking frequency to maximize oxygen availability, and pH—is performed using response surface analysis [53,105]. This analysis is based on the actual and predicted responses obtained from a Box and Behnken design experiment, which involves three factors each at three levels [53]. An alternative lies in continuously feeding the fresh medium onto the pre-formed BC pellicle at the air–liquid interface [106]. The activated bacteria can access sufficient nutrients and oxygen from the fresh medium to sustain BC synthesis. Consequently, newly formed BC continuously occurs at the air–liquid interface, regardless of the location of previously existing pellicles. Implementing intermittent feeding in this manner facilitates continuous BC synthesis and enhances overall BC production [106].
Fed-batch or intermittent feeding strategies supplying the culture with fresh media under sterile conditions to sustain BC synthesis are promising strategies to scale-up BC production [106]. By using enzymatically hydrolyzed paper sludge (EHPS) as a culture medium, in fed-batch fermentation, the HS medium outperforms EHPS/HS and EHPS. BC production increases substantially in fed-batch compared to batch fermentation across all media types. The BC production in fed-batch fermentation increases approximately 2.1, 4.6, and 2.9-fold compared to batch fermentation when using HS, EHPS/HS, and EHPS media, respectively [92]. For Alcohol Lees, a by-product of alcohol production, BC is produced using an Intermittent Feeding Strategy [107]. Through the utilization of the respond surface methodology, the synergistic effects between Alcohol Lees dilution, pH, intermittent feeding height, and intermittent feeding interval are analyzed. The optimal conditions identified for maximum BC yield are Alcohol Lees dilution of 86%, pH of 5.6, intermittent feeding height of 1.4 mm, and intermittent feeding interval of 22.2 h, resulting in a BC yield of 4.4 g/L. Comparatively, conventional static cultures of BC yield 2.0 g/L in a one-week culture period [107]. The BC production through G. xylinus using glycerol as a carbon source is studied in a 3 L bench-top bioreactor under fill-and-draw and pulse-feed fed-batch cultures at 30 °C [108]. The fill-and-draw fed-batch culture accumulated 24.2 g/L of BC with a yield and productivity of 0.2 g/g and 2.69 g/L/day, respectively, after 9 days. The pulse-feed fed-batch culture yielded 24.38 g/L of BC with a yield of 0.38 g/g and productivity of 2.71 g/L/day. The study suggests that fed-batch cultivation, particularly pulse-feed, holds promise for industrial BC production, potentially reducing costs [108].
A rotating biological contactor (RBC) has been employed as a bioreactor to enhance the production of BC (Figure 3A) [109]. The RBC utilizes mechanical rotations to introduce air to the biofilm, promoting oxygenation to support growth [110]. The increased rotation speed enhances oxygen transfer, while RBC offers advantages such as low energy consumption, minimal maintenance, operational cost efficiency, and a compact design, making it an environmentally friendly option for BC production [109,111]. Factors influencing RBC performance include disc rotation speed, disc immersion, surface roughness, material, spacing, biofilm characteristics, and dissolved oxygen (DO) levels [112]. The impact of disk rotation speed, disk spacing, disk material, surface roughness, and aeration is individually examined using a one-factor-at-a-time approach [112]. The findings reveal that the greatest quantity of BC is generated on the surfaces of integrated polyethylene discs rotating at 13 rpm. Moreover, optimal BC production occurs when the distance between adjacent discs is set at 1 cm, with a total of 16 discs. Using an aquarium pump to aerate the RBC, comprised of 12 integrated polyethylene discs, operating at the optimal rotation speed of 13 rpm, results in an increase in both the wet weight and dry weight of BC. Specifically, this aerated fermentation leads to a more than 64% increase in wet weight and a 47% increase in dry weight compared to non-aerated RBC [112].
In the airlift bioreactor, Vázquez et. al. (2023) [113] investigated the viability of utilizing a culture medium derived from Garnacha Tintorera grape bagasse, rich in phenolic compounds, and waste potatoes, a glucose source, for simultaneous production of BC and gluconic acid through K. xylinus fermentation (Figure 3B). Acid hydrolysis of the grape bagasse enhances the phenolic content while minimizing growth-inhibiting by-products, e.g., furfural and HMF. Statistical analysis identifies acid concentration and temperature as pivotal variables, suggesting the optimal hydrolysis conditions of 2% sulfuric acid concentration for 60 min at 125 °C. These conditions enable a significant increase in total phenolic content compared to aqueous extraction [113].
Utilizing different force fields, like the magnetic force field, presents a promising alternative to enhance biomass production compared to conventional methods [114,115,116]. Studies have demonstrated the efficacy of static magnetic fields (SMF) in stimulating microbial growth and enhancing enzyme production. Furthermore, the potential of rotating magnetic fields (RMF) in diverse biotechnological applications, including the biogenesis of BC is studied in a bioreactor equipped with RMF generators (Figure 3C) [116,117,118]. The utilization of RMF has been shown to enhance the growth and metabolic activity of K. xylinus [118]. Furthermore, employing RMF-assisted bioreactors has led to an increase in BC yield compared to conventional static methods because they enhance both the quantity and quality of K. xylinus inoculum, consequently leading to a higher yield of the resulting BC [115]. However, the RMF application does not result in a higher incidence of cellulose non-producing mutants within the 72 h exposure period [119].
Figure 3. Bioreactors for BC production. Rotating biological contactor (A) (Reprinted with permission from ref. [109] Copyright © 2023 Elsevier). Airlift bioreactor (B) (Reprinted from ref. [113] open access Copyright © 2023 by the authors. Licensee MDPI). Bioreactor equipped with generators of rotating magnetic fields: front view (a), top view (b), schematic of the experimental setup (c). 1—RMF generator, 2—phase inverter, 3—PC, 4—generator tank, 5—inner container, 6—CX-701 multimeter, 7—temperature probe, 8—circulation pump, 9—plate heat exchanger, 10—three-way valve, 11—valve controller, 12—plastic probe (with cells suspension) (C) (Reprinted with permission from ref. [116] open access Copyright © 2022 Elsevier).
Figure 3. Bioreactors for BC production. Rotating biological contactor (A) (Reprinted with permission from ref. [109] Copyright © 2023 Elsevier). Airlift bioreactor (B) (Reprinted from ref. [113] open access Copyright © 2023 by the authors. Licensee MDPI). Bioreactor equipped with generators of rotating magnetic fields: front view (a), top view (b), schematic of the experimental setup (c). 1—RMF generator, 2—phase inverter, 3—PC, 4—generator tank, 5—inner container, 6—CX-701 multimeter, 7—temperature probe, 8—circulation pump, 9—plate heat exchanger, 10—three-way valve, 11—valve controller, 12—plastic probe (with cells suspension) (C) (Reprinted with permission from ref. [116] open access Copyright © 2022 Elsevier).
Fermentation 10 00316 g003
A magnetically assisted external-loop airlift bioreactor (EL-ALB) equipped with RMF generators, facilitating K. xylinus inoculum development was studied by Żywicka, Ciecholewska-Juśko et al. [115]. The study introduces the EL-ALB equipped with RMF generators for K. xylinus inoculum production in three-cycle repeated fed-batch cultures within a 40 L working volume. Analyses reveal improved fluid hydrodynamics in the RMF-assisted EL-ALB compared to a non-RMF EL-ALB, suggesting enhanced mass transfer rates during fermentation and increased K. xylinus cell multiplication [120]. Fermentation in the RMF-assisted EL-ALB yields inoculum with over 200 times higher cellular density compared to traditional methods, maintaining high and stable metabolic activity across repeated batch fermentation cycles. Importantly, RMF-assisted inoculum production does not induce cellulose-deficient mutants, and K. xylinus demonstrates consistent BC production regardless of inoculum age, achieving 7.26 g/L of dry mass [115].

3.2. Genetic Engineering

Studies into genes associated with BC synthesis offers valuable insights for the industrial utilization of BC-producing bacteria to generate BC from diverse carbon sources, including agro-waste feedstocks [121]. Based on these studies, an alternative method to improve BC production includes the utilization of synthetic biology tools and the alteration of metabolic pathways responsible for BC synthesis [122,123]. The genetic engineering involves leveraging adaptation mechanisms that bacteria develop in response to specific environmental conditions such as the adaptation ability to various carbon sources (Figure 4A) [124]. The effectiveness of genetic modification in fine-tuning K. xylinus is demonstrated. For example, in one study, the bcsA, bcsAB, and bcsABCD genes, which encode a complete cellulose synthase operon, are overexpressed in K. xylinus DSM 2325. This modification led to an accelerated production rate and a two to four-fold increase in BC yield [125]. Another study focuses on heterologously expressing the vgb gene from Vitreoscilla hemoglobin, resulting in a 50–70% boost in BC production under low oxygen conditions [126]. Furthermore, research on the pgi gene, responsible for encoding Escherichia coli (E. coli) phosphoglucose isomerase, demonstrates that its high expression enhances the metabolic pathway of BC synthesis, resulting in a remarkable 115.8% increase in BC production (Figure 4B) [127].
A recombinant strain for BC production in mannose-rich media is developed (Figure 4C) [128]. The introduction of the mak and pmi genes from E. coli into K. xylinus ATCC 23770 enhances the metabolic efficiency of mannose to fructose 6-phosphate. Comparing wild-type and recombinant strains reveals a 1.6-fold increase in the mannose utilization rate and an 84% boost in BC yield in the recombinant strain. Moreover, BC produced by the recombinant strain exhibits improved mechanical properties, with a 1.7-fold increase in tensile strength and elongation, and a 1.3-fold increase in Young’s modulus. The results suggest that the expression of the mak and pmi genes leads to better adaptation of the recombinant strain to the mannose carbon source and enhances the mechanical properties of BC. However, while the pBBR1MCS-2 plasmid carrying the exogenous gene confers kanamycin resistance, only K. xylinus ATCC 23770 exhibits tolerance to kanamycin upon transformation, suggesting potential limitations for strain transformation in other K. xylinus strains [128]. A study focuses on the characterization of Komagataeibacter sp. nov. CGMCC 17276 [129]. Whole-genome analysis reveals four BC synthase operons, particularly emphasizing the significance of bcs II and bcs III in BC synthesis under static and agitated culture conditions. Metabolic pathway analysis identifies glycerol as the most effective carbon source for BC production among eight potential options including glucose, sucrose, fructose, trehalose, maltose, mannitol, and ethanol. Glycerol is as an effective carbon source for BC production, with waste glycerol also presenting eco-friendly strategies for industrial BC production [129].

4. Environmental Impact Assessment

Global industrial focus is shifting toward more efficient production methods that promote sustainability, aiming to minimize the adverse environmental effects of manufacturing processes. The recent report from the Intergovernmental Panel on Climate Change emphasized the significance of decreasing net anthropogenic greenhouse gas emissions [130]. Life cycle assessment (LCA) has been utilized to evaluate the environmental implications of processes used to manufacture bio-based products, whether on a laboratory or industrial scale [131]. Currently, LCA studies have primarily focused on assessing the influence of various fermentation media on BC production. Overall, the BC production process has a relatively small impact on resource consumption and environmental factors in the global life cycle. The carbon footprint of BC production is often lower compared to cellulose extraction from plant biomass. Microcrystalline cellulose obtained from sugarcane bagasse, following acid–alkali pretreatment and depolymerization, exhibits the estimated carbon footprint of 261 kg CO2-eq/kg of cellulose [132] and nanocellulose derived from wood pulp by acid–alkali treatment and homogenization causes 80 kg CO2-eq/kg of nanocellulose [133]. The carbon footprint of BC typically falls within the range of 34–296 kg CO2-eq/kg BC [131,134,135]. These considerable variations can be attributed to differences in production methods, product yields, equipment, the specific LCA software used, inventory databases, and the system boundaries considered [131,134,135]. The lowest carbon footprint at 39 kg CO2-eq/kg BC is reported when utilizing nata-de-fique agro-waste for BC production [135].
A basic, small-scale LCA of BC production was conducted by using the SIMAPRO® software [136]. The study investigated the use of corn steep liquor (CSL) as a replacement for yeast extract and peptone in BC production [136]. The LCA compared two scenarios: one using CSL in the culture medium and one using the conventional HS medium. The results showed reductions in environmental impacts with the CSL-containing medium, including decreases in global warming, freshwater eutrophication, freshwater ecotoxicity, and human carcinogenic toxicity. The study also noted an increased environmental impact from the energy-intensive electric greenhouse used during fermentation, highlighting the importance of energy efficiency in minimizing environmental effects [136]. From the ReCiPe 2016 methodology, the production of BC requires significant resources, with water being the primary component, accounting for 36.1 tons per kilogram of BC produced [131]. Most of this water (98%) is returned to fresh waters after treatment. Raw material production contributes significantly to resource consumption, with 17.8 tons of water per kilogram of BC used, primarily for carton packaging, culture medium raw materials, and sodium hydroxide for BC washing [131]. Fermentation and downstream processes also utilize substantial amounts of water. However, substituting petroleum-derived plastics with bio-based alternatives could potentially decrease CO2-equivalent emissions by 30%, thereby reducing the global warming potential [131].
The LCA of the developed bioprocess demonstrates that using brewer’s spent grain extract (BSGE) and brewer’s spent yeast (BSY) in the culture medium can reduce the carbon footprint of 1 kg dry BC by 76% compared to commercial-based media [63]. The highest carbon footprint is observed with the HS medium, while the BSGE- and BSGE + BSYE-derived media resulted in 4.2 to 4.4-fold lower carbon footprints. This reduction is attributed to the higher BC concentration achieved and the substitution of commercial carbon and nitrogen sources with BSGE and BSYE [63]. Substituting raw materials in the HS medium with brewery-derived side-streams led to further reductions in total carbon footprint, with medium sterilization and inoculation accounting for 1–3% of the total carbon footprint in all cases [63].

5. Applications

The versatility and sustainability of cellulose and its derivatives make them promising candidates for addressing contemporary challenges in materials science, ranging from eco-friendly packaging solutions to advanced biomedical applications [137,138]. As research in this field progresses, leveraging the unique properties of BC holds immense potential for creating innovative, sustainable materials with diverse applications across industries.

5.1. Medical Application

The increasing demand for biodegradable and biocompatible biopolymers manufactured through environmentally friendly methods has surged owing to their minimal greenhouse gas emissions [139]. Especially, the COVID-19 pandemic has led to a notable increase in plastic pollution due to the use of single-use plastics for personal protective equipment and packaging. With increased awareness about the origins of medical products, there is a shift away from plastic-based materials [140]. Consequently, numerous researchers have focused on developing biodegradable polymers for medical applications [141]. Hence, BC is one of the suitable options to meet these criteria due to not only its animal-free and human-free sourcing, but also its biodegradability [142].
Several findings support the development of BC composites for targeted biomedical applications, particularly as antimicrobial wound dressings. The development of BC-based composites incorporating bactericidal nanoparticles like Se [142], Cu [143], Ag [144], ZnO [145], Au [146], and TiO2 [138] have been extensively researched. Nevertheless, the potential toxicity of these nanoparticles to healthy mammalian cells, resulting from the production of reactive oxygen species, cannot be overlooked despite their antibacterial properties [147,148]. There has been growing interest in utilizing green plant-based antimicrobials to create antimicrobial BC-based composites due to their potent antimicrobial properties and minimal side effects [149]. Studies have explored incorporating animal-based products such as propolis [150] and chitosan [151] or extracts from various plants into BC-based composites as summarized in Table 2. Various bioactive compounds such as curcumin [152,153], lignin [154], quercetin [155], carrageenan [156], and other plant extracts [157,158] have been employed in creating BC-based composites for wound healing and delivery of bioactive compounds. In particular, the addition of the natural extracts not only enhances antimicrobial activity [157] but also improves mechanical and other properties [159]. The plant extracts contain functional groups that serve as active antimicrobial agents, including polyphenols (flavonoids), chalcones, and other nitrogen-containing moieties (alkaloids, amines, etc.), which confer biological activity [157,158].
Research into the applications of pomegranate (Punica granatum L.) in medicine and antimicrobial treatments dates back centuries and continues in modern research [160]. Recent investigations suggest that pomegranate containing several phenolic compounds can serve as a natural alternative to synthetic antibacterial agents against various microbial pathogens [160,161]. Studies have also highlighted the potential of the pomegranate in treating conditions such as cancer, bacterial wound infections, among others, owing to its reported anti-carcinogenic, anti-inflammatory, and antioxidant properties [162]. An antibacterial composite using BC infused with pomegranate peel extract (PGPE) for potential biomedical uses was developed by Ul-Islam, Alhajaim et al. [163]. The BC was efficiently derived from food waste, and PGPE was incorporated into a BC hydrogel. The resulting BC-PGPE composite exhibited notable antibacterial properties against both Staphylococcus aureus (S. aureus) and E. coli strains. The composite could retain 97% of its dry weight in water and maintained this retention for over 48 h. Additionally, it showed superior reswelling ability compared to pure BC after three consecutive cycles of re-wetting [163]. A wound dressing that contains active compounds from mangosteen peel was developed by Marisca Evalina, Yulanda et al. [164]. The extract derived from mangosteen (Garcinia mangostana), primarily cultivated in tropical regions of Southeast Asia, offers a range of advantageous biological properties including antibacterial, antioxidant, anti-inflammatory, and anticancer effects attributed to its bioactive constituents such as xanthones, mangostin, tannins, and other polyphenolic compounds [165,166]. The wound dressing made from BC loaded with mangosteen peel extract exhibits flexibility and suitable thickness [164]. The incorporation of mangosteen peel extract into the dressing enhances wound healing after 15 days, as evidenced by improvements in wound closure. Furthermore, the optimal concentration of mangosteen peel extract for the dressing is determined to be 10%, suggesting its potential application in the development of treatments for diabetic ulcers [164].
Table 2. Natural products incorporated in BC for biomedical applications.
Table 2. Natural products incorporated in BC for biomedical applications.
Natutal ProductsIncoporation MethodsImproved PropertiesPotential ApplicationsRef.
PropolisImpregnation of BC into propolis/ZnO solutionAntimicrobial activity against E. coli, Bacillus subtilis (B. subtilis), and Candida albicans (C. albicans)Antimicrobial film[150]
ChitosanImpregnation of BC into chitosan solutionAntimicrobial activity against S. aureus, Pseudomonas aeruginosa (P. aeruginosa), and C. albicansAntimicrobial dressing[151]
CurcuminImpregnation of BC into curcumin solutionAntimicrobial activity against E. coli, P. aeruginosa, Salmonella typhimurium (S. typhimurium), and S. aureus, wound healing property (accelerated wound closure up to 64% after 15 days)Antimicrobial wound dressing[152,153]
LigninImpregnation of BC into lignin solutionAntimicrobial activity against P. aeruginosa, S. aureus, Serratia sp., Listeria monocytogenes (L. monocytogenes), and Salmonella typhimurium (S. typhimurium)Antimicrobial chronic wound dressing[154]
QuercetinPhase inversion of mixed BC and quercetin solutionDrug loading capacity and controlled drug release propertiesControlled-release drug delivery[155]
CarrageenanImpregnation of BC into carrageenan/gelatin solutionAntimicrobial activity against E. coli, S. aureus, and Klebsiella pneumonia, drug loading capacity and controlled drug release propertiesWound dressing and tissue regeneration[156]
Green tea leaf extract, roselle flower petals, and Hibiscus rosa-sinensis L. flower extractIn situ biosynthesis with BCAntimicrobial activity against P. aeruginosa and E. coli and antioxidant propertyWound dressing and face mask[157]
Combretaceae and Solanaceae extractImpregnation of BC into the plant extract solutionAntimicrobial activity against E. coli and S. aureusBiomedical materials[158]
Pomegranate peel extractImpregnation of BC into the plant extract solutionAntimicrobial activity against E. coli and S. aureusAntimicrobial wound dressing[163]
Mangosteen peel extractImpregnation of BC into the plant extract solutionWound healing property (accelerated wound closure up to >90% after 15 days)Wound dressing[164]
Euclea schimperi extractImpregnation of BC into the plant extract solutionAntimicrobial activity against E. coli and S. aureusAntimicrobial wound dressing[167]
Asparagaceae leave extractImpregnation of BC into the plant extract solutionAntimicrobial activity against E. coli and S. aureusAntimicrobial wound dressing[168]

5.2. Energy Storage and Electronics

The focus on renewable green energy presents a crucial avenue for addressing the contemporary energy crisis. Solar energy, as a widely accessible clean energy source, holds significant promise and has garnered extensive attention for its utilization [169]. Particularly, photothermal conversion emerges as a key method for harnessing solar energy, prompting intensive research efforts [170]. Various photothermal materials, including polymer materials, carbon-based materials, plasma metals, and semiconductor materials, have been explored [171]. A multifunctional photothermal phase change material (PCM), composed of silylated BC, hydroxylated carbon nanotubes (HCNT), and polyethylene glycol (PEG), has been developed for solar energy utilization (Figure 5A) [172]. This PCM features a cross-linked network structure, enhancing support, dispersion, and interlocking within the material. It demonstrates a high output voltage of 423 mV and power density of 30.26 W/m2 under solar irradiation, suitable for solar–thermal–electric energy conversion. Moreover, it finds applications in thermal management for solar cells and light-emitting diode (LED) chips, showing its versatility in waste heat conversion into electricity [172].
Rechargeable lithium-ion batteries (LIBs) are recognized as one of the most promising energy storage solutions for contemporary mobile electronics and hybrid electric vehicles due to their outstanding attributes, including high energy densities, extended cycling life, and minimal self-discharge rates [173]. One crucial element of LIBs is the separator membrane, a porous membrane located between the anode and the cathode. Its primary function is to prevent internal short circuits by ensuring that the two electrodes do not make direct contact with each other [174]. The study of Ajkidkarn and Manuspiya [175] explored the use of BC nanocrystals (BCNCs) derived from nata de coco waste as a bio-filler to enhance polyether block amide (PEBAX) separator membranes for LIBs. The BCNCs are obtained through sulfuric acid hydrolysis, yielding rod-like/needle-like structures with negatively charged surfaces. Non-solvent-induced phase separation and film casting methods are employed to fabricate BCNCs/PEBAX membranes, which exhibit improved thermal and dimensional stability up to 150°C without shrinkage. The addition of BCNCs enhanced the ionic conductivity and porosity of the membrane, making them promising candidates for replacing commercial polyolefin-based separator membranes in LIBs due to their superior electrochemical performance and reinforcement effect [175].
A proton exchange membrane requires high conductivity, thermal stability, absorbency, and cost-effectiveness [176]. BC shows promise as a sustainable material for these electronic applications and conducting materials [5,177,178]. However, pure BC lacks optical transparency, electrical conductivity, magnetism, hydrophobicity, and antimicrobial properties [176]. To enhance its capabilities for use in conducting materials and electrical devices, BC must undergo modification. This involves incorporating conductive materials like nano carbon fillers [5,179] and doping with acid and conducting polymers in nanoparticle or nanowire form [180]. Gadim et al. [181] fabricated proton-conducting electrolytes using a composite of BC and poly(4-styrene sulfonic acid) at room temperature (Figure 5B). A yield of 40 mW/cm2 at 125 mA/cm2 could be achieved. Following this, Naumi et al. [182] developed a polymer electrolyte membrane fuel cell utilizing BC with sulfonated polystyrene and phosphoric acid (Figure 5C). The addition of phosphoric acid was found to enhance membrane proton conductivity [182]. Phosphorylation has an impact on ion exchange capacity, ionic conductivity, swelling index, and contact angle of the membrane [176]. The phosphorylated BC produced from cassava liquid waste and doped with 20 mmol of H3PO4 leads to an enhancement in proton conductivity, with a maximum conductivity reaching 7.9 × 10−2 S/cm at 80 °C. Furthermore, a membrane electrode assembly was constructed using this membrane, with 60 %w/w Pt/C loading of 0.5 mg/cm2 for both the anode and cathode. A single-cell performance achieves a high power density of 25 mW/cm2 at an operating temperature of 40 °C [176]. A carbon-fiber-embedded BC/polyaniline nanocomposite electrode was developed for microbial fuel cell (MFC) applications, that are promising renewable energy sources to simultaneously treat organic waste and generate electricity [183]. The MFC performance has a maximum current density of 0.009 mA/cm2. The adaptation period of microorganisms using E. coli and biofilm development influenced the initial current density, with an increase observed in the nanocomposite compared to carbon fiber alone. The nanocomposite facilitates electron adhesion and conduction, leading to higher conductivity and improved MFC performance compared to the other electrodes tested, including carbon fiber/BC and BC/polyaniline membranes. The enhanced conductivity of this nanocomposite is attributed to the incorporation of carbon fibers and the doping of BC with polyaniline, promoting electron transfer to the conductive material [183]. The in situ fermentation BC with graphene, reduced graphene oxide, polyaniline, and carbon nanofibers using a novel layer-by-layer self-assembly approach was proposed by Luo et al. [5]. Introduction of graphene dispersion into the BC layers was achieved via a repetitive spraying method, resulting in the formation of robust three-dimensional hydrogels [5,184]. Morphological and Raman spectroscopy analyses confirmed the presence of few-layered graphene nanosheets, enhancing hydrogen bonding interactions within the composite [185]. The incorporation of polyaniline leads to the fabrication of freestanding BC-based electrodes exhibiting good electrochemical properties and high charge storage capacity with the aim of developing BC-based nanocomposites with enhanced properties for potential applications in electrodes and supercapacitors [186,187]. The strategic integration of graphene via sustainable in situ fermentation routes leads to the production of composites with properties comparable to those achieved through chemical-based ex situ modifications [188]. Graphene/BC nanocomposites produced via fermentation exhibited a high capacitance of approximately 215 F/g, with exceptional stability over nearly 5000 cycles [189]. However, the inclusion of carbon-based nanomaterials as fermentation additives in the fermentation media affected microbial growth kinetics and process behavior, which are crucial during scaled-up operations [190].
Figure 5. Energy storage application of chemically modified BC nanocomposites. Silylated BC/hydroxylated carbon nanotube/polyethylene glycol (SBTP) film (A): synthesis of the SBTP film (a), mechanism of solar–thermal energy storage, conversion of solar–thermal energy to electricity (b), and infrared thermograph depicting the phase change storage process of SBTP film under solar irradiation (c) (Reprinted with permission from ref. [172] Copyright © 2023 Wiley). Poly(4-styrene sulfonic acid) (PSSA)/BC membrane electrode (B): fabrication scheme (d) and image showing the pristine BC (e) and PSSA/BC after fuel cell testing with no evident signs of degradation (f), except for the groove indicated by the arrow, possibly caused by the rubber seal (Reprinted with permission from ref. [181] Copyright © 2017 Elsevier). Phosphorylated BC (C): phosphorylation process of BC (Reprinted from ref. [172] open access Copyright © 2023 by the authors. Licensee MDPI).
Figure 5. Energy storage application of chemically modified BC nanocomposites. Silylated BC/hydroxylated carbon nanotube/polyethylene glycol (SBTP) film (A): synthesis of the SBTP film (a), mechanism of solar–thermal energy storage, conversion of solar–thermal energy to electricity (b), and infrared thermograph depicting the phase change storage process of SBTP film under solar irradiation (c) (Reprinted with permission from ref. [172] Copyright © 2023 Wiley). Poly(4-styrene sulfonic acid) (PSSA)/BC membrane electrode (B): fabrication scheme (d) and image showing the pristine BC (e) and PSSA/BC after fuel cell testing with no evident signs of degradation (f), except for the groove indicated by the arrow, possibly caused by the rubber seal (Reprinted with permission from ref. [181] Copyright © 2017 Elsevier). Phosphorylated BC (C): phosphorylation process of BC (Reprinted from ref. [172] open access Copyright © 2023 by the authors. Licensee MDPI).
Fermentation 10 00316 g005
The development of organic light-emitting diodes (OLEDs) typically involves a multilayered structure with an organic semiconductor, placed between two electrodes on a transparent rigid substrate [191]. However, a significant challenge lies in replacing glass with a more flexible substrate to enable OLEDs for use on curved surfaces, which is crucial for the design of wearable devices and real-time communication [192]. As a result, there has been a growing interest in developing flexible and conformable substrates for OLED fabrication [193]. Highly transparent BC (HTBC) biocompatible membranes are developed for use as substrates in OLEDs [194]. These membranes, composed of BC and an organic–inorganic sol containing boehmite nanoparticles and epoxy-modified siloxane, are enhanced with SiO2 and indium tin oxide (ITO) thin films deposited via radio frequency magnetron sputtering. The HTBC/SiO2/ITO substrate exhibited improved visible light transmission by 88%. Electrical characterization reveals that the ITO-deposited films are n-type doped semiconductors with properties comparable to commercial ITO on glass substrates [194]. The blending of polystyrene with BC presents an appealing option for OLED applications due to the comparable refractive indices of these polymers [195]. Utilizing d-limonene, a renewable solvent capable of dissolving polystyrene, has been explored to enhance the transparency and hydrophobic properties of BC, aiming to achieve a flexible substrate for OLED applications [195].

5.3. Filtration Membranes

The utilization of membrane technology has been recognized as an efficient and energy-saving approach for separation processes [196]. However, its widespread adoption on a large scale is impeded by various challenges, encompassing complex operational procedures, high operational costs, concerns regarding toxicity, limited recyclability, and the generation of hazardous by-products. Hence, there is a critical imperative to explore materials that offer cost-effectiveness and environmental compatibility [197]. Addressing some of these challenges, BC-based separation membranes have shown promise. For the separation of emulsions, previous efforts have leveraged the hydrophilicity of inorganic coatings like SiO2 to enhance membrane performance [198]. For instance, Hou et al. demonstrated the deposition of modified SiO2 on dried BC membranes using chlorodimethyloctadecylsilane through hydrolysis and condensation processes (Figure 6A) [198]. These nanofiber network membranes exhibit pore sizes of 0.5–1 μm and show underwater superoleophobicity and under-oil superhydrophobicity properties, enabling the separation of water-in-oil (W/O) emulsions. Similarly, Wahid et al. developed superhydrophilic membranes based on BC by blending commercial dopamine hydrochloride with micro-sized SiO2 particles to fabricate BC-SiO2 composite membranes for oil-in-water (O/W) emulsion separation (Figure 6B) [199]. However, the use of microparticles may lead to increased accumulation of micro-sized particles, potentially posing challenges for waste disposal in terrestrial and aquatic environments, including freshwater, sediments, and soil [200]. Spherical SiO2 nanoparticles are produced through the hydrolysis of tetraethoxysilane in the presence of ammonia and subsequently applied onto the surface of BC membranes [200]. The surface roughness and wettability are enhanced, facilitating the separation of oil-in-water (O/W) emulsions. The straightforward deposition of SiO2 nanoparticles offers a practical method for creating superhydrophilic membranes capable of efficiently separating emulsified oil/water mixtures, achieving separation rates of up to 99% [200]. Moreover, composite of BC produced from banana peel waste and SiO2 nanoparticles was also applied for desalination [201]. The flexible paper-like composite membranes are produced with varying proportions of BC, microcellulose, and SiO2. The membrane demonstrated the high salt rejection value of 4.89% initially, although this decreased over time due to membrane surface clogging [201].
The presence of non-biodegradable, toxic, carcinogenic, mutagenic, and potentially teratogenic dyes and heavy metals poses significant threats to both human health and aquatic ecosystems [202]. Therefore, there is an urgent need to remove metal ions and dyes from wastewater to mitigate their harmful effects [203]. This shows the critical importance and high demand for effective wastewater treatment methods that can address these contaminants and protect environmental and public health. Biosurfactants (BS) derived from Candida lipolytica fermentation and BC, both cultivated in media supplemented with industrial waste, have been applied for treating textile effluents [204]. Using solely BC as a filtration agent achieved up to 45% removal efficiency of navy blue dye, while treatment with BS prior to filtration with BC resulted in 65% color removal [204]. Using coffee as a natural dye to produce eco-friendly coffee-dyed BC (BC-COF) bio-leather, aiming for sustainable textile applications was investigated [205]. The optimal dyeing and conditions were determined through various parameters such as temperature and time. The analysis confirmed the colorization of BC with coffee without altering its chemical and crystalline structures, with coffee molecules effectively incorporating into the BC fiber structures. The BC-COF bio-leather exhibits efficient adsorption of methylene blue dye with multiple reusability, suggesting its potential as a sustainable dye adsorbent [205]. A biocomposite of dialdehyde cellulose nanocrystals (DCNC) and wool keratin is fabricated for the purpose of removing Cd+2 and Crystal violet from water. The removal efficiency of Cd2+ and Crystal violet by the Keratin/DCNC adsorbent slightly decreases after five cycles of recycling [74]. These results show the potential of BC combination for wastewater treatment, offering a cost-effective and sustainable filtration solution.
Metal oxides, particularly magnetite, have been recognized for their ability to improve the adsorption of various contaminants such as metals, textile dyes, organic substances, nitrogen, and phosphorus, while also exhibiting antibacterial properties. Syukri et al. [206] conducted a study to assess the effectiveness of incorporating Fe3O4 into nanocellulose membranes derived from pineapple peel waste for reducing bacteria and dye contaminants in wastewater. The inclusion of Fe3O4 showed promising results in reducing microbial counts and anionic dye concentrations in water samples obtained from local rivers in Padang city, adjacent to industries such as rubber, cement, and tofu production [206]. Similar observations are noted in the utilization of membranes for removing dyes and bacteria carrying negative charges [207]. The impact of incorporating TiO2 nanoparticles and graphene into BC acetate-based nanocomposite membranes designed for bacterial filtration in water was investigated. BC was derived from pineapple peel waste through a fermentation process, followed by a high-pressure homogenization step to reduce nanocellulose size, and an esterification process to produce cellulose acetate. Nanocomposite membranes were then synthesized, reinforced with 1% TiO2 nanoparticles and 1% graphene nanopowder. The addition of TiO2 nanoparticles and graphene to the cellulose acetate membrane resulted in improved filtration efficiency against coliform bacteria in water [207].

5.4. Food Packaging Materials

It is crucial to develop bio-based primary packaging to enhance food safety and quality while also addressing environmental pollution from the growing accumulation of packaging waste [208,209]. Bioplastic food packaging can be made from BC, driven by its biodegradability. BC offers a promising alternative to synthetic plastics commonly used in various industries, including food packaging, due to its lignin-free cellulose, high mechanical strength, and environmentally friendly nature [208,209]. To meet the requirements of food packaging standards, modifications are necessary for pristine BC film due to its non-transparency, lack of antibacterial properties, and typically thicker consistency compared to plastic film [210].
The BC films with enhanced properties are developed through surface modification using cyclic anhydrides and dimethyl sulfoxide as a solvent [211]. Different anhydrides have been explored to optimize the film characteristics. The modification with dodecenyl succinic anhydride increases the dry and wet tensile strength, while octadecenyl succinic anhydride reduces water vapor permeability and exhibits strong antimicrobial effects. Films modified with maleic anhydride have improved oxygen barrier properties and facilitate the effective preservation of strawberries (Figure 7A). The films display complete biodegradability in soil within one month. Additionally, surface grafting with different alkyl moieties (C0–C18) leads to films with varied properties, including mechanical strength, water vapor permeability, and antimicrobial activity [211]. The regenerated BC films are fabricated utilizing citric acid as the crosslinking agent to enhance the mechanical properties of the resulting film [210]. The film exhibits a fracture strength of 93.40 MPa and a Young’s modulus of 4.2 GPa, surpassing the performance of commercial polyvinyl chloride (PVC) plastic film. Moreover, the modified film demonstrated biodegradability in soil within a mere 2 weeks. Furthermore, the citric-acid-crosslinked film exhibits heat resistance and food preservation capabilities compared to PVC plastic wrap [210]. Modified BC films are investigated, incorporating carboxymethyl cellulose (CMC) as a stabilizer and glycerol as a plasticizer [212]. Various composite films were prepared with different concentrations of CMC and glycerol, and their physicochemical and mechanical properties were assessed. A tensile strength of 17.47 MPa, elongation at break of 25.60%, and Young’s modulus of 6.54 GPa were found. The modified BC films were applied as packaging for meat sausages, demonstrating their ability to maintain sausage quality during a 6-day storage period at room temperature (Figure 7B) [212]. Antimicrobial nanocomposite films were developed using polyvinyl alcohol reinforced with glycerol, BC nanocrystals, and boric acid. The film composition optimized by the response surface methodology exhibited reduced water solubility and moisture content, along with enhanced water vapor permeability and water vapor transmission rate. Additionally, the antimicrobial activity of the film against B. subtilis and C. albicans was observed. The optimized film also demonstrated increased biodegradability compared to neat polyvinyl alcohol film. The addition of BCNCs and boric acid enhanced the thermal stability of the film [213]. The BC-based food packaging material that is biocompatible, biodegradable, bioactive, and non-toxic is also developed by reducing silver nitrate with sodium chloride to prepare a BC/Ag nanocomposite [214]. The nanocomposite exhibits improved water solubility (4.6%) and tensile strength (25.7 MPa). It shows no toxicity toward NIH-3T3 fibroblasts and displays antimicrobial activity against various bacterial and fungal strains. BC/Ag-nanocomposite-film-coated oranges maintain acceptable sensory qualities during storage for up to 9 weeks [214].
The incorporation of natural antimicrobial agents into matrix formulations for food packaging was studied. Biodegradable food packaging films are created using potato peel powder, with BC added as a reinforcement agent [215]. Natural extracts from food processing wastes, such as curcumin, were utilized as antioxidant additives. BC interacts with biopolymers in the potato peel matrix primarily through inter- and intra-molecular bonding interactions, leading to enhancements in mechanical properties, as well as reductions in oxygen permeability, water vapor permeability, and light transparency of the films. The addition of curcumin imparts good antioxidant capacity to the potato peel/BC films, effectively inhibiting lipid oxidation in fresh pork during storage [215]. Whey protein concentrate and BC nanowhiskers were investigated as raw materials for edible film production, with a focus on modifying these films with oregano essential oils to impart antimicrobial properties against various foodborne pathogens [40]. The inclusion of BC nanowhiskers in whey protein concentrate/oregano essential oil films improves the mechanical properties of the films [216] and increases their antimicrobial effectiveness, with up to 80% inhibition [40]. Structural analysis confirms the homogenous distribution of the oregano essential oils and BC within the film matrix, resulting in rough films with enhanced release capacity for the oregano essential oils [40]. A dual-action method was used by impregnating BC with citrus pectin (CP) and then functionalizing it with an active emulsion coating of thyme essential oil (TEO) encapsulated in CP-TEO [217]. This film exhibits good physicochemical properties, a continuous and dense microstructure, optimal mechanical strength, and the lowest water vapor permeability. The BC composite film demonstrates antibacterial activity against S. aureus and E. coli, antibiofilm properties, effective UV-blocking, and improves visible light transmittance. It shows superior potential in preserving the quality of grapes over 9 days of storage compared to pristine BC and conventional plastic packaging. During the 9-day storage period, the film maintains the freshness, color, and overall structural integrity of the grape samples while exhibiting minimal weight loss [217]. Moreover, agar/foamed BC incorporating carvacrol, the thyme essential oils, exhibits a gradual release of carvacrol, making it suitable for food packaging for fish. The agar/BC/carvacrol combination inhibits microbial growth and minimizes nitrogen compounds resulting from protein decomposition when microbial spoilage occurs in sea bass fillets after day 6. Additionally, it decelerates lipid peroxidation, thereby prolonging the shelf life of the fish [218]. Since BC possesses unique characteristics such as fibrillated organization, porous structure, and abundant hydrogen bonding sites, BC is also identified as an effective biopolymer for delivering various bioactive compounds such as catechin [219] and cinnamon essential oil [220].
Films based on chitosan/BC have the potential to serve as antimicrobial packaging, enhancing the preservation of packaged foods. An antimicrobial film using dialdehyde BC, chitosan, and ε-polylysine (ε-PL), a cationic natural antimicrobial peptide derived from microbial metabolites, was studied [221]. The antibacterial activity against S. aureus and E. coli was enhanced with increasing ε-PL content. However, the excessive ε-PL tended to aggregate, disrupting intermolecular interactions among polymer chains and consequently reducing tensile strength and water solubility. The optimal formulation was found to exhibit a maximum tensile strength of 36.77 MPa, along with low water vapor permeability and oxygen permeability values. Furthermore, the films effectively extended the shelf life of tilapia fillets during refrigerated storage [221]. A biocomposite consisting of BC nanofibers incorporated into tapioca starch/chitosan-based films was studied [222]. Through ultrasonication, the fillers were uniformly dispersed throughout the film matrix. The addition of the dried BC resulted in an improved tensile strength (4.7 MPa), thermal resistance, moisture resistance (8.9% increase after 8 h), and water vapor barrier (27% improvement after 24 h). Additionally, all the chitosan/BC films exhibited antibacterial activity [222]. The BC and silver nanoparticles were incorporated into chitosan to develop nanocomposite films [223]. The addition of BC and silver nanoparticles enhanced the strength of the chitosan matrix while reducing moisture content, water solubility, and water vapor permeability. Incorporating glycerol and grape bagasse extract (GE) led to higher moisture content due to their hydrophilic nature. The addition of GE increased water vapor permeability. Glycerol and GE incorporation significantly enhanced the swelling capacity and solubility. Using the multilayer method, BC and chitosan compounds exhibited greater swelling capacity and lower solubility compared to compounds developed using the immersion method, where a BC film was immersed in a chitosan solution [224]. The nanocomposite films demonstrated potent antibacterial effects against foodborne pathogens [223] and antioxidant properties [224], which held promise as active packaging materials, capable of extending the shelf life of packaged foods.

6. Conclusions and Challenges

Utilizing biopolymers like cellulose offers a sustainable alternative to petroleum-based polymers, aligning with the UN’s Sustainable Development Goals and addressing environmental concerns. Cellulose-based materials present diverse applications and the potential to reduce carbon emissions. They can be sourced from plants or synthesized by bacteria, with BC offering unique properties. Cost-effective BC production is crucial, with efforts focusing on mainly utilizing waste sugars as culture media. Research interest in BC has grown, highlighting its versatile applications and potential for large-scale production using agro-industrial waste. To improve BC yield, dynamic bioreactor systems have been developed, alongside genetic engineering approaches to modify metabolic pathways. These strategies offer promising avenues for maximizing BC production while minimizing energy consumption and environmental impact. The versatility and sustainability of BC make it promising for addressing contemporary materials science challenges in many applications such as medical materials, energy storage/electronic devices, filtration, and food packaging.
As research in this field progresses, harnessing waste-derived resources for BC production holds significant promise for advancing sustainable bioprocessing practices in cellulose biosynthesis. However, there are some challenges related to efficient and economical production. Understanding the intricate interplay between waste utilization, enzymatic processes, and chemical composition is pivotal for devising efficient strategies to enhance BC production while minimizing costs and environmental impacts. The enhancement of BC production yield and its effective application necessitate further exploration through foundational research, particularly focusing on the bacterial cells employed in BC production. The mechanisms underlying the growth of strains of BC-producing bacteria and the synthesis of BC remain incompletely understood. In-depth investigations into the physiology and metabolic pathways of the BC-producing bacterial strains can offer valuable insights into the optimization of culture conditions, substrate utilization, and the regulation of BC biosynthesis. Additionally, elucidating the genetic determinants and regulatory networks governing BC production can pave the way for genetic engineering approaches aimed at enhancing BC yield and modifying its properties for tailored applications. Moreover, understanding the interactions between the BC-producing bacteria and their environment, including nutrient availability, pH, and temperature, is crucial for optimizing BC production processes and scaling them up for industrial applications. Therefore, further fundamental research efforts are warranted to unlock the full potential of BC as a sustainable biomaterial in diverse fields.

Funding

This review received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Wang, F.; Harindintwali, J.D.; Yuan, Z.; Wang, M.; Wang, F.; Li, S.; Yin, Z.; Huang, L.; Fu, Y.; Li, L.; et al. Technologies and perspectives for achieving carbon neutrality. Innovation 2021, 2, 100180. [Google Scholar] [CrossRef] [PubMed]
  2. de Sousa, F.D.B. The role of plastic concerning the sustainable development goals: The literature point of view. Clean. Responsible Consum. 2021, 3, 100020. [Google Scholar] [CrossRef]
  3. Morrow, R.; Ribul, M.; Eastmond, H.; Lanot, A.; Baurley, S. Bio-Producing Bacterial Cellulose Filaments through Co-Designing with Biological Characteristics. Materials 2023, 16, 4893. [Google Scholar] [CrossRef] [PubMed]
  4. Atta, O.M.; Manan, S.; Shahzad, A.; Ul-Islam, M.; Ullah, M.W.; Yang, G. Biobased materials for active food packaging: A review. Food Hydrocoll. 2022, 125, 107419. [Google Scholar] [CrossRef]
  5. Luo, H.; Xie, J.; Xiong, L.; Zhu, Y.; Yang, Z.; Wan, Y. Fabrication of flexible, ultra-strong, and highly conductive bacterial cellulose-based paper by engineering dispersion of graphene nanosheets. Compos. Part B Eng. 2019, 162, 484–490. [Google Scholar] [CrossRef]
  6. Eslahi, N.; Mahmoodi, A.; Mahmoudi, N.; Zandi, N.; Simchi, A. Processing and Properties of Nanofibrous Bacterial Cellulose-Containing Polymer Composites: A Review of Recent Advances for Biomedical Applications. Polym. Rev. 2020, 60, 144–170. [Google Scholar] [CrossRef]
  7. Verma, J.; Petru, M.; Goel, S. Cellulose based materials to accelerate the transition towards sustainability. Ind. Crops Prod. 2024, 210, 118078. [Google Scholar] [CrossRef]
  8. Camargo, M.S.A.; Cercal, A.P.; Silveira, V.F.; Mancinelli, K.C.B.; Gern, R.M.M.; Garcia, M.C.F.; Apati, G.P.; dos Santos Schneider, A.L.; Pezzin, A.P.T. Evaluation of Wet Bacterial Cellulose Degradation in Different Environmental Conditions. Macromol. Symp. 2020, 394, 2000149. [Google Scholar] [CrossRef]
  9. Gorgieva, S.; Trček, J. Bacterial Cellulose: Production, Modification and Perspectives in Biomedical Applications. Nanomaterials 2019, 9, 1352. [Google Scholar] [CrossRef]
  10. Greser, A.B.; Avcioglu, N.H. Optimization and physicochemical characterization of bacterial cellulose by Komagataeibacter nataicola and Komagataeibacter maltaceti strains isolated from grape, thorn apple and apple vinegars. Arch. Microbiol. 2022, 204, 465. [Google Scholar] [CrossRef]
  11. Cannazza, P.; Rissanen, A.J.; Guizelini, D.; Losoi, P.; Sarlin, E.; Romano, D.; Santala, V.; Mangayil, R. Characterization of Komagataeibacter Isolate Reveals New Prospects in Waste Stream Valorization for Bacterial Cellulose Production. Microorganisms 2021, 9, 2230. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, Y.; Ullah, M.W.; Ul-Islam, M.; Khan, S.; Jang, J.H.; Park, J.K. Self-assembly of bio-cellulose nanofibrils through intermediate phase in a cell-free enzyme system. Biochem. Eng. J. 2019, 142, 135–144. [Google Scholar] [CrossRef]
  13. Zhang, Y.; Chen, Y.; Cao, G.; Ma, X.; Zhou, J.; Xu, W. Bacterial cellulose production from terylene ammonia hydrolysate by Taonella mepensis WT-6. Int. J. Biol. Macromol. 2021, 166, 251–258. [Google Scholar] [CrossRef] [PubMed]
  14. El-Gendi, H.; Salama, A.; El-Fakharany, E.M.; Saleh, A.K. Optimization of bacterial cellulose production from prickly pear peels and its ex situ impregnation with fruit byproducts for antimicrobial and strawberry packaging applications. Carbohydr. Polym. 2023, 302, 120383. [Google Scholar] [CrossRef] [PubMed]
  15. Ammar, G.A.G.; Saleh, A.K.; Taha, T.H.; El-Zawawy, W.K.; Abdel-Fattah, Y.R. Developed applicability of a bacterial cellulose matrix as a gelling substitute for plant tissue culture media. Cellulose 2022, 29, 7883–7900. [Google Scholar] [CrossRef]
  16. Płoska, J.; Garbowska, M.; Klempová, S.; Stasiak-Różańska, L. Obtaining Bacterial Cellulose through Selected Strains of Acetic Acid Bacteria in Classical and Waste Media. Appl. Sci. 2023, 13, 6429. [Google Scholar] [CrossRef]
  17. Sari, A.K.; Majlan, E.H.; Loh, K.S.; Wong, W.Y.; Alva, S.; Khaerudini, D.S.; Yunus, R.M. Effect of acid treatments on thermal properties of bacterial cellulose produced from cassava liquid waste. Mater. Today Proc. 2022, 57, 1174–1178. [Google Scholar] [CrossRef]
  18. Ajkidkarn, P.; Manuspiya, H. Solution plasma synthesis of bacterial cellulose acetate derived from nata de coco waste incorporated with polyether block amide. Int. J. Biol. Macromol. 2022, 209, 1486–1497. [Google Scholar] [CrossRef] [PubMed]
  19. Potočnik, V.; Gorgieva, S.; Trček, J. From Nature to Lab: Sustainable Bacterial Cellulose Production and Modification with Synthetic Biology. Polymers 2023, 15, 3466. [Google Scholar] [CrossRef]
  20. Ul-Islam, M.; Khan, S.; Ullah, M.W.; Park, J.K. Comparative study of plant and bacterial cellulose pellicles regenerated from dissolved states. Int. J. Biol. Macromol. 2019, 137, 247–252. [Google Scholar] [CrossRef]
  21. Taokaew, S.; Thienchaimongkol, J.; Nakson, N.; Kobayashi, T. Valorisation of okara Waste as an Alternative Nitrogen Source in the Biosynthesis of Nanocellulose. Chem. Eng. Trans. 2022, 92, 649–654. [Google Scholar]
  22. Sperotto, G.; Stasiak, L.G.; Godoi, J.P.M.G.; Gabiatti, N.C.; De Souza, S.S. A review of culture media for bacterial cellulose production: Complex, chemically defined and minimal media modulations. Cellulose 2021, 28, 2649–2673. [Google Scholar] [CrossRef]
  23. Fernandes, I.d.A.A.; Pedro, A.C.; Ribeiro, V.R.; Bortolini, D.G.; Ozaki, M.S.C.; Maciel, G.M.; Haminiuk, C.W.I. Bacterial cellulose: From production optimization to new applications. Int. J. Biol. Macromol. 2020, 164, 2598–2611. [Google Scholar] [CrossRef] [PubMed]
  24. Hussain, Z.; Sajjad, W.; Khan, T.; Wahid, F. Production of bacterial cellulose from industrial wastes: A review. Cellulose 2019, 26, 2895–2911. [Google Scholar] [CrossRef]
  25. Katyal, M.; Singh, R.; Mahajan, R.; Sharma, A.; Gupta, R.; Aggarwal, N.K.; Yadav, A. Bacterial cellulose: Nature’s greener tool for industries. Biotechnol. Appl. Biochem. 2023, 70, 1629–1640. [Google Scholar] [CrossRef]
  26. Patel, A.; Patel, P.; Shukla, A.; Wong, J.W.C.; Varjani, S.; Gosai, H. Sustainable Bioconversion of Industrial Wastes into Bacterial Cellulose for Diverse Applications: A Way Towards Pollution Control and Abatement. Curr. Pollut. Rep. 2023, 9, 226–242. [Google Scholar] [CrossRef]
  27. Filippi, K.; Georgaka, N.; Alexandri, M.; Papapostolou, H.; Koutinas, A. Valorisation of grape stalks and pomace for the production of bio-based succinic acid by Actinobacillus succinogenes. Ind. Crops Prod. 2021, 168, 113578. [Google Scholar] [CrossRef]
  28. Bharathiraja, B.; Iyyappan, J.; Jayamuthunagai, J.; Kumar, R.P.; Sirohi, R.; Gnansounou, E.; Pandey, A. Critical review on bioconversion of winery wastes into value-added products. Ind. Crops Prod. 2020, 158, 112954. [Google Scholar] [CrossRef]
  29. Arun, K.B.; Madhavan, A.; Sindhu, R.; Binod, P.; Pandey, A.; R, R.; Sirohi, R. Remodeling agro-industrial and food wastes into value-added bioactives and biopolymers. Ind. Crops Prod. 2020, 154, 112621. [Google Scholar] [CrossRef]
  30. Nirmal, N.P.; Khanashyam, A.C.; Mundanat, A.S.; Shah, K.; Babu, K.S.; Thorakkattu, P.; Al-Asmari, F.; Pandiselvam, R. Valorization of Fruit Waste for Bioactive Compounds and Their Applications in the Food Industry. Foods 2023, 12, 556. [Google Scholar] [CrossRef]
  31. Islam, M.U.; Ullah, M.W.; Khan, S.; Shah, N.; Park, J.K. Strategies for cost-effective and enhanced production of bacterial cellulose. Int. J. Biol. Macromol. 2017, 102, 1166–1173. [Google Scholar] [CrossRef] [PubMed]
  32. Gorgieva, S.; Jančič, U.; Cepec, E.; Trček, J. Production efficiency and properties of bacterial cellulose membranes in a novel grape pomace hydrolysate by Komagataeibacter melomenusus AV436T and Komagataeibacter xylinus LMG 1518. Int. J. Biol. Macromol. 2023, 244, 125368. [Google Scholar] [CrossRef]
  33. Ogrizek, L.; Lamovšek, J.; Čuš, F.; Leskovšek, M.; Gorjanc, M. Properties of Bacterial Cellulose Produced Using White and Red Grape Bagasse as a Nutrient Source. Processes 2021, 9, 1088. [Google Scholar] [CrossRef]
  34. Cerrutti, P.; Roldán, P.; García, R.M.; Galvagno, M.A.; Vázquez, A.; Foresti, M.L. Production of bacterial nanocellulose from wine industry residues: Importance of fermentation time on pellicle characteristics. J. Appl. Polym. Sci. 2016, 133. [Google Scholar] [CrossRef]
  35. Niculescu, V.-C.; Ionete, R.-E. An Overview on Management and Valorisation of Winery Wastes. Appl. Sci. 2023, 13, 5063. [Google Scholar] [CrossRef]
  36. Cazón, P.; Puertas, G.; Vázquez, M. Production and Characterization of Active Bacterial Cellulose Films Obtained from the Fermentation of Wine Bagasse and Discarded Potatoes by Komagateibacter xylinus. Polymers 2022, 14, 5194. [Google Scholar] [CrossRef] [PubMed]
  37. Abdelraof, M.; Hasanin, M.S.; El-Saied, H. Ecofriendly green conversion of potato peel wastes to high productivity bacterial cellulose. Carbohydr. Polym. 2019, 211, 75–83. [Google Scholar] [CrossRef] [PubMed]
  38. Ciecholewska-Juśko, D.; Broda, M.; Żywicka, A.; Styburski, D.; Sobolewski, P.; Gorący, K.; Migdał, P.; Junka, A.; Fijałkowski, K. Potato Juice, a Starch Industry Waste, as a Cost-Effective Medium for the Biosynthesis of Bacterial Cellulose. Int. J. Mol. Sci. 2021, 22, 10807. [Google Scholar] [CrossRef] [PubMed]
  39. Long, L.; Chen, J.; Cao, Y.; Huang, C.; Feng, S.; Yang, H.; Tian, D. Valorization of straw biomass into lignin bio-ink, bacterial cellulose and activated nanocarbon through the trade-off alkali-catalyzed glycerol organosolv biorefinery. Chem. Eng. J. 2024, 484, 149549. [Google Scholar] [CrossRef]
  40. Papadaki, A.; Lappa, I.K.; Manikas, A.C.; Pastore Carbone, M.G.; Natsia, A.; Kachrimanidou, V.; Kopsahelis, N. Grafting bacterial cellulose nanowhiskers into whey protein/essential oil film composites: Effect on structure, essential oil release and antibacterial properties of films. Food Hydrocoll. 2024, 147, 109374. [Google Scholar] [CrossRef]
  41. Huang, Y.; Yang, H.; Yu, Y.; Li, H.; Li, H.; Bai, J.; Shi, F.; Liu, J. Bacterial cellulose biomass aerogels for oil-water separation and thermal insulation. J. Environ. Chem. Eng. 2023, 11, 110403. [Google Scholar] [CrossRef]
  42. Karanicola, P.; Patsalou, M.; Stergiou, P.-Y.; Kavallieratou, A.; Evripidou, N.; Christou, P.; Panagiotou, G.; Damianou, C.; Papamichael, E.M.; Koutinas, M. Ultrasound-assisted dilute acid hydrolysis for production of essential oils, pectin and bacterial cellulose via a citrus processing waste biorefinery. Bioresour. Technol. 2021, 342, 126010. [Google Scholar] [CrossRef] [PubMed]
  43. Tsouko, E.; Maina, S.; Ladakis, D.; Kookos, I.K.; Koutinas, A. Integrated biorefinery development for the extraction of value-added components and bacterial cellulose production from orange peel waste streams. Renew. Energy 2020, 160, 944–954. [Google Scholar] [CrossRef]
  44. Kuo, C.-H.; Huang, C.-Y.; Shieh, C.-J.; Wang, H.-M.D.; Tseng, C.-Y. Hydrolysis of Orange Peel with Cellulase and Pectinase to Produce Bacterial Cellulose using Gluconacetobacter xylinus. Waste Biomass Valorization 2019, 10, 85–93. [Google Scholar] [CrossRef]
  45. Le, H.V.; Dao, N.T.; Bui, H.T.; Kim Le, P.T.; Le, K.A.; Tuong Tran, A.T.; Nguyen, K.D.; Mai Nguyen, H.H.; Ho, P.H. Bacterial Cellulose Aerogels Derived from Pineapple Peel Waste for the Adsorption of Dyes. ACS Omega 2023, 8, 33412–33425. [Google Scholar] [CrossRef] [PubMed]
  46. Patel, A.; Patel, P.; Parmar, M.; Gosai, H. Employing RSM and ANN-based applications for modelling enhanced bacterial cellulose production from pineapple peel waste using Komagateibacter saccharivorans APPK1. Chem. Eng. J. 2024, 480, 148057. [Google Scholar] [CrossRef]
  47. Anwar, B.; Bundjali, B.; Sunarya, Y.; Arcana, I.M. Properties of Bacterial Cellulose and Its Nanocrystalline Obtained from Pineapple Peel Waste Juice. Fibers Polym. 2021, 22, 1228–1236. [Google Scholar] [CrossRef]
  48. Mardawati, E.; Rahmah, D.M.; Rachmadona, N.; Saharina, E.; Pertiwi, T.Y.R.; Zahrad, S.A.; Ramdhani, W.; Srikandace, Y.; Ratnaningrum, D.; Endah, E.S.; et al. Pineapple core from the canning industrial waste for bacterial cellulose production by Komagataeibacter xylinus. Heliyon 2023, 9, e22010. [Google Scholar] [CrossRef] [PubMed]
  49. Hasanin, M.S.; Abdelraof, M.; Hashem, A.H.; El Saied, H. Sustainable bacterial cellulose production by Achromobacter using mango peel waste. Microb. Cell Factories 2023, 22, 24. [Google Scholar] [CrossRef]
  50. García-Sánchez, M.E.; Robledo-Ortíz, J.R.; Jiménez-Palomar, I.; González-Reynoso, O.; González-García, Y. Production of bacterial cellulose by Komagataeibacter xylinus using mango waste as alternative culture medium. Rev. Mex. Ing. Quim. 2020, 19, 851–865. [Google Scholar] [CrossRef]
  51. Nasharudin, M.I.H.; Mahmud, N.; Rahim, M.H.A. Watermelon waste as a growth media substitute for bacterial cellulose production. AIP Conf. Proc. 2024, 3023, 020001. [Google Scholar]
  52. Quiñones-Cerna, C.; Rodríguez-Soto, J.C.; Barraza-Jáuregui, G.; Huanes-Carranza, J.; Cruz-Monzón, J.A.; Ugarte-López, W.; Hurtado-Butrón, F.; Samanamud-Moreno, F.; Haro-Carranza, D.; Valdivieso-Moreno, S.; et al. Bioconversion of Agroindustrial Asparagus Waste into Bacterial Cellulose by Komagataeibacter rhaeticus. Sustainability 2024, 16, 736. [Google Scholar] [CrossRef]
  53. Pandey, A.; Singh, A.; Singh, M.K. Novel low-cost green method for production bacterial cellulose. Polym. Bull. 2023, 81, 6721–6741. [Google Scholar] [CrossRef]
  54. Avcioglu, N.H. Eco-friendly Production of Bacterial Cellulose with Komagataeibacter intermedius Strain by Using Jasminum sambac and Camellia sinensis Plants. J. Polym. Environ. 2024, 32, 460–477. [Google Scholar] [CrossRef]
  55. Agüero, A.; Lascano, D.; Ivorra-Martinez, J.; Gómez-Caturla, J.; Arrieta, M.P.; Balart, R. Use of bacterial cellulose obtained from kombucha fermentation in spent coffee grounds for active composites based on PLA and maleinized linseed oil. Ind. Crops Prod. 2023, 202, 116971. [Google Scholar] [CrossRef]
  56. De Souza, K.C.; Trindade, N.M.; De Amorim, J.D.P.; Do Nascimento, H.A.; Costa, A.F.S.; Henrique, M.A.; Caetano, V.F.; Sarubbo, L.A.; Vinhas, G.M. Kinetic study of a bacterial cellulose production by komagataeibacter rhaeticus using coffee grounds and sugarcane molasses. Mater. Res. 2021, 24, e202000454. [Google Scholar] [CrossRef]
  57. Singhaboot, P.; Phanomarpornchai, A.; Phuangsiri, C.; Boonthongtho, K.; Kroeksakul, P. The Potential of Liquid Waste from the Fruit Preserves Production Process as a Low-cost Raw Material for the Production of Bacterial Cellulose. Pertanika J. Trop. Agric. Sci. 2022, 45, 1125–1136. [Google Scholar] [CrossRef]
  58. Sar, T.; Yesilcimen Akbas, M. Potential use of olive oil mill wastewater for bacterial cellulose production. Bioengineered 2022, 13, 7659–7669. [Google Scholar] [CrossRef]
  59. Voon, W.W.Y.; Muhialdin, B.J.; Yusof, N.L.; Rukayadi, Y.; Meor Hussin, A.S. Bio-cellulose Production by Beijerinckia fluminensis WAUPM53 and Gluconacetobacter xylinus 0416 in Sago By-product Medium. Appl. Biochem. Biotechnol. 2019, 187, 211–220. [Google Scholar] [CrossRef]
  60. Yanti, N.A.; Ahmad, S.W.; Muhiddin, N.H.; Ramadhan, L.O.A.N.; Walhidayah, T. Characterization of bacterial cellulose produced by acetobacter xylinum strain lkn6 using sago liquid waste as nutrient source. Pak. J. Biol. Sci. 2021, 24, 335–344. [Google Scholar] [CrossRef]
  61. Dórame-Miranda, R.F.; Gámez-Meza, N.; Medina-Juárez, L.Á.; Ezquerra-Brauer, J.M.; Ovando-Martínez, M.; Lizardi-Mendoza, J. Bacterial cellulose production by Gluconacetobacter entanii using pecan nutshell as carbon source and its chemical functionalization. Carbohydr. Polym. 2019, 207, 91–99. [Google Scholar] [CrossRef] [PubMed]
  62. Guimarães, D.T.; de Oliveira Barros, M.; de Araújo e Silva, R.; Silva, S.M.F.; de Almeida, J.S.; de Freitas Rosa, M.; Gonçalves, L.R.B.; Brígida, A.I.S. Superabsorbent bacterial cellulose film produced from industrial residue of cashew apple juice processing. Int. J. Biol. Macromol. 2023, 242, 124405. [Google Scholar] [CrossRef] [PubMed]
  63. Tsouko, E.; Pilafidis, S.; Dimopoulou, M.; Kourmentza, K.; Sarris, D. Bioconversion of underutilized brewing by-products into bacterial cellulose by a newly isolated Komagataeibacter rhaeticus strain: A preliminary evaluation of the bioprocess environmental impact. Bioresour. Technol. 2023, 387, 129667. [Google Scholar] [CrossRef] [PubMed]
  64. Heydorn, R.L.; Lammers, D.; Gottschling, M.; Dohnt, K. Effect of food industry by-products on bacterial cellulose production and its structural properties. Cellulose 2023, 30, 4159–4179. [Google Scholar] [CrossRef]
  65. Unlu, F.; Boran, F.; Yesilada, O.; Koytepe, S. Laccase immobilization on bacterial cellulose produced in a mulberry pomace waste extract medium: Characterization and use for dye decolorization. J. Appl. Polym. Sci. 2023, 140, e53952. [Google Scholar] [CrossRef]
  66. Salari, M.; Sowti Khiabani, M.; Rezaei Mokarram, R.; Ghanbarzadeh, B.; Samadi Kafil, H. Preparation and characterization of cellulose nanocrystals from bacterial cellulose produced in sugar beet molasses and cheese whey media. Int. J. Biol. Macromol. 2019, 122, 280–288. [Google Scholar] [CrossRef] [PubMed]
  67. Apriyana, A.Y.; Andriani, D.; Karina, M. Production of bacterial cellulose from tofu liquid waste and rice-washed water: Morphological property and its functional groups analysis. IOP Conf. Ser. Earth Environ. Sci. 2020, 483, 012005. [Google Scholar] [CrossRef]
  68. Taokaew, S.; Nakson, N.; Thienchaimongkol, J.; Kobayashi, T. Enhanced production of fibrous bacterial cellulose in Gluconacetobacter xylinus culture medium containing modified protein of okara waste. J. Biosci. Bioeng. 2023, 135, 71–78. [Google Scholar] [CrossRef]
  69. Ho Jin, Y.; Lee, T.; Kim, J.R.; Choi, Y.-E.; Park, C. Improved production of bacterial cellulose from waste glycerol through investigation of inhibitory effects of crude glycerol-derived compounds by Gluconacetobacter xylinus. J. Ind. Eng. Chem. 2019, 75, 158–163. [Google Scholar] [CrossRef]
  70. Ye, J.; Zheng, S.; Zhang, Z.; Yang, F.; Ma, K.; Feng, Y.; Zheng, J.; Mao, D.; Yang, X. Bacterial cellulose production by Acetobacter xylinum ATCC 23767 using tobacco waste extract as culture medium. Bioresour. Technol. 2019, 274, 518–524. [Google Scholar] [CrossRef]
  71. Li, Z.-Y.; Azi, F.; Ge, Z.-W.; Liu, Y.-F.; Yin, X.-T.; Dong, M.-S. Bio-conversion of kitchen waste into bacterial cellulose using a new multiple carbon utilizing Komagataeibacter rhaeticus: Fermentation profiles and genome-wide analysis. Int. J. Biol. Macromol. 2021, 191, 211–221. [Google Scholar] [CrossRef]
  72. Wang, Q.; Tian, D.; Hu, J.; Huang, M.; Shen, F.; Zeng, Y.; Yang, G.; Zhang, Y.; He, J. Harvesting Bacterial Cellulose from Kitchen Waste to Prepare Superhydrophobic Aerogel for Recovering Waste Cooking Oil toward a Closed-Loop Biorefinery. ACS Sustain. Chem. Eng. 2020, 8, 13400–13407. [Google Scholar] [CrossRef]
  73. Nguyen Ngo, T.T.; Phan, T.H.; Thong Le, T.M.; Tu Le, T.N.; Huynh, Q.; Trang Phan, T.P.; Hoang, M.; Vo, T.P.; Nguyen, D.Q. Producing bacterial cellulose from industrial recycling paper waste sludge. Heliyon 2023, 9, e17663. [Google Scholar] [CrossRef]
  74. Peiravi-Rivash, O.; Mashreghi, M.; Baigenzhenov, O.; Hosseini-Bandegharaei, A. Producing bacterial nano-cellulose and keratin from wastes to synthesize keratin/cellulose nanobiocomposite for removal of dyes and heavy metal ions from waters and wastewaters. Colloids Surf. A Physicochem. Eng. Asp. 2023, 656, 130355. [Google Scholar] [CrossRef]
  75. Garmaroody, E.R.; Jafarzadeh, A.E.; Kermanian, H.; Ramezani, O. Spent black liquor as an alternative carbon source for the synthesis of bacterial cellulose. Cellul. Chem. Technol. 2022, 56, 749–756. [Google Scholar] [CrossRef]
  76. Zhou, J.; Chen, Y.; Zhang, Y.; Sun, S.; Ullah, M.W.; Xu, W. Biotransformation of nylon-6,6 hydrolysate to bacterial cellulose. Green Chem. 2021, 23, 7805–7815. [Google Scholar] [CrossRef]
  77. Ojo, A.O. An Overview of Lignocellulose and Its Biotechnological Importance in High-Value Product Production. Fermentation 2023, 9, 990. [Google Scholar] [CrossRef]
  78. Steinbach, D.; Kruse, A.; Sauer, J. Pretreatment technologies of lignocellulosic biomass in water in view of furfural and 5-hydroxymethylfurfural production- A review. Biomass Convers. Biorefinery 2017, 7, 247–274. [Google Scholar] [CrossRef]
  79. 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]
  80. Arminda, M.; Josúe, C.; Cristina, D.; Fabiana, S.; Yolanda, M. Use of activated carbons for detoxification of a lignocellulosic hydrolysate: Statistical optimisation. J. Environ. Manag. 2021, 296, 113320. [Google Scholar] [CrossRef]
  81. 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]
  82. Santoso, S.P.; Lin, S.-P.; Wang, T.-Y.; Ting, Y.; Hsieh, C.-W.; Yu, R.-C.; Angkawijaya, A.E.; Soetaredjo, F.E.; Hsu, H.-Y.; Cheng, K.-C. Atmospheric cold plasma-assisted pineapple peel waste hydrolysate detoxification for the production of bacterial cellulose. Int. J. Biol. Macromol. 2021, 175, 526–534. [Google Scholar] [CrossRef]
  83. Yang, X.-Y.; Huang, C.; Guo, H.-J.; Xiong, L.; Luo, J.; Wang, B.; Lin, X.-Q.; Chen, X.-F.; Chen, X.-D. Bacterial cellulose production from the litchi extract by Gluconacetobacter xylinus. Prep. Biochem. Biotechnol. 2016, 46, 39–43. [Google Scholar] [CrossRef]
  84. Panesar, P.S.; Chavan, Y.; Chopra, H.K.; Kennedy, J.F. Production of microbial cellulose: Response surface methodology approach. Carbohydr. Polym. 2012, 87, 930–934. [Google Scholar] [CrossRef]
  85. Fan, X.; Gao, Y.; He, W.; Hu, H.; Tian, M.; Wang, K.; Pan, S. Production of nano bacterial cellulose from beverage industrial waste of citrus peel and pomace using Komagataeibacter xylinus. Carbohydr. Polym. 2016, 151, 1068–1072. [Google Scholar] [CrossRef]
  86. Santoso, S.P.; Chou, C.-C.; Lin, S.-P.; Soetaredjo, F.E.; Ismadji, S.; Hsieh, C.-W.; Cheng, K.C. Enhanced production of bacterial cellulose by Komactobacter intermedius using statistical modeling. Cellulose 2020, 27, 2497–2509. [Google Scholar] [CrossRef]
  87. Nie, W.; Zheng, X.; Feng, W.; Liu, Y.; Li, Y.; Liang, X. Characterization of bacterial cellulose produced by Acetobacter pasteurianus MGC-N8819 utilizing lotus rhizome. LWT 2022, 165, 113763. [Google Scholar] [CrossRef]
  88. Amândio, M.S.T.; Rocha, J.M.S.; Xavier, A.M.R.B. Enzymatic Hydrolysis Strategies for Cellulosic Sugars Production to Obtain Bioethanol from Eucalyptus globulus Bark. Fermentation 2023, 9, 241. [Google Scholar] [CrossRef]
  89. Efthymiou, M.-N.; Tsouko, E.; Pateraki, C.; Papagiannopoulos, A.; Tzamalis, P.; Pispas, S.; Bethanis, K.; Mantala, I.; Koutinas, A. Property evaluation of bacterial cellulose nanostructures produced from confectionery wastes. Biochem. Eng. J. 2022, 186, 108575. [Google Scholar] [CrossRef]
  90. Sadalage, P.S.; Pawar, K.D. Production of microcrystalline cellulose and bacterial nanocellulose through biological valorization of lignocellulosic biomass wastes. J. Clean. Prod. 2021, 327, 129462. [Google Scholar] [CrossRef]
  91. Yan, S.; Xu, Y.; Yu, X.-W. Rational engineering of xylanase hyper-producing system in Trichoderma reesei for efficient biomass degradation. Biotechnol. Biofuels 2021, 14, 90. [Google Scholar] [CrossRef]
  92. Saleh, A.K.; Salama, A.; Badawy, A.S.; Diab, M.A.; El-Gendi, H. Paper sludge saccharification for batch and fed-batch production of bacterial cellulose decorated with magnetite for dye decolorization by experimental design. Cellulose 2023, 30, 10841–10866. [Google Scholar] [CrossRef]
  93. Zhou, J.; Gu, F.; Yang, H.; Cao, G.; Xu, W.; Sun, S.; Zhang, Y. Sequential fermentation strategy improves microbial conversion of waste jasmine flower to bacterial cellulose with antibacterial properties. Ind. Crops Prod. 2022, 185, 115147. [Google Scholar] [CrossRef]
  94. Duncan, S.M.; Alkasrawi, M.; Gurram, R.; Almomani, F.; Wiberley-Bradford, A.E.; Singsaas, E. Paper Mill Sludge as a Source of Sugars for Use in the Production of Bioethanol and Isoprene. Energies 2020, 13, 4662. [Google Scholar] [CrossRef]
  95. Saleh, A.K.; El-Gendi, H.; Ray, J.B.; Taha, T.H. A low-cost effective media from starch kitchen waste for bacterial cellulose production and its application as simultaneous absorbance for methylene blue dye removal. Biomass Convers. Biorefinery 2023, 13, 12437–12449. [Google Scholar] [CrossRef]
  96. El-Bestawy, E.; Eltaweil, A.S.; Khallaf, N.S. Effective production of bacterial cellulose using acidic dairy industry by-products and agro wastes. Sustain. Chem. Pharm. 2023, 33, 101064. [Google Scholar] [CrossRef]
  97. Ganta, A.; Bashir, Y.; Das, S. Dairy Wastewater as a Potential Feedstock for Valuable Production with Concurrent Wastewater Treatment through Microbial Electrochemical Technologies. Energies 2022, 15, 9084. [Google Scholar] [CrossRef]
  98. Astuti, E.N.J.; Nugroho, D.A.; Ahmadi, T.P. Application of real-time image processing for monitoring bacterial cellulose growth in various nitrogen sources using soybean-boiled wastewater medium during fermentation. IOP Conf. Ser. Earth Environ. Sci. 2023, 1183, 012064. [Google Scholar] [CrossRef]
  99. Jin Chung, W.; Shim, J.; Ravindran, B. Characterization of cheese processed wastewater and treatment using calcium nanoparticles synthesised by Senna auriculata L flower extract. J. King Saud Univ. Sci. 2022, 34, 101793. [Google Scholar] [CrossRef]
  100. Lotfy, V.F.; Basta, A.H.; Abdel-Monem, M.O.; Abdel-Hamed, G.Z. Utilization of bacteria in rotten Guava for production of bacterial cellulose from isolated and protein waste. Carbohydr. Polym. Technol. Appl. 2021, 2, 100076. [Google Scholar] [CrossRef]
  101. Taokaew, S.; Zhang, X.; Chuenkaek, T.; Kobayashi, T. Chitin from fermentative extraction of crab shells using okara as a nutrient source and comparative analysis of structural differences from chemically extracted chitin. Biochem. Eng. J. 2020, 159, 107588. [Google Scholar] [CrossRef]
  102. Skiba, E.A.; Shavyrkina, N.A.; Budaeva, V.V.; Sitnikova, A.E.; Korchagina, A.A.; Bychin, N.V.; Gladysheva, E.K.; Pavlov, I.N.; Zharikov, A.N.; Lubyansky, V.G.; et al. Biosynthesis of Bacterial Cellulose by Extended Cultivation with Multiple Removal of BC Pellicles. Polymers 2021, 13, 2118. [Google Scholar] [CrossRef] [PubMed]
  103. Shavyrkina, N.A.; Skiba, E.A.; Kazantseva, A.E.; Gladysheva, E.K.; Budaeva, V.V.; Bychin, N.V.; Gismatulina, Y.A.; Kashcheyeva, E.I.; Mironova, G.F.; Korchagina, A.A.; et al. Static Culture Combined with Aeration in Biosynthesis of Bacterial Cellulose. Polymers 2021, 13, 4241. [Google Scholar] [CrossRef] [PubMed]
  104. Aswini, K.; Gopal, N.O.; Uthandi, S. Optimized culture conditions for bacterial cellulose production by Acetobacter senegalensis MA1. BMC Biotechnol. 2020, 20, 46. [Google Scholar] [CrossRef] [PubMed]
  105. Bagherniya, M.; Babaeipour, V.; Soleimani, A. Optimization of Bacterial Nano-Cellulose Production in Bench-Scale Rotating Biological Contact Bioreactor by Response Surface Methodology. Iran. J. Chem. Chem. Eng. 2021, 40, 407–416. [Google Scholar]
  106. Kumar, V.; Sharma, D.K.; Bansal, V.; Mehta, D.; Sangwan, R.S.; Yadav, S.K. Efficient and economic process for the production of bacterial cellulose from isolated strain of Acetobacter pasteurianus of RSV-4 bacterium. Bioresour. Technol. 2019, 275, 430–433. [Google Scholar] [CrossRef] [PubMed]
  107. Kim, J.-H.; Han, K.-A. Optimization of bacterial cellulose production from alcohol lees by intermittent feeding strategy. Braz. J. Chem. Eng. 2023, 40, 685–694. [Google Scholar] [CrossRef]
  108. Adnan, A.; Nair, G.; Lay, M.; Swan, J. Bacterial Cellulose Synthesis by Gluconacetobacter xylinus: Enhancement via Fed-batch Fermentation Strategies in Glycerol Media. Trends Sci. 2021, 18, 453. [Google Scholar] [CrossRef]
  109. Hosseinzadeh, A.; Chegeni, A.; Babaeipour, V. Mathematical modeling and simulation of oxygen mass transfer in rotating biological contactor (RBC) for bacterial cellulose production. Biochem. Eng. J. 2023, 200, 109076. [Google Scholar] [CrossRef]
  110. Waqas, S.; Bilad, M.R.; Aqsha, A.; Yub Harun, N.; Ayoub, M.; Wirzal, M.D.H.; Jaafar, J.; Mulyati, S.; Elma, M. Effect of membrane properties in a membrane rotating biological contactor for wastewater treatment. J. Environ. Chem. Eng. 2021, 9, 104869. [Google Scholar] [CrossRef]
  111. Han, Y.; Ma, J.; Xiao, B.; Huo, X.; Guo, X. New Integrated Self-Refluxing Rotating Biological Contactor for rural sewage treatment. J. Clean. Prod. 2019, 217, 324–334. [Google Scholar] [CrossRef]
  112. Soleimani, A.; Hamedi, S.; Babaeipour, V.; Rouhi, M. Design, construction and optimization a flexible bench-scale rotating biological contactor (RBC) for enhanced production of bacterial cellulose by Acetobacter Xylinium. Bioprocess Biosyst. Eng. 2021, 44, 1071–1080. [Google Scholar] [CrossRef]
  113. Vázquez, M.; Puertas, G.; Cazón, P. Processing of Grape Bagasse and Potato Wastes for the Co-Production of Bacterial Cellulose and Gluconic Acid in an Airlift Bioreactor. Polymers 2023, 15, 3944. [Google Scholar] [CrossRef]
  114. Kthiri, A.; Hidouri, S.; Wiem, T.; Jeridi, R.; Sheehan, D.; Landouls, A. Biochemical and biomolecular effects induced by a static magnetic field in Saccharomyces cerevisiae: Evidence for oxidative stress. PLoS ONE 2019, 14, e0209843. [Google Scholar] [CrossRef]
  115. Żywicka, A.; Ciecholewska-Juśko, D.; Drozd, R.; Rakoczy, R.; Konopacki, M.; Kordas, M.; Junka, A.; Migdał, P.; Fijałkowski, K. Preparation of Komagataeibacter xylinus Inoculum for Bacterial Cellulose Biosynthesis Using Magnetically Assisted External-Loop Airlift Bioreactor. Polymers 2021, 13, 3950. [Google Scholar] [CrossRef]
  116. Konopacki, M.; Grygorcewicz, B.; Kordas, M.; Ossowicz-Rupniewska, P.; Nowak, A.; Perużyńska, M.; Rakoczy, R. Intensification of bacterial cellulose production process with sequential electromagnetic field exposure aided by dynamic modelling. Biochem. Eng. J. 2022, 182, 108432. [Google Scholar] [CrossRef]
  117. Drozd, R.; Rakoczy, R.; Wasak, A.; Junka, A.; Fijałkowski, K. The application of magnetically modified bacterial cellulose for immobilization of laccase. Int. J. Biol. Macromol. 2018, 108, 462–470. [Google Scholar] [CrossRef]
  118. Drozd, R.; Szymańska, M.; Żywicka, A.; Kowalska, U.; Rakoczy, R.; Kordas, M.; Konopacki, M.; Junka, A.F.; Fijałkowski, K. Exposure to non-continuous rotating magnetic field induces metabolic strain-specific response of Komagataeibacter xylinus. Biochem. Eng. J. 2021, 166, 107855. [Google Scholar] [CrossRef]
  119. Fijałkowski, K.; Rakoczy, R.; Żywicka, A.; Drozd, R.; Zielińska, B.; Wenelska, K.; Cendrowski, K.; Peitler, D.; Kordas, M.; Konopacki, M.; et al. Time Dependent Influence of Rotating Magnetic Field on Bacterial Cellulose. Int. J. Polym. Sci. 2016, 2016, 7536397. [Google Scholar] [CrossRef]
  120. Lechowska, J.; Kordas, M.; Konopacki, M.; Fijałkowski, K.; Drozd, R.; Rakoczy, R. Hydrodynamic studies in magnetically assisted external-loop airlift reactor. Chem. Eng. J. 2019, 362, 298–309. [Google Scholar] [CrossRef]
  121. Gullo, M.; La China, S.; Petroni, G.; Di Gregorio, S.; Giudici, P. Exploring K2G30 Genome: A High Bacterial Cellulose Producing Strain in Glucose and Mannitol Based Media. Front. Microbiol. 2019, 10, 58. [Google Scholar] [CrossRef]
  122. Ryngajłło, M.; Jędrzejczak-Krzepkowska, M.; Kubiak, K.; Ludwicka, K.; Bielecki, S. Towards control of cellulose biosynthesis by Komagataeibacter using systems-level and strain engineering strategies: Current progress and perspectives. Appl. Microbiol. Biotechnol. 2020, 104, 6565–6585. [Google Scholar] [CrossRef]
  123. Singh, A.; Walker, K.T.; Ledesma-Amaro, R.; Ellis, T. Engineering Bacterial Cellulose by Synthetic Biology. Int. J. Mol. Sci. 2020, 21, 9185. [Google Scholar] [CrossRef]
  124. Anguluri, K.; La China, S.; Brugnoli, M.; Cassanelli, S.; Gullo, M. Better under stress: Improving bacterial cellulose production by Komagataeibacter xylinus K2G30 (UMCC 2756) using adaptive laboratory evolution. Front. Microbiol. 2022, 13, 994097. [Google Scholar] [CrossRef]
  125. Jacek, P.; Dourado, F.; Gama, M.; Bielecki, S. Molecular aspects of bacterial nanocellulose biosynthesis. Microb. Biotechnol. 2019, 12, 633–649. [Google Scholar] [CrossRef]
  126. Liu, M.; Li, S.; Xie, Y.; Jia, S.; Hou, Y.; Zou, Y.; Zhong, C. Enhanced bacterial cellulose production by Gluconacetobacter xylinus via expression of Vitreoscilla hemoglobin and oxygen tension regulation. Appl. Microbiol. Biotechnol. 2018, 102, 1155–1165. [Google Scholar] [CrossRef]
  127. Jang, W.D.; Kim, T.Y.; Kim, H.U.; Shim, W.Y.; Ryu, J.Y.; Park, J.H.; Lee, S.Y. Genomic and metabolic analysis of Komagataeibacter xylinus DSM 2325 producing bacterial cellulose nanofiber. Biotechnol. Bioeng. 2019, 116, 3372–3381. [Google Scholar] [CrossRef]
  128. Yang, F.; Cao, Z.; Li, C.; Chen, L.; Wu, G.; Zhou, X.; Hong, F.F. A recombinant strain of Komagataeibacter xylinus ATCC 23770 for production of bacterial cellulose from mannose-rich resources. New Biotechnol. 2023, 76, 72–81. [Google Scholar] [CrossRef]
  129. Lu, T.; Gao, H.; Liao, B.; Wu, J.; Zhang, W.; Huang, J.; Liu, M.; Huang, J.; Chang, Z.; Jin, M.; et al. Characterization and optimization of production of bacterial cellulose from strain CGMCC 17276 based on whole-genome analysis. Carbohydr. Polym. 2020, 232, 115788. [Google Scholar] [CrossRef]
  130. Summary for Policymakers. In Climate Change 2022—Impacts, Adaptation and Vulnerability: Working Group II Contribution to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Intergovernmental Panel on Climate Change (Ed.) Cambridge University Press: Cambridge, UK, 2023; pp. 3–34. [Google Scholar]
  131. Forte, A.; Dourado, F.; Mota, A.; Neto, B.; Gama, M.; Ferreira, E.C. Life cycle assessment of bacterial cellulose production. Int. J. Life Cycle Assess. 2021, 26, 864–878. [Google Scholar] [CrossRef]
  132. Katakojwala, R.; Mohan, S.V. Microcrystalline cellulose production from sugarcane bagasse: Sustainable process development and life cycle assessment. J. Clean. Prod. 2020, 249, 119342. [Google Scholar] [CrossRef]
  133. Turk, J.; Oven, P.; Poljanšek, I.; Lešek, A.; Knez, F.; Malovrh Rebec, K. Evaluation of an environmental profile comparison for nanocellulose production and supply chain by applying different life cycle assessment methods. J. Clean. Prod. 2020, 247, 119107. [Google Scholar] [CrossRef]
  134. de Araújo e Silva, R.; Santa Brígida, A.I.; de Freitas Rosa, M.; da Silva Neto, R.M.; Spinosa, W.A.; Benício de Sá Filho, E.; Brito de Figueirêdo, M.C. An approach for implementing ecodesign at early research stage: A case study of bacterial cellulose production. J. Clean. Prod. 2020, 269, 122245. [Google Scholar] [CrossRef]
  135. Martínez, E.; Posada, L.; Botero, J.C.; Rios-Arango, J.A.; Zapata-Benabithe, Z.; López, S.; Molina-Ramírez, C.; Osorio, M.A.; Castro, C.I. Nata de fique: A cost-effective alternative for the large-scale production of bacterial nanocellulose. Ind. Crops Prod. 2023, 192, 116015. [Google Scholar] [CrossRef]
  136. Aragão, J.V.S.; Costa, A.F.S.; Silva, G.L.; Silva, S.M.; Macêdo, J.S.; Galdino, C.J.S.; Milanez, V.F.A.; Sarubbo, L.A. Analysis of the environmental life cycle of bacterial cellulose production. Chem. Eng. Trans. 2020, 79, 445–450. [Google Scholar]
  137. Wada, N.; Fujie, T.; Sasaki, R.; Matsushima, T.; Takahashi, K. Direct synthesis of a robust cellulosic composite from cellulose acetate and a nanofibrillated bacterial cellulose sol. Polym. J. 2022, 54, 735–740. [Google Scholar] [CrossRef]
  138. Feng, Q.; Wu, W.; Cui, Y.; Zhou, Y.; Zhang, Y.; Xu, S.; Lin, L.; Zhou, M.; Li, Z. Reversible wettability control of self-assembled TiO2 scaffolds on bacterial cellulose from superhydrophobicity to superhydrophilicity. Cellulose 2024, 31, 2907–2920. [Google Scholar] [CrossRef]
  139. Acharjee, S.A.; Bharali, P.; Gogoi, B.; Sorhie, V.; Walling, B.; Alemtoshi. PHA-Based Bioplastic: A Potential Alternative to Address Microplastic Pollution. Water Air Soil Pollut. 2022, 234, 21. [Google Scholar] [CrossRef]
  140. Patrício Silva, A.L.; Prata, J.C.; Walker, T.R.; Duarte, A.C.; Ouyang, W.; Barcelò, D.; Rocha-Santos, T. Increased plastic pollution due to COVID-19 pandemic: Challenges and recommendations. Chem. Eng. J. 2021, 405, 126683. [Google Scholar] [CrossRef]
  141. Oleksy, M.; Dynarowicz, K.; Aebisher, D. Advances in Biodegradable Polymers and Biomaterials for Medical Applications—A Review. Molecules 2023, 28, 6213. [Google Scholar] [CrossRef]
  142. Mao, L.; Wang, L.; Zhang, M.; Ullah, M.W.; Liu, L.; Zhao, W.; Li, Y.; Ahmed, A.A.Q.; Cheng, H.; Shi, Z.; et al. In Situ Synthesized Selenium Nanoparticles-Decorated Bacterial Cellulose/Gelatin Hydrogel with Enhanced Antibacterial, Antioxidant, and Anti-Inflammatory Capabilities for Facilitating Skin Wound Healing. Adv. Healthc. Mater. 2021, 10, 2100402. [Google Scholar] [CrossRef] [PubMed]
  143. Gutierrez, E.; Burdiles, P.A.; Quero, F.; Palma, P.; Olate-Moya, F.; Palza, H. 3D Printing of Antimicrobial Alginate/Bacterial-Cellulose Composite Hydrogels by Incorporating Copper Nanostructures. ACS Biomater. Sci. Eng. 2019, 5, 6290–6299. [Google Scholar] [CrossRef] [PubMed]
  144. Supanakorn, G.; Taokaew, S.; Phisalaphong, M. Multifunctional Cellulosic Natural Rubber and Silver Nanoparticle Films with Superior Chemical Resistance and Antibacterial Properties. Nanomaterials 2023, 13, 521. [Google Scholar] [CrossRef] [PubMed]
  145. Dincă, V.; Mocanu, A.; Isopencu, G.; Busuioc, C.; Brajnicov, S.; Vlad, A.; Icriverzi, M.; Roseanu, A.; Dinescu, M.; Stroescu, M.; et al. Biocompatible pure ZnO nanoparticles-3D bacterial cellulose biointerfaces with antibacterial properties. Arab. J. Chem. 2020, 13, 3521–3533. [Google Scholar] [CrossRef]
  146. Windarsih, A.; Indrianingsih, A.W.; Maryana, R.; Apriyana, W.; Rosyida, V.T.; Nurhayati, S.; Jatmiko, T.H.; Ratih, D.; Suwanto, A. Gold modified bacterial cellulose from coconut water waste and its antibacterial activity. Waste Biomass Valorization 2022, 13, 4157–4164. [Google Scholar] [CrossRef]
  147. Skeeters, S.S.; Rosu, A.C.; Divyanshi; Yang, J.; Zhang, K. Comparative Determination of Cytotoxicity of Sub-10 nm Copper Nanoparticles to Prokaryotic and Eukaryotic Systems. ACS Appl. Mater. Interfaces 2020, 12, 50203–50211. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, G.; Jin, W.; Qasim, A.M.; Gao, A.; Peng, X.; Li, W.; Feng, H.; Chu, P.K. Antibacterial effects of titanium embedded with silver nanoparticles based on electron-transfer-induced reactive oxygen species. Biomaterials 2017, 124, 25–34. [Google Scholar] [CrossRef] [PubMed]
  149. Chandra, H.; Bishnoi, P.; Yadav, A.; Patni, B.; Mishra, A.P.; Nautiyal, A.R. Antimicrobial Resistance and the Alternative Resources with Special Emphasis on Plant-Based Antimicrobials—A Review. Plants 2017, 6, 16. [Google Scholar] [CrossRef]
  150. Mocanu, A.; Isopencu, G.; Busuioc, C.; Popa, O.-M.; Dietrich, P.; Socaciu-Siebert, L. Bacterial cellulose films with ZnO nanoparticles and propolis extracts: Synergistic antimicrobial effect. Sci. Rep. 2019, 9, 17687. [Google Scholar] [CrossRef]
  151. Cabañas-Romero, L.V.; Valls, C.; Valenzuela, S.V.; Roncero, M.B.; Pastor, F.I.J.; Diaz, P.; Martínez, J. Bacterial Cellulose–Chitosan Paper with Antimicrobial and Antioxidant Activities. Biomacromolecules 2020, 21, 1568–1577. [Google Scholar] [CrossRef]
  152. Sajjad, W.; He, F.; Ullah, M.W.; Ikram, M.; Shah, S.M.; Khan, R.; Khan, T.; Khalid, A.; Yang, G.; Wahid, F. Fabrication of Bacterial Cellulose-Curcumin Nanocomposite as a Novel Dressing for Partial Thickness Skin Burn. Front. Bioeng. Biotechnol. 2020, 8, 553037. [Google Scholar] [CrossRef] [PubMed]
  153. Gupta, A.; Briffa, S.M.; Swingler, S.; Gibson, H.; Kannappan, V.; Adamus, G.; Kowalczuk, M.; Martin, C.; Radecka, I. Synthesis of Silver Nanoparticles Using Curcumin-Cyclodextrins Loaded into Bacterial Cellulose-Based Hydrogels for Wound Dressing Applications. Biomacromolecules 2020, 21, 1802–1811. [Google Scholar] [CrossRef] [PubMed]
  154. Zmejkoski, D.; Spasojević, D.; Orlovska, I.; Kozyrovska, N.; Soković, M.; Glamočlija, J.; Dmitrović, S.; Matović, B.; Tasić, N.; Maksimović, V.; et al. Bacterial cellulose-lignin composite hydrogel as a promising agent in chronic wound healing. Int. J. Biol. Macromol. 2018, 118, 494–503. [Google Scholar] [CrossRef] [PubMed]
  155. Jantarat, C.; Attakitmongkol, K.; Nichsapa, S.; Sirathanarun, P.; Srivaro, S. Molecularly imprinted bacterial cellulose for sustained-release delivery of quercetin. J. Biomater. Sci. Polym. Ed. 2020, 31, 1961–1976. [Google Scholar] [CrossRef] [PubMed]
  156. Ngwabebhoh, F.A.; Patwa, R.; Zandraa, O.; Saha, N.; Saha, P. Preparation and characterization of injectable self-antibacterial gelatin/carrageenan/bacterial cellulose hydrogel scaffolds for wound healing application. J. Drug Deliv. Sci. Technol. 2021, 63, 102415. [Google Scholar] [CrossRef]
  157. Indrianingsih, A.W.; Rosyida, V.T.; Apriyana, W.; Hayati, S.N.; Darsih, C.; Nisa, K.; Ratih, D. Antioxidant and antibacterial properties of bacterial cellulose—Indonesian plant extract composites for mask sheet. J. Appl. Pharm. Sci. 2020, 10, 037–042. [Google Scholar] [CrossRef]
  158. Fatima, A.; Yasir, S.; Ul-Islam, M.; Kamal, T.; Ahmad, M.W.; Abbas, Y.; Manan, S.; Ullah, M.W.; Yang, G. Ex situ development and characterization of green antibacterial bacterial cellulose-based composites for potential biomedical applications. Adv. Compos. Hybrid Mater. 2022, 5, 307–321. [Google Scholar] [CrossRef]
  159. Ul-Islam, M.; Ahmad, F.; Fatima, A.; Shah, N.; Yasir, S.; Ahmad, M.W.; Manan, S.; Ullah, M.W. Ex situ Synthesis and Characterization of High Strength Multipurpose Bacterial Cellulose-Aloe vera Hydrogels. Front. Bioeng. Biotechnol. 2021, 9, 601988. [Google Scholar] [CrossRef] [PubMed]
  160. Kumar, N.; Daniloski, D.; Pratibha; Neeraj; D’Cunha, N.M.; Naumovski, N.; Petkoska, A.T. Pomegranate peel extract—A natural bioactive addition to novel active edible packaging. Food Res. Int. 2022, 156, 111378. [Google Scholar] [CrossRef]
  161. Glazer, I.; Masaphy, S.; Marciano, P.; Bar-Ilan, I.; Holland, D.; Kerem, Z.; Amir, R. Partial Identification of Antifungal Compounds from Punica granatum Peel Extracts. J. Agric. Food Chem. 2012, 60, 4841–4848. [Google Scholar] [CrossRef]
  162. Habib, H.M.; El-Gendi, H.; El-Fakharany, E.M.; El-Ziney, M.G.; El-Yazbi, A.F.; Al Meqbaali, F.T.; Ibrahim, W.H. Antioxidant, Anti-Inflammatory, Antimicrobial, and Anticancer Activities of Pomegranate Juice Concentrate. Nutrients 2023, 15, 2709. [Google Scholar] [CrossRef] [PubMed]
  163. Ul-Islam, M.; Alhajaim, W.; Fatima, A.; Yasir, S.; Kamal, T.; Abbas, Y.; Khan, S.; Khan, A.H.; Manan, S.; Ullah, M.W.; et al. Development of low-cost bacterial cellulose-pomegranate peel extract-based antibacterial composite for potential biomedical applications. Int. J. Biol. Macromol. 2023, 231, 123269. [Google Scholar] [CrossRef] [PubMed]
  164. Marisca Evalina, G.; Yulanda, A.; Yuana Elly, A. Effectivity of Patch Herbal Mixture Composed of Mangosteen Peel Extract and Bacterial Cellulose for Wound Healing. Pharmacogn. J. 2023, 15, 461–466. [Google Scholar]
  165. Taokaew, S.; Chiaoprakobkij, N.; Siripong, P.; Sanchavanakit, N.; Pavasant, P.; Phisalaphong, M. Multifunctional cellulosic nanofiber film with enhanced antimicrobial and anticancer properties by incorporation of ethanolic extract of Garcinia mangostana peel. Mater. Sci. Eng. C 2021, 120, 111783. [Google Scholar] [CrossRef] [PubMed]
  166. Taokaew, S.; Piyaviriyakul, S.; Siripong, P.; Phisalaphong, M. Aqueous and ethanolic extracts of mangosteen peels as natural antimicrobial/anticancer materials against pathogenic microbes and B16F10 murine melanoma. Chiang Mai J. Sci. 2018, 45, 1345–1358. [Google Scholar]
  167. Fatima, A.; Yasir, S.; Khan, M.S.; Manan, S.; Ullah, M.W.; Ul-Islam, M. Plant extract-loaded bacterial cellulose composite membrane for potential biomedical applications. J. Bioresour. Bioprod. 2021, 6, 26–32. [Google Scholar] [CrossRef]
  168. Kamal, T.; Ul-Islam, M.; Khan, S.B.; Bakhsh, E.M.; Chani, M.T.S. Development of plant extract impregnated bacterial cellulose as a green antimicrobial composite for potential biomedical applications. Ind. Crops Prod. 2022, 187, 115337. [Google Scholar] [CrossRef]
  169. Carrillo, A.J.; González-Aguilar, J.; Romero, M.; Coronado, J.M. Solar Energy on Demand: A Review on High Temperature Thermochemical Heat Storage Systems and Materials. Chem. Rev. 2019, 119, 4777–4816. [Google Scholar] [CrossRef] [PubMed]
  170. Han, X.; Ding, S.; Hu, H.; Wang, S. Recent advances in structural regulation and optimization of high-performance solar-driven interfacial evaporation systems. J. Mater. Chem. A 2022, 10, 18509–18541. [Google Scholar] [CrossRef]
  171. Wang, G.; Tang, Z.; Gao, Y.; Liu, P.; Li, Y.; Li, A.; Chen, X. Phase Change Thermal Storage Materials for Interdisciplinary Applications. Chem. Rev. 2023, 123, 6953–7024. [Google Scholar] [CrossRef]
  172. Zhang, D.; Zhang, S.; Liang, Q.; Song, J.; Guan, M.; Zhang, T.; Chen, S.; Wang, H. One-Step Synthesis of Multifunctional Bacterial Cellulose Film-Based Phase Change Materials with Cross-Linked Network Structure for Solar–Thermal Energy Conversion, Storage, and Utilization. Small 2023, 20, 2307259. [Google Scholar] [CrossRef] [PubMed]
  173. Hossain, M.H.; Chowdhury, M.A.; Hossain, N.; Islam, M.A.; Mobarak, M.H. Advances of lithium-ion batteries anode materials—A review. Chem. Eng. J. Adv. 2023, 16, 100569. [Google Scholar] [CrossRef]
  174. Lingappan, N.; Lee, W.; Passerini, S.; Pecht, M. A comprehensive review of separator membranes in lithium-ion batteries. Renew. Sustain. Energy Rev. 2023, 187, 113726. [Google Scholar] [CrossRef]
  175. Ajkidkarn, P.; Manuspiya, H. Novel bacterial cellulose nanocrystals/polyether block amide microporous membranes as separators for lithium-ion batteries. Int. J. Biol. Macromol. 2020, 164, 3580–3588. [Google Scholar] [CrossRef] [PubMed]
  176. Kartika Sari, A.; Mohamad Yunus, R.; Majlan, E.H.; Loh, K.S.; Wong, W.Y.; Saidin, N.U.; Alva, S.; Khaerudini, D.S. Nata de Cassava Type of Bacterial Cellulose Doped with Phosphoric Acid as a Proton Exchange Membrane. Membranes 2023, 13, 43. [Google Scholar] [CrossRef] [PubMed]
  177. Hosseini, H.; Teymouri, M.; Saboor, S.; Khalili, A.; Goodarzi, V.; Poudineh Hajipoor, F.; Khonakdar, H.A.; Shojaei, S.; Asefnejad, A.; Bagheri, H. Challenge between sequence presences of conductive additives on flexibility, dielectric and supercapacitance behaviors of nanofibrillated template of bacterial cellulose aerogels. Eur. Polym. J. 2019, 115, 335–345. [Google Scholar] [CrossRef]
  178. Ye, J.; Guo, L.; Zheng, S.; Feng, Y.; Zhang, T.; Yang, Z.; Yuan, Q.; Shen, G.; Zhang, Z. Synthesis of bacterial cellulose based SnO2-PPy nanocomposites as potential flexible, highly conductive material. Mater. Lett. 2019, 253, 372–376. [Google Scholar] [CrossRef]
  179. Sheng, N.; Chen, S.; Yao, J.; Guan, F.; Zhang, M.; Wang, B.; Wu, Z.; Ji, P.; Wang, H. Polypyrrole@TEMPO-oxidized bacterial cellulose/reduced graphene oxide macrofibers for flexible all-solid-state supercapacitors. Chem. Eng. J. 2019, 368, 1022–1032. [Google Scholar] [CrossRef]
  180. Vilela, C.; Silva, A.C.Q.; Domingues, E.M.; Gonçalves, G.; Martins, M.A.; Figueiredo, F.M.L.; Santos, S.A.O.; Freire, C.S.R. Conductive polysaccharides-based proton-exchange membranes for fuel cell applications: The case of bacterial cellulose and fucoidan. Carbohydr. Polym. 2020, 230, 115604. [Google Scholar] [CrossRef]
  181. Gadim, T.D.O.; Loureiro, F.J.A.; Vilela, C.; Rosero-Navarro, N.; Silvestre, A.J.D.; Freire, C.S.R.; Figueiredo, F.M.L. Protonic conductivity and fuel cell tests of nanocomposite membranes based on bacterial cellulose. Electrochim. Acta 2017, 233, 52–61. [Google Scholar] [CrossRef]
  182. Naumi, F.; Natanael, C.L.; Rahayu, I.; Indrarti, L.; Hendrana, S. Polymer Electrolyte Membrane Fuel Cell based on Sulfonated Polystyrene and Phosphoric Acid with Biocellulose as a Matrix. Res. J. Chem. Env. 2018, 22, 289–293. [Google Scholar]
  183. Trindade, E.C.A.; Antônio, R.V.; Brandes, R.; de Souza, L.; Neto, G.; Vargas, V.M.M.; Carminatti, C.A.; de Oliveira Souza Recouvreux, D. Carbon fiber-embedded bacterial cellulose/polyaniline nanocomposite with tailored for microbial fuel cells electrode. J. Appl. Polym. Sci. 2020, 137, 49036. [Google Scholar] [CrossRef]
  184. Luo, H.; Dong, J.; Xu, X.; Wang, J.; Yang, Z.; Wan, Y. Exploring excellent dispersion of graphene nanosheets in three-dimensional bacterial cellulose for ultra-strong nanocomposite hydrogels. Compos. Part A Appl. Sci. Manuf. 2018, 109, 290–297. [Google Scholar] [CrossRef]
  185. Luo, H.; Dong, J.; Yao, F.; Yang, Z.; Li, W.; Wang, J.; Xu, X.; Hu, J.; Wan, Y. Layer-by-Layer Assembled Bacterial Cellulose/Graphene Oxide Hydrogels with Extremely Enhanced Mechanical Properties. Nano-Micro Lett. 2018, 10, 42. [Google Scholar] [CrossRef]
  186. Luo, H.; Dong, J.; Zhang, Y.; Li, G.; Guo, R.; Zuo, G.; Ye, M.; Wang, Z.; Yang, Z.; Wan, Y. Constructing 3D bacterial cellulose/graphene/polyaniline nanocomposites by novel layer-by-layer in situ culture toward mechanically robust and highly flexible freestanding electrodes for supercapacitors. Chem. Eng. J. 2018, 334, 1148–1158. [Google Scholar] [CrossRef]
  187. Wan, Y.; Li, J.; Yang, Z.; Ao, H.; Xiong, L.; Luo, H. Simultaneously depositing polyaniline onto bacterial cellulose nanofibers and graphene nanosheets toward electrically conductive nanocomposites. Curr. Appl. Phys. 2018, 18, 933–940. [Google Scholar] [CrossRef]
  188. Tan, H.; Tang, J.; Henzie, J.; Li, Y.; Xu, X.; Chen, T.; Wang, Z.; Wang, J.; Ide, Y.; Bando, Y.; et al. Assembly of Hollow Carbon Nanospheres on Graphene Nanosheets and Creation of Iron–Nitrogen-Doped Porous Carbon for Oxygen Reduction. ACS Nano 2018, 12, 5674–5683. [Google Scholar] [CrossRef]
  189. Luo, H.; Xiong, P.; Xie, J.; Yang, Z.; Huang, Y.; Hu, J.; Wan, Y.; Xu, Y. Uniformly Dispersed Freestanding Carbon Nanofiber/Graphene Electrodes Made by a Scalable Biological Method for High-Performance Flexible Supercapacitors. Adv. Funct. Mater. 2018, 28, 1803075. [Google Scholar] [CrossRef]
  190. Dhar, P.; Pratto, B.; Gonçalves Cruz, A.J.; Bankar, S. Valorization of sugarcane straw to produce highly conductive bacterial cellulose / graphene nanocomposite films through in situ fermentation: Kinetic analysis and property evaluation. J. Clean. Prod. 2019, 238, 117859. [Google Scholar] [CrossRef]
  191. Wei, Q.; Fei, N.; Islam, A.; Lei, T.; Hong, L.; Peng, R.; Fan, X.; Chen, L.; Gao, P.; Ge, Z. Small-Molecule Emitters with High Quantum Efficiency: Mechanisms, Structures, and Applications in OLED Devices. Adv. Opt. Mater. 2018, 6, 1800512. [Google Scholar] [CrossRef]
  192. Lim, Y.-W.; Jin, J.; Bae, B.-S. Optically Transparent Multiscale Composite Films for Flexible and Wearable Electronics. Adv. Mater. 2020, 32, 1907143. [Google Scholar] [CrossRef]
  193. Zhao, D.; Zhu, Y.; Cheng, W.; Chen, W.; Wu, Y.; Yu, H. Cellulose-Based Flexible Functional Materials for Emerging Intelligent Electronics. Adv. Mater. 2021, 33, 2000619. [Google Scholar] [CrossRef]
  194. Legnani, C.; Barud, H.S.; Caiut, J.M.A.; Calil, V.L.; Maciel, I.O.; Quirino, W.G.; Ribeiro, S.J.L.; Cremona, M. Transparent bacterial cellulose nanocomposites used as substrate for organic light-emitting diodes. J. Mater. Sci. Mater. Electron. 2019, 30, 16718–16723. [Google Scholar] [CrossRef]
  195. Cebrian, A.V.S.; Carvalho, R.S.; Barreto, A.R.J.; Maturi, F.E.; Barud, H.S.; Silva, R.R.; Legnani, C.; Cremona, M.; Ribeiro, S.J.L. Development of Conformable Substrates for OLEDs Using Highly Transparent Bacterial Cellulose Modified with Recycled Polystyrene. Adv. Sustain. Syst. 2022, 6, 2000258. [Google Scholar] [CrossRef]
  196. Osman, A.I.; Chen, Z.; Elgarahy, A.M.; Farghali, M.; Mohamed, I.M.A.; Priya, A.K.; Hawash, H.B.; Yap, P.-S. Membrane Technology for Energy Saving: Principles, Techniques, Applications, Challenges, and Prospects. Adv. Energy Sustain. Res. 2024, 5, 2400011. [Google Scholar] [CrossRef]
  197. Khanzada, N.K.; Al-Juboori, R.A.; Khatri, M.; Ahmed, F.E.; Ibrahim, Y.; Hilal, N. Sustainability in Membrane Technology: Membrane Recycling and Fabrication Using Recycled Waste. Membranes 2024, 14, 52. [Google Scholar] [CrossRef]
  198. Hou, Y.; Duan, C.; Zhu, G.; Luo, H.; Liang, S.; Jin, Y.; Zhao, N.; Xu, J. Functional bacterial cellulose membranes with 3D porous architectures: Conventional drying, tunable wettability and water/oil separation. J. Membr. Sci. 2019, 591, 117312. [Google Scholar] [CrossRef]
  199. Wahid, F.; Zhao, X.-J.; Duan, Y.-X.; Zhao, X.-Q.; Jia, S.-R.; Zhong, C. Designing of bacterial cellulose-based superhydrophilic/underwater superoleophobic membrane for oil/water separation. Carbohydr. Polym. 2021, 257, 117611. [Google Scholar] [CrossRef]
  200. Than-ardna, B.; Weder, C.; Manuspiya, H. Superhydrophilic bacterial cellulose membranes efficiently separate oil-in-water emulsions. J. Mater. Sci. 2023, 58, 5086–5103. [Google Scholar] [CrossRef]
  201. Sijabat, E.K.; Nuruddin, A.; Aditiawati, P.; Sunendar Purwasasmita, B. Flat sheet membrane composite for desalinisation applications based on Bacterial Nanocellulose (BNC) from banana peel waste, cellulose, and silica. Mater. Res. Express 2020, 7, 105004. [Google Scholar] [CrossRef]
  202. Liu, J.; Chen, T.-W.; Yang, Y.-L.; Bai, Z.-C.; Xia, L.-R.; Wang, M.; Lv, X.-L.; Li, L. Removal of heavy metal ions and anionic dyes from aqueous solutions using amide-functionalized cellulose-based adsorbents. Carbohydr. Polym. 2020, 230, 115619. [Google Scholar] [CrossRef] [PubMed]
  203. Tissera, N.D.; Wijesena, R.N.; Yasasri, H.; de Silva, K.M.N.; de Silva, R.M. Fibrous keratin protein bio micro structure for efficient removal of hazardous dye waste from water: Surface charge mediated interfaces for multiple adsorption desorption cycles. Mater. Chem. Phys. 2020, 246, 122790. [Google Scholar] [CrossRef]
  204. Durval, I.J.B.; Silvestre, G.U.; Medeiros, A.D.; da Silva, C.G.; Amorim, J.D.; Rufino, R.D.; Costa, A.F.S.; Sarubbo, L.A. Biosurfactant and Bacterial Cellulose Applied to Textile Effluent Treatment. Chem. Eng. Trans. 2023, 100, 415–420. [Google Scholar]
  205. Kim, H.; Kim, H.R. Production of coffee-dyed bacterial cellulose as a bio-leather and using it as a dye adsorbent. PLoS ONE 2022, 17, e0265743. [Google Scholar] [CrossRef] [PubMed]
  206. Syukri, D.; Suryanto, H.; Kurniawan, F.; Hari, P.D.; Fiana, R.M.; Rini. Bacterial reduction in river water using nanocellulose membrane from pineapple biomass with ferrous-ferric oxide reinforcement. Glob. J. Environ. Sci. Manag. 2024, 10, 643–656. [Google Scholar]
  207. Suryanto, H.; Kurniawan, F.; Syukri, D.; Binoj, J.S.; Hari, P.D.; Yanuhar, U. Properties of bacterial cellulose acetate nanocomposite with TiO2 nanoparticle and graphene reinforcement. Int. J. Biol. Macromol. 2023, 235, 123705. [Google Scholar] [CrossRef] [PubMed]
  208. Cazón, P.; Vázquez, M. Bacterial cellulose as a biodegradable food packaging material: A review. Food Hydrocoll. 2021, 113, 106530. [Google Scholar] [CrossRef]
  209. Ludwicka, K.; Kaczmarek, M.; Białkowska, A. Bacterial Nanocellulose—A Biobased Polymer for Active and Intelligent Food Packaging Applications: Recent Advances and Developments. Polymers 2020, 12, 2209. [Google Scholar] [CrossRef]
  210. Yu, Q.; Yang, L.; Wang, S.; Zhang, L.; Sun, D. Citric acid cross-linked regenerated bacterial cellulose as biodegradable film for food packaging. Cellulose 2023, 30, 10273–10284. [Google Scholar] [CrossRef]
  211. Jiang, Z.; Cheung, K.M.; Ngai, T. Development of strong and high-barrier food packaging films from cyclic-anhydride modified bacterial cellulose. RSC Sustain. 2024, 2, 139–152. [Google Scholar] [CrossRef]
  212. Yanti, N.A.; Ahmad, S.W.; Ramadhan, L.O.; Jamili; Muzuni; Walhidayah, T.; Mamangkey, J. Properties and Application of Edible Modified Bacterial Cellulose Film Based Sago Liquid Waste as Food Packaging. Polymers 2021, 13, 3570. [Google Scholar] [CrossRef]
  213. Rouhi, M.; Garavand, F.; Heydari, M.; Mohammadi, R.; Sarlak, Z.; Cacciotti, I.; Razavi, S.H.; Mousavi, M.; Parandi, E. Fabrication of novel antimicrobial nanocomposite films based on polyvinyl alcohol, bacterial cellulose nanocrystals, and boric acid for food packaging. J. Food Meas. Charact. 2024, 18, 2146–2161. [Google Scholar] [CrossRef]
  214. Atta, O.M.; Manan, S.; Ul-Islam, M.U.; Ahmed, A.A.Q.; Ullah, M.W.; Yang, G. Silver Decorated Bacterial Cellulose Nanocomposites as Antimicrobial Food Packaging Materials. ES Food Agrofor. 2021, 6, 12–26. [Google Scholar] [CrossRef]
  215. Xie, Y.; Niu, X.; Yang, J.; Fan, R.; Shi, J.; Ullah, N.; Feng, X.; Chen, L. Active biodegradable films based on the whole potato peel incorporated with bacterial cellulose and curcumin. Int. J. Biol. Macromol. 2020, 150, 480–491. [Google Scholar] [CrossRef]
  216. Papadaki, A.; Manikas, A.C.; Papazoglou, E.; Kachrimanidou, V.; Lappa, I.; Galiotis, C.; Mandala, I.; Kopsahelis, N. Whey protein films reinforced with bacterial cellulose nanowhiskers: Improving edible film properties via a circular economy approach. Food Chem. 2022, 385, 132604. [Google Scholar] [CrossRef]
  217. Muhammed, A.P.; Thangarasu, S.; Manoharan, R.K.; Oh, T.H. Ex-situ fabrication of engineered green network of multifaceted bacterial cellulose film with enhanced antimicrobial properties for post-harvest preservation of table grapes. Food Packag. Shelf Life 2024, 43, 101284. [Google Scholar] [CrossRef]
  218. Kusuma, A.C.; Chou, Y.-C.; Hsieh, C.-C.; Santoso, S.P.; Go, A.W.; Lin, H.-T.V.; Hsiao, I.L.; Lin, S.-P. Agar-altered foaming bacterial cellulose with carvacrol for active food packaging applications. Food Packag. Shelf Life 2024, 42, 101269. [Google Scholar] [CrossRef]
  219. Paximada, P.; Echegoyen, Y.; Koutinas, A.A.; Mandala, I.G.; Lagaron, J.M. Encapsulation of hydrophilic and lipophilized catechin into nanoparticles through emulsion electrospraying. Food Hydrocoll. 2017, 64, 123–132. [Google Scholar] [CrossRef]
  220. Razavi, M.S.; Golmohammadi, A.; Nematollahzadeh, A.; Fiori, F.; Rovera, C.; Farris, S. Preparation of cinnamon essential oil emulsion by bacterial cellulose nanocrystals and fish gelatin. Food Hydrocoll. 2020, 109, 106111. [Google Scholar] [CrossRef]
  221. Liao, W.; Liu, X.; Zhao, Q.; Lu, Z.; Feng, A.; Sun, X. Physicochemical, antibacterial and food preservation properties of active packaging films based on chitosan/ε-polylysine-grafted bacterial cellulose. Int. J. Biol. Macromol. 2023, 253, 127231. [Google Scholar] [CrossRef]
  222. Abral, H.; Pratama, A.B.; Handayani, D.; Mahardika, M.; Aminah, I.; Sandrawati, N.; Sugiarti, E.; Muslimin, A.N.; Sapuan, S.M.; Ilyas, R.A. Antimicrobial Edible Film Prepared from Bacterial Cellulose Nanofibers/Starch/Chitosan for a Food Packaging Alternative. Int. J. Polym. Sci. 2021, 2021, 6641284. [Google Scholar] [CrossRef]
  223. Salari, M.; Sowti Khiabani, M.; Rezaei Mokarram, R.; Ghanbarzadeh, B.; Samadi Kafil, H. Development and evaluation of chitosan based active nanocomposite films containing bacterial cellulose nanocrystals and silver nanoparticles. Food Hydrocoll. 2018, 84, 414–423. [Google Scholar] [CrossRef]
  224. Cazón, P.; Puertas, G.; Vázquez, M. Characterization of multilayer bacterial cellulose-chitosan films loaded with grape bagasse antioxidant extract: Insights into spectral and water properties, microstructure, and antioxidant activity. Int. J. Biol. Macromol. 2024, 268, 131774. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM images of BC (A) and plant cellulose (B). The inset images show the high-resolution photographs of the samples (Reprinted with permission from ref. [20] Copyright © 2019 Elsevier).
Figure 1. SEM images of BC (A) and plant cellulose (B). The inset images show the high-resolution photographs of the samples (Reprinted with permission from ref. [20] Copyright © 2019 Elsevier).
Fermentation 10 00316 g001
Figure 2. Trend of publications from 2013 to 2023, showing the cumulative number of articles related to BC (A). Additionally, it provides a breakdown of the percentage of publications by subject area (B), obtained through keyword searches for “Bacterial nanocellulose” on the Scopus platform.
Figure 2. Trend of publications from 2013 to 2023, showing the cumulative number of articles related to BC (A). Additionally, it provides a breakdown of the percentage of publications by subject area (B), obtained through keyword searches for “Bacterial nanocellulose” on the Scopus platform.
Fermentation 10 00316 g002
Figure 4. BC production strategies by genetic engineering of K. xylinus strains under different conditions. Adaptive laboratory evolution stress conditions in strain K2G30 (UMCC 2756) (A): schematic representation of the metabolic pathway for fructose and mannose (a), BC production using fructose and a comparison with mannitol and glucose (b), and schematic depiction of the arrangement of the pgi-related gene in the genome of K. xylinus K2G30 (c) (Reprinted from ref. [124] open access Copyright © 2022 Frontiers). Genomic and metabolic engineering of strain DSM 2325 (B): central carbon metabolism and BC biosynthetic pathway of K. xylinus (d), Systems Biology Markup Language file depiction of reversible reactions in a genome-scale metabolic model KxyMBEL1810 (e), random sampling and correlation analysis results for glycolysis and pentose phosphate pathway reactions, identifying overexpression targets PGI and GND (f). BC concentration and yield comparisons among wild-type and genetically modified strains (Reprinted with permission from ref. [127] Copyright © 2019 Wiley). Recombination of strain ATCC 23770 to use mannose-rich source (C) (Reprinted with permission from ref. [128] Copyright © 2023 Elsevier).
Figure 4. BC production strategies by genetic engineering of K. xylinus strains under different conditions. Adaptive laboratory evolution stress conditions in strain K2G30 (UMCC 2756) (A): schematic representation of the metabolic pathway for fructose and mannose (a), BC production using fructose and a comparison with mannitol and glucose (b), and schematic depiction of the arrangement of the pgi-related gene in the genome of K. xylinus K2G30 (c) (Reprinted from ref. [124] open access Copyright © 2022 Frontiers). Genomic and metabolic engineering of strain DSM 2325 (B): central carbon metabolism and BC biosynthetic pathway of K. xylinus (d), Systems Biology Markup Language file depiction of reversible reactions in a genome-scale metabolic model KxyMBEL1810 (e), random sampling and correlation analysis results for glycolysis and pentose phosphate pathway reactions, identifying overexpression targets PGI and GND (f). BC concentration and yield comparisons among wild-type and genetically modified strains (Reprinted with permission from ref. [127] Copyright © 2019 Wiley). Recombination of strain ATCC 23770 to use mannose-rich source (C) (Reprinted with permission from ref. [128] Copyright © 2023 Elsevier).
Fermentation 10 00316 g004
Figure 6. Filtration membrane application of BC loaded with SiO2 nanoparticles (A) (Reprinted with permission from ref. [198] Copyright © 2019 Elsevier) and SiO2 microparticles (B) (Reprinted with permission from ref. [199] Copyright © 2021 Elsevier) for emulsion separation.
Figure 6. Filtration membrane application of BC loaded with SiO2 nanoparticles (A) (Reprinted with permission from ref. [198] Copyright © 2019 Elsevier) and SiO2 microparticles (B) (Reprinted with permission from ref. [199] Copyright © 2021 Elsevier) for emulsion separation.
Fermentation 10 00316 g006
Figure 7. Food packaging application of modified BC films. Cyclic-anhydride-modified BC (A): the surfaces and inner structure of strawberries left uncovered (a) or covered with the pristine BC film (b), BC films modified with maleic anhydride (c), BC films modified with 2-Octenylsuccinic anhydride (d), and low-density polyethylene film (e) for storage durations of 0 and 7 days at 4 °C, 40% relative humidity, followed by storage at room temperature (24 °C, 55% relative humidity) for 2 days (Reprinted from ref. [211] open access Copyright © 2024 by Royal Society of Chemistry). BC/carboxymethyl cellulose/glycerol modified film (B): meat sausage packaged by low-density polyethylene film, native BC film, and modified BC film for storage durations of 1 to 7 days at room temperature (Reprinted from ref. [212] open access Copyright © 2021 by the authors. Licensee MDPI).
Figure 7. Food packaging application of modified BC films. Cyclic-anhydride-modified BC (A): the surfaces and inner structure of strawberries left uncovered (a) or covered with the pristine BC film (b), BC films modified with maleic anhydride (c), BC films modified with 2-Octenylsuccinic anhydride (d), and low-density polyethylene film (e) for storage durations of 0 and 7 days at 4 °C, 40% relative humidity, followed by storage at room temperature (24 °C, 55% relative humidity) for 2 days (Reprinted from ref. [211] open access Copyright © 2024 by Royal Society of Chemistry). BC/carboxymethyl cellulose/glycerol modified film (B): meat sausage packaged by low-density polyethylene film, native BC film, and modified BC film for storage durations of 1 to 7 days at room temperature (Reprinted from ref. [212] open access Copyright © 2021 by the authors. Licensee MDPI).
Fermentation 10 00316 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Taokaew, S. Bacterial Nanocellulose Produced by Cost-Effective and Sustainable Methods and Its Applications: A Review. Fermentation 2024, 10, 316. https://doi.org/10.3390/fermentation10060316

AMA Style

Taokaew S. Bacterial Nanocellulose Produced by Cost-Effective and Sustainable Methods and Its Applications: A Review. Fermentation. 2024; 10(6):316. https://doi.org/10.3390/fermentation10060316

Chicago/Turabian Style

Taokaew, Siriporn. 2024. "Bacterial Nanocellulose Produced by Cost-Effective and Sustainable Methods and Its Applications: A Review" Fermentation 10, no. 6: 316. https://doi.org/10.3390/fermentation10060316

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