Next Article in Journal
Performance Evaluation of Jute/Glass-Fiber-Reinforced Polybutylene Succinate (PBS) Hybrid Composites with Different Layering Configurations
Previous Article in Journal
Magnetic Adsorbents for Wastewater Treatment: Advancements in Their Synthesis Methods
Previous Article in Special Issue
An Overview Regarding Microbial Aspects of Production and Applications of Bacterial Cellulose
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Bacterial Cellulose—A Remarkable Polymer as a Source for Biomaterials Tailoring

by
Lăcrămioara Popa
,
Mihaela Violeta Ghica
*,
Elena-Emilia Tudoroiu
,
Diana-Georgiana Ionescu
and
Cristina-Elena Dinu-Pîrvu
Department of Physical and Colloidal Chemistry, Faculty of Pharmacy, Carol Davila University of Medicine and Pharmacy Bucharest, 6 Traian Vuia Str., 020956 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Materials 2022, 15(3), 1054; https://doi.org/10.3390/ma15031054
Submission received: 10 December 2021 / Revised: 19 January 2022 / Accepted: 27 January 2022 / Published: 29 January 2022
(This article belongs to the Special Issue Advances in Bacterial Cellulose Composites)

Abstract

:
Nowadays, the development of new eco-friendly and biocompatible materials using ‘green’ technologies represents a significant challenge for the biomedical and pharmaceutical fields to reduce the destructive actions of scientific research on the human body and the environment. Thus, bacterial cellulose (BC) has a central place among these novel tailored biomaterials. BC is a non-pathogenic bacteria-produced polysaccharide with a 3D nanofibrous structure, chemically identical to plant cellulose, but exhibiting greater purity and crystallinity. Bacterial cellulose possesses excellent physicochemical and mechanical properties, adequate capacity to absorb a large quantity of water, non-toxicity, chemical inertness, biocompatibility, biodegradability, proper capacity to form films and to stabilize emulsions, high porosity, and a large surface area. Due to its suitable characteristics, this ecological material can combine with multiple polymers and diverse bioactive agents to develop new materials and composites. Bacterial cellulose alone, and with its mixtures, exhibits numerous applications, including in the food and electronic industries and in the biotechnological and biomedical areas (such as in wound dressing, tissue engineering, dental implants, drug delivery systems, and cell culture). This review presents an overview of the main properties and uses of bacterial cellulose and the latest promising future applications, such as in biological diagnosis, biosensors, personalized regenerative medicine, and nerve and ocular tissue engineering.

1. Introduction

Over the last decades, due to the advancement of technology (artificial intelligence or robotics) [1], wide novel, multifunctional, and biomimetic biomaterials (natural, modified natural, or synthetic) have been developed [2] with enhanced properties and applications [3] suitable for use in areas from the food industry to regenerative medicine and bioprinting [4]. These biomaterials can be successfully substitute for the traditional materials [5]. The term ‘biomaterial’ refers to an eco-friendly material, which is based on sustainable resources (agricultural raw materials, fossil, and electronic reserves) [6]. The extensive development of novel chemical and physical methods furnishes new opportunities for the scientific community to study and research particular elements to design effective and safer materials to ensure the regeneration of impaired skin [7]. The researchers have a tremendous interest in the entire groups of biomolecules (monomers, oligomers, and macromolecules), such as carbohydrates, amino acids, proteins, nucleotides, nucleic acids, and lipids [8,9]. The common classification of biomaterials consists of four distinct classes: polymers, composites, ceramics, and metals [7]. For the biomedical and pharmaceutical fields (tissue engineering, wound dressings, bioimaging, drug delivery systems, implants, biosensors, biomedical diagnoses, and treatment of various conditions) [10,11,12], the new projected materials should display similar biological and structural characteristics, such as the indigenous extracellular matrix [13]. The novel biomaterials should have the capacity to sustain their structural stability to assure cellular proliferation and the development of new skin tissues [14]. Fundamentally, these biomaterials show biocompatibility, biodegradability, non-immunogenicity, non-cytotoxicity, biological inertness, histoconductivity, histoinductivity [15], bioactivity, optimum physicochemical, and mechanical properties in order to be used in the biomedical area. Their main purpose is to restore or to replace the skin tissue functions to increase the patient’s quality of life [16]. An essential feature of these newly designed biomaterials consists in their biosynthesis using harmless and safe technologies for the environment, known as ‘green’ methods, that considerably reduce the negative consequences of pollution on the climate and the human body [17].
Among these newly tailored biomaterials, a central place is occupied by bacterial cellulose (BC), an ecological polysaccharide broadly studied for multiple applications due to its excellent physicochemical and biological properties [18]. Bacterial cellulose is derived from cellulose, a natural polymer ubiquitously found in our surroundings, known as ‘the most plentiful biopolymer’ on Earth [19]. It is synthesized by all classes of plants, ranging from fungi, algae, and bacteria to cotton, wood, and hemp. Due to its abundant, renewable, degradable, and recyclable character, cellulose has gained attention as a sustainable material [20]. Cellulose is a hydrophilic polysaccharide consisting of linear macromolecular chains of 1–4 linked β-D-glucopyranosyl units forming linear chains in the cell wall [21]. Its high strength, stiffness, crystallinity, and durability are due to saccharide chains held together by Van der Waals forces and hydrogen bonds. Its overall reactivity is results from the presence of hydroxyl groups and their distribution [22,23,24]. Since the structural and chemical particularities of cellulose are covered in many papers, it will not be detailed in the current article.
Among the characteristics of bacterial cellulose are high biocompatibility, biodegradability, non-toxicity, viscoelasticity, flexibility, chemical stability [25], adequate hydrogel traits [26], unidirectional polarity, and fluctuating density [27]. In comparison with the cellulose produced by plants, bacterial cellulose has a higher crystallinity [28], purity, tensile strength [29], value of degree of polymerization, and Young’s modulus [30]. It has a hydrophilic porous structure that allows it to retain a large quantity of water (>90%). BC has a large applicability in fields ranging from electronics, paper, and food [31,32] to applications in the biomedical industry (tissue engineering, bone and cartilage reconstruction, implants, wound dressings, cornea restoration, artificial blood vessels, orthodontic treatment, drug delivery devices, antibacterial products, biosensors, biological diagnoses, regenerative medicine), the pharmaceutical field, veterinary medicine, the leather industry, and pollution control [33,34,35].
In this review, we will further present the main aspects concerning bacterial cellulose biosynthesis and its properties and applications with emphasis on its biomedical uses, such as in dressing materials and artificial scaffolds [36]. We will discuss various combinations of BC and different biopolymers (natural and synthetic) with several bioactive agents (metals, inorganic substances, plants extract, or drugs) to develop new materials and composites with a large applicability in the biomedical and biotechnological domains [37].

2. Bacterial Cellulose—Pioneer for Continuously Developing Macromolecules

2.1. State of the Art

Bacterial cellulose (BC), also known as bacterial nanocellulose (BNC) due to its nanostructured network [38], represents a particular biopolymer [39] produced by certain bacterial strains through fermentation processes; it is lignin and hemicellulose free [40]. It is also called microbial cellulose [41]. First reported by Brown in 1988, this natural polymer resembles plant and wood-derived cellulose, exhibiting the highest purity compared to the latter. Structurally, bacterial cellulose mainly consists of nanofibrillar polysaccharides, which have a diameter between 20 and 100 nm. Thus, bacterial nanocellulose is much thinner than the cellulose extracted from plants [38]. Bacteria produce extracellular cellulose fibers, in static or even dynamic conditions, with different yields, depending on the species and also on the culture media substrate. Komagataeibacter xylinus (also known as Gluconacetobacter xylinus or Acetobacter xylinum) [39], a strictly aerobic Gram-negative bacterium, is known as the strongest cellulose producer, along with other species: Komagataeibacter medellinensis, Komagataeibacter hanseii, Komagataeibacter oboediens, Komagataeibacter rhaeticus, and Komagataeibacter pomaceti, classified as safe bacteria (GRAS). Other bacteria known to produce BC, Azotobacter, Escherichia, Pseudomonas, Rhizobium, Salmonella, Agrobacterium, Klebsiella, and Sarcina ventriculi, are reported, along with recently discovered Lactobacillus hilgardii [42,43,44].

2.2. Biosynthesis

Bacterial cellulose is synthesized by oxidative fermentation in a synthetic and non-synthetic medium. The earlier-presented non-photosynthetic microorganism Komagataeibacter xylinus is fermented at pH = 3–7 at a temperature of 25–30 °C using a saccharide as a carbon source and producing a large quantity of cellulose microfibrils [45,46,47].
Bacterial cellulose originates in the bacterial cytoplasm and is carried out in the membrane of the microorganisms. With glucose as the substrate, the cytoplasm becomes the host of a reaction cycle: phosphorylation–glucokinase, isomerization–phosphoglucomutase, and uridine diphosphate UDP–glucose (UDPG) are produced [48]. The bacterial cellulose cytoplasm synthesis stage at a cellular (microscopic) level is illustrated in Figure 1.
The next stage occurs in the membrane where cellulose synthase operates. This enzyme is the main determinant of the cellulose type, along with the starting substrate. Thus, the nucleotide-activated glucose chains are extruded through pores, resulting in d-glucose units interconnected through β-1,4-glycosidic bonds. More bonds are formed between these after leaving the ‘mother-organism,’ resulting in longer chains which are inter- and intra-linked by hydrogen bonds measuring approximately 25 nm in width and up to 9 µm in length. Further, the linkage between these structures determines the formation of ribbon-shaped fibrillar formations (<100 nm in width). During culture, bacterial cellulose morphology transforms from the ‘floccus’ to the ‘pellicle’ phase, translated in a 3D-network to resemble a pellicle at the surface of the culture media [49,50,51,52].
After fermentation occurs in the culture media and bacterial cellulose is obtained, a purification process is mandatory since the product is not of high purity and may contain culture media residues (for example, lignin and hemicellulose), unwanted cells, or side products [53,54]. A standard, but expensive, purification procedure requires the following steps: harvesting the bacterial cellulose pellicles from the culture vessel, washing them in distilled water to remove any residual medium, treating them with NaOH/KOH/Na2CO3 at 100 °C for 15–20 min to kill the microorganisms, filtering them using an aspirator, and finally neutralizing the filtrate using distilled water. A drying method is further applied before obtaining the final product [55,56]. Another method of purification also uses an alkaline medium, NaOH or K2CO3 at 80 °C for 60 min; the procedure is performed twice to eliminate all the bacteria [57]. The purification of bacterial cellulose is of crucial importance when its cultivation targets are the fabrication of wound-healing materials, especially cartilage implants [58]. Researchers point out that supercritical CO2 processing and treatment with carbonic acid under high pressure are sufficient purification procedures for microbial cellulose that will be used in biomedical applications. Recent experiments indicate that treatment in biphasic systems is more effective in terms of maintaining the main structure of the cellulose network [59].
One of the aforementioned factors impacting bacterial cellulose synthesis is the culture media. In other words, the carbon and nitrogen sources are the most important. The culture substrates involved is the main reason why obtaining bacterial cellulose is an expensive process. Following eco-friendly fabrication protocols, wastewater from distilleries was reported as an efficient culture media for bacterial cellulose. The method is advantageous because it produces quality bacterial cellulose and diminishes waste disposal by transforming the wastewater into a cheap fermentation promoter [60]. Other parameters such as pH, culture type (static/dynamic, surfaced/submerged), shear forces, and oxygenating rates of the support also influence the synthesis process.
In terms of bacterial cellulose networks, the literature data include different patterns, depending on the type of culture conditions. For example, solid-state cultures of Gluconacetobacter xylinus have been reported to produce bacterial cellulose with a honeycomb geometry, following a polyurethane support pattern. Also, the use of support enhances biomass recovery [61]. Other honeycomb-like matrices were obtained from bacterial cellulose and gelatin, exhibiting large surface areas and uniform pore arrangement. Ampicillin was added, resulting in a retarded release sponge with potential antibacterial applications [62].

2.3. Properties

As expected, the properties of microbial cellulose reside in its structure. It consists of bundles of cellulose nano-fibrils (2–4 nm), which become ribbon-like structures of around 100 µm in diameter and 100 µm in length [63,64,65,66]. Due to this particular nanostructure, bacterial cellulose has the capacity to retain its dry weight inside water; this fact gives this biopolymer superior elasticity, flexibility, and resistance in humid conditions; therefore, bacterial nanocellulose represents an excellent natural resource for designing new bandages for the accelerated healing of lesions [38]. It exhibits an ultrafine web structure which is very difficult to disperse in water [67]. The crystallographic structure is similar to that of cellulose type Iα [43,68]. Compared to other biopolymers, bacterial cellulose exhibits a combination of unique properties due to its fibrillar structure and its light weight, its purity, and its macromolecular properties, including its polymerization degree of up to 8000 [69]; its other characteristics such as its hydrophilicity, crystallinity (up to 90%) [70], moldability, non-toxicity, and biodegradability also set it apart [71]. The pure cellulose nanofibers confer intrinsic high purity and strength without the need of further refining treatments [42,63,71,72]. Chemical modifications are also permitted due to vacant hydroxyl groups that are not engaged in hydrogen bonds [49]. Its unique optical, electrical, and mechanical properties with numerous improvement possibilities have attracted the attention of scientists in many fields [73]. Bacterial cellulose morphology can be changed as needed. The literature data report these biogenesis interventions: growing cellulose in a chitosan-modified culture media or a methylcellulose-carboxymethylcellulose-poly(vinyl alcohol)-modified media for its application in wound dressings or other medical applications [74,75]. The main advantages of bacterial cellulose are summarized in Figure 2.

2.4. Applications

Individually, bacterial cellulose has been an excellent starter for many applications due to its particular physicochemical and biological properties. The direction of its applicability points primarily towards the biomedical field [76]. These are a few examples of how bacterial cellulose can be included in innovative biomedical applications.
First of all, microbial cellulose is a suitable material for 3D-bioprinting. The literature data describe bacterial cellulose as a medical material used in bioprinting costal, auricle, and nasal cartilage, due to its special 3D-structured network and unique properties [77]. By increasing its cellulose content by 17%, modified bacterial cellulose showed similarities to human auricular cartilage one week after its implantation. High compatibility was also demonstrated, along with non-absorbable properties, as the implant was accepted by the surrounding soft tissues [78].
Other challenging therapies where BC may play a vital role involve oral implants and newly guided bone regeneration techniques. BC has proven to exhibit suitable characteristics in the role of a barrier membrane, a component of capital importance in implantology. BC enhances tissue and bone regeneration by separating them from other surrounding tissues [79]. Moreover, BC is an important candidate to replace collagen (cytotoxic) as a shielding membrane, since its rate of biodegradation can be reduced [80]. It also exhibits a promising potential as a root canal treatment material, with perspectives in replacing commercially available disposable paper points [81]. Experimental studies showed that potato starch added to a BC culture media produces an increased viscosity and fills many networks’ vacant spaces. Moreover, scaffolds are obtained after culturing muscle cells onto the surface. This pattern complies with the specifications of hollow organ reconstruction material [82].
BC also represents a feasible material for tissue engineering [83], a field that is currently focused on discovering materials and techniques to artificially mimic a suitable environment for stem cell culture [84]. It is well known that stem cells stand out because of their self-renewal characteristics emphasized by their ability to further differentiate into numerous cell types, depending on the specificity of the organism in question, with a high potential for multiple applications [85,86]. A research team demonstrated how bacterial cellulose can be exploited in this direction: a nanofibrous bacterial cellulose membrane is reported to have the ability to inhibit the differentiation of mouse embryonic stem cells. At the same time, the mouse embryonic fibroblast cultivation was improved, in comparison to the conventional culture media. The pluripotency of the cells was confirmed, along with their ability for quick manipulation, significantly enabling other handling maneuvers [87]. BC also has properties which make it adequate to serve as a scaffold for tissue engineering at the level of the cartilage. Chondrocyte proliferation studies developed using BC supports have been carried out. Bacterial cellulose loaded with bone marrow mesenchymal stem cells also represents an innovative resource for developing scaffolds and quality testing techniques for bone reconstruction materials. A study using horse stem cells proved BC’s cell adhesive and life supporting platform properties [88]. Due to it being non-cytotoxic, BC may function as a durable scaffold in the slow-healing processes [89]. Outstanding studies were carried out to improve and discover new methods for diagnosing neurodegenerative diseases. Results showed that neuroblastoma cells (SH-SY5Y) attached and proliferated on a bacterial nanocellulosic 3D-scaffold, resulting in mature neurons. This special model was designed for the investigation of neurodegenerative disease mechanisms, paving the way for discovering new treatments for neurological conditions [90].
Another branch of the biomedical field where BC occupies a central place is in the development of wounds dressings for the treatment of lesions of different etiologies (burns, chronic skin ulcers, surgical incisions, and traumatic wounds) [91]. BC exhibits a high capacity to maintain an optimal moisture at the lesion site, to absorb wound exudates, to allow a good exchange of gases, to provide thermal isolation, and to prevent a strong adhesion to the skin tissue. BC-based wound dressings supply an excellent protection against contamination and infection, reducing the occurrence of local pain and inflammation. Therefore, all these advantages lead to an increase in skin restoration and re-epithelialization, accelerating the wound healing process [92].
Along with the biomedical applications of BC presented above, BC can also be a promising material for controlled drug delivery [93]. Thus, BC can be loaded with various drugs, such as benzalkonium chloride, tetracycline, ibuprofen, diclofenac, paracetamol, propranolol, lidocaine hydrochloride, caffeine, silver sulfadiazine, and amoxicillin [92,94].
The second direction of applicability for bacterial cellulose is in the pharmaceutical domain. The use of intelligent technology in the pharmaceutical industry represents the future in terms of delivering active moieties to specific sites of absorption and action. Among these, Pickering emulsions have gained interest in recent years due to their versatility and possible application in the pharmaceutical and cosmetic industry. Recent approaches to these disperse systems include liquid marbles as precursors [95] and cellulose nanofibers/nanocrystals as stabilizers. Investigations in this direction include the study of oil droplets trapped by cellulose nanocrystals. Further studies revealed that the Pickering emulsions can be stabilized using TEMPO-oxidized (2,2,6,6-tetramethyl piperidine oxide) bacterial cellulose nanofibers. The assessment of the quality parameters of these emulsions indicates stability and viscoelasticity, and therefore, great potential for its use in new drug delivery systems [96].
The third direction of applicability of bacterial cellulose is in the biotechnological industry [97]. The petrochemical-based industry flourished for many decades, but nowadays a migration towards a bio-based ‘green’ economy is unfolding. This implies that avoiding natural material exploitation comes first when developing quality bioproducts to be obtained according to modern world requirements. In recent decades, the researchers’ interest was focused on BC as a promising ecological biomaterial due to its eco-friendly production process. Furthermore, BC possesses adequate mechanoelectrical and electromechanical transduction characteristics [98]. Following this direction, biofibers are currently important candidates as reinforcements for numerous applications of polymer composites [99]. Due to its suitable electrical properties and its renewable capacity, BC has a high potential for use in the expansion of new biosensors, wearable electronics, biomedical and energy storage devices, electrodes, and supercapacitors [100]. Biosensors have a large applicability in tissue engineering and regenerative medicine. Among them are enzymes, receptors, and antibodies. These devices are broadly used in bioanalysis because they can verify the biological signals in real time, informing health care providers about the current health status of the patient, helping the medical team to discover the disease in time and to initiate the treatment [101].

3. Bacterial Cellulose Composites—Important Emerging Materials for Biomedical Design and Other Impacting Applications

It is well known that ‘composites’ represent an umbrella term, defining a class of compounds obtained by placing two or more types of materials together: one plays the role of the matrix and the other is the reinforcement. The final product meets high-quality standards when compared to the raw materials it is made of [102].
Taking into account the aforementioned aspects and the fact that bacterial cellulose is a 3D-structured network, BC may function as a matrix to ‘trap’ other potential compounds [103]. Thus, microbial cellulose composites came to life, aiming to improve and eliminate some of the components’ disadvantages or to adapt a range of properties, depending on the desired final product [104]. The direct addition of various materials applied to BC results in a ‘combination product’ called BC composite, as illustrated in Figure 3.
Bacterial cellulose presents a high capacity to act as a host for some oxide nanoparticles (metals, carbon derivates, minerals). The many possibilities for using BC as a mesh for obtaining composites are becoming popular, and most of them involve other known polymers [105].

3.1. Biopolymers for Tailoring Bacterial Cellulose-Based Composites

The term ‘polymer’ is ubiquitous and defines a widespread class of materials used in many industrial areas. As many would say, we live in a so-called ‘polymer age,’ which is exhibited by several changes that a polymer can suffer, such as physical, chemical, thermal, or photochemical modifications [106]. The expanding role of these macromolecules in our lives is the reason why, during the last decades, scientists became interested in developing value-added products based on polymers. Currently, the main concern is obtaining biocompatible polymers via ecological and economical techniques. These biopolymers are environmentally friendly with low toxicity, making them suitable for multiple applications [107]. The waste disposal problem arising from industrial expansion is one of the main reasons why self-biodegradable materials became a necessity. Thus, in the 1980s, biodegradable plastics were created, originating from natural and synthetic polymers [108,109]. The newly designed materials exhibited the characteristic of self-biodegradation in the presence of living microorganisms which act in collaboration with chemical factors or enzymes [110].
Currently, worldwide efforts concern preference of biodegradable materials over non-biodegradable ones. The scope is to diminish pollution and uncontrolled waste. Polymers were once considered pollutant agents and most of them still maintain their ‘bad reputation’ [111]. As it is well known, a polymer is composed of repetitive similar/non-similar entities, called monomers, which connect through non-covalent bonds, generating the polymolecular entities [112]. A brief classification of the most widespread polymers includes cellulose, collagen, chitosan, and hyaluronic acid as natural polymers [113]; cellulose derivatives (sodium carboxymethylcellulose (NaCMC), hydroxyethylcellulose (HEC), hydroxypropylcellulose (HPC), hydroxypropylmethylcellulose (HPMC), methylcellulose (MC), and ethylcellulose (EC)) as modified natural polymers obtained by modifying the pure cellulose through the etherification reaction with alkyl groups [114]; and poly(vinyl alcohol) (PVA), polyvinylpyrrolidone (PVP), and Carbopol® as synthetic polymers [115].
All biopolymers stand out because of their remarkable previously discovered applications, but also distinguish themselves by perspectives of these properties, which should be brought to light. Among them, this paper points out a few innovative applications. The focus will be placed on emerging materials in the biomedical field (tissue engineering, wound dressings, drug delivery systems, regenerative medicine) where biopolymers act as precursors [116]. Thus, these biomaterials present biological functions which allow the healing of any impaired organ or tissue of the human body [117].
Due to the excellent physicochemical and biological properties of all types of biopolymers (natural, modified natural polymers, and synthetic), they can combine with bacterial cellulose to develop new composites with results that surmount their drawbacks and extend their applications [102]. BC-based composites have a large applicability in the wound healing process as wound dressings, as well as in tissue engineering, implants (bone, cartilage, cornea, teeth), 3D-bioprinting, the treatment of cardiovascular and neurological diseases, drug delivery, biosensors, electronics, and biofuels [102,118], as will be described below.
Cellulose, one of the main natural polymers, may experience different structural changes through specific processes to attain better quality (to improve its water solubility) and to extend its applicability in multiple domains [119]. Among the methods applied, lyophilization was mentioned in the literature data as a method to increase the porosity of a cellulose matrix for favorable bone regeneration [120]. Mello et al. obtained lyophilized cellulose and assessed it as a wrapper for peripheral nerve injuries (sectioned sciatic nerves) in animal studies. It was demonstrated that lyophilized cellulose caused a moderate fibrous reaction when implanted in peripheral nerve lesions with loss of substance. It proved to be effective as protection in those lesions in the presence of an inserted neural graft. Axons regrowth was reported, along with motor response after a period of recovery [121].
Other applications that indicate the use of cellulose-based materials pertain to the restorative medicine field [122]. One such material is oxidized regenerated cellulose (ORC), a natural biopolymer with carboxyl groups, which results through the chemical oxidation of pure cellulose. This material possesses several optimal characteristics (non-toxicity, biodegradability, antibacterial activity, and biocompatibility); thus, ORC has multiple medical uses, but it is principally a hemostatic agent [123]. A retrospective study on patients that underwent skin graft reconstruction treatments used ORC and a collagen-based composite loaded with silver. Results indicated a reduction in pain medication usage during healing, along with a decreased necessity for dressing replacement [124].
Collagen is a natural polymer known as the major constitutional protein found in human tissues. It makes up the extracellular matrix of connective tissues. Due to its high biocompatibility, it has been used in various biological applications, including as a wound dressing material, part of drug delivery systems, and materials-based scaffolds for tissue engineering [125]. Due to its suitable properties, collagen can be combined with other polymers, for example, with dextran in a spongious matrix loaded with flufenamic acid. Release profiles were obtained from animal testing: gradual delivery of the anti-inflammatory moiety accelerated the wound healing process and also enhanced re-epithelialization. The aim of designing these matrices was to reduce burn lesion progression, pain, and inflammation. The exact mechanism involved in tissue restoration through these particular systems will require further investigations [126].
Along with the above association (collagen and dextran), collagen can be also combined with BC for possible applications in the biomedical field. Different forms of collagen (hydrolyzed, solution, gel) embedded onto BC sheets increased their quality in terms of thermal stability and improved mechanical properties over plain lyophilized BC [127]. New research is constantly being carried out to guide collagen and other natural polymers towards inclusion in high-quality wound dressings. Rheological parameters in correlation with biological behavior and structure were studied for a spongious collagen-dextran-zinc oxide (50%) composite. Important perspectives for skin regeneration and antibacterial properties were indicated [128].
Moreover, collagen, the main component of bone tissue, is included in numerous composites (with BC), which are proposed as regeneration processes enhancers [77,129]. BC-collagen scaffolds impregnated with human umbilical cord blood-derived mesenchymal stem cells successfully functioned in osteogenic differentiation. Subcutaneous transplantation of these scaffolds enabled prolific neovascularization in early bone regeneration [130]. Vascular endothelial growth factor was added to BC scaffolds in studies carried out on femoral fractures in mice and rabbits, proving promotor properties in vascularization, ossification, and maturation of newly developed bones [130,131]. Playing a role similar that of plain BC acting as a separator between tissues in oral implants recovery, BC-collagen is applied as a carrier material for the osteogenic growth peptide with the goal of conserving bone defect space during the healing process, as well as enabling hyperplasia [132]. Moraes et al. developed a hydrogel dressing based on BC and collagen and compared it with a commercial product regarding their effect on the healing of rat dorsum. In vivo studies showed better skin healing using the newly designed hydrogel on day 7 after surgery [133].
Apart from physical contact therapies in wound healing and scaffolds, research groups also studied drug absorption and release mechanisms through porous microspheres based on collagen and bacterial cellulose. Absorption equilibriums analysis indicates promising applications for this system in future active moiety delivery systems for wounds [134].
Along with the natural polymers mentioned above, chitosan is also widespread in nature, second only to cellulose. Its origin is in the natural chitin. Chitosan exhibits many excellent properties, such as biocompatibility, non-toxicity, biodegradability, particular solubility, along with antimicrobial [135,136], antioxidant [137], antiviral, and antifungal effects [138]. Chitosan is well known as an excipient for drop-type ophthalmic products [139], and also for complex entities like liposomes [140], microemulsions [141], hydrogels [142], and implants [143]. Due to its excellent characteristics, chitosan can be used as a scaffold alone or in association with other polymers to develop new materials with promising biomedical uses. For example, chitosan in a gel form made was included as part of a drug delivery system destined for periodontal diseases treatment through intra-pocket drug release. As such, one proposed approach included chitosan-based formulations containing a chemotherapeutic agent (metronidazole benzoate) and an antibiotic (tetracycline hydrochloride). The optimum chitosan concentration was established through kinetic profile analysis. Thus, the gel with 3% w/w chitosan represents an excellent local treatment for periodontitis [144].
Moreover, ophthalmic pharmaceutical forms represent a challenge in terms of formulation and organ specificity. Apart from the active ingredients, auxiliary components are the ones entitled to encompass these boundaries. Chitosan won an important position in this direction due to its biodegradability, bioavailability, and permeation enhancement ability. Moreover, its antibacterial and antifungal properties, along with its intrinsically adhesive nature, promote chitosan’s inclusion in modern ophthalmic drug delivery systems. Since most in situ ophthalmic chitosan gels commonly deliver only one active substance, future investigations will try to incorporate more active ingredients, paving the way to attaining a local synergistic action [145,146,147].
In the meantime, chitosan can also be combined with BC. Thus, it was included in BC-based composites with various applications. The literature data included a BC-chitosan film that was compared to plain BC and other hydrocolloid films (Tegaderm®). It was demonstrated that BC did not dehydrate wounds but maintained a suitable moist healing environment with good permeability. The BC-chitosan film enabled skin regeneration and provided a better wound-healing effect [148]. Bacterial cellulose and chitosan, along with ciprofloxacin, were successfully integrated into a patch with dual antibacterial properties [149]. The mixture of BC, chitosan, and carboxymethylcellulose led to an antimicrobial film with a higher tensile strength and water vapor transmission rate [150].
In the category of the modified natural polymers, a central place is occupied by the cellulose derivatives (CMC, HPMC, MC, HEC, HPC, EC) that are of great interest, mainly in the food industry [151]. Nowadays, every industry is trying hard to maintain a standard of low waste, suitable economy, and high quality. The food industry concentrated its efforts on the development of cellulose derivatives as proper materials for food packaging and freshness maintenance due to their antimicrobial effect [152]. In this direction, many tests were performed, such as the scanning electron microscopy (SEM) analysis on gluten networks, showing that CMC is suitable to be used as a flour dough rheology regulator [153], whereas HPMC functions as a texture enhancer for whipped cream [154]. Following the idea of improving food preserving methods, methylcellulose-coated eggs exhibited promising shelf-life freshness when compared to eggs with uncoated shells [155]. Another functional food interface was developed targeting antioxidant activity using methylcellulose plasticized films, showing efficacy in preserving tocopherol content in walnut oil [156].
CMC also qualifies as an important polymer for use in the biomedical field due to its biocompatibility with human skin, its biodegradability, and its non-toxicity. Its advantages also include its high-water uptake, which is very important in ensuring a favorable environment for re-epithelialization [157]. NaCMC has a great ability to combine with other polymers to generate new composites showing suitable applicability in the biomedical field. For example, to meet all of the qualitative requirements of wound healing materials, spongious matrices comprised of collagen-NaCMC-mefenamic acid were developed and tested. Ninety-five percent of the anti-inflammatory agent was released. Along with favorable degradation and swelling ability, these hybrid matrices were proposed for further in vivo and in vitro testing [125]. Moreover, this combination of collagen and NaCMC and the same anti-inflammatory drug, mefenamic acid, was comparatively tested on rats in terms of its burn healing efficacy. Results indicated valuable effects in the hemostasis and inflammation stages, accelerating wound healing through the reduction of pain and minimal scar formation [158]. Local treatment for burns and soft-tissue injuries also included a multiparticulate system based on a collagen-dextran matrix embedded with flufenamic acid. Apart from the polymeric matrix, it consists of microcapsules based on gelatin-NaCMC-alginate with an embedded anti-inflammatory drug. This model was proposed as a promising design for future similar applications [159]. CMC can also be blended with BC. Thus, Pavaloiu et al. designed a new composite hydrogel loaded with ibuprofen sodium to study its drug release and swelling characteristics. It was found that the mechanism of swelling is controlled by pseudo-Fickian diffusion [160]. Juncu et al. formulated composite films based on NaCMC, BC, and ibuprofen sodium. The swelling behavior was studied using non-linear diffusion. The main results showed that the drug delivery depends on the content of BC; thus, the increase in the concentration of BC led to a decrease in the ibuprofen release rate [161].
Of special interest is the association between sodium alginate and BC. Cartilage restructuring properties were discovered in the BC-alginates double-layered composites inoculated with human nasal septal chondrocytes. In vitro culture revealed the stents’ potential as a model in treating severe auricle defects, as new healthy cartilage was formed through the porous layer of the biocomposite [162]. Porous sponges were obtained by combining BC and sodium alginate and cross-linking this with CaCl2 solutions. The sponge is conceived as a tear resistant, daily removed dressing for covering oral cavity surgical wounds [163]. The BC-alginate-chitosan-copper composites were confirmed to be suitable for use as biocompatible wound dressings due to their antibacterial activity against Escherichia coli and Methicillin-resistant Staphylococcus aureus [164]. Alginates were included, along with bacterial cellulose, in composite films loaded with silver sulfadiazine, proving cytotoxicity and an extended antimicrobial spectrum (Escherichia coli, Staphylococcus aureus, and Candida albicans) [165]. Kim et al. designed a BC and alginate-based nanocomposite, with normal spherical shapes and size, that showed high biocompatibility, biodegradability, optimal capacity to absorb water, higher crystallinity, and greater surface area. All these suitable properties enhanced the field of use for this nanocomposite, with promising applications in the biomedical, pharmaceutical, and biocatalytic industries [166].
Included in the category of synthetic polymers are poly(vinyl alcohol) (PVA), polyvinylpyrrolidone (polyvidone) (PVP), polyacrylic acid (Carbopol®), poly(ethylene glycol) (PEG), polyacrylamide, and polyurethanes, which are hydrophilic substances with swelling properties in the presence of biological fluids, thus generating hydrogels [113,115]. Two of them, PVA and PVP, are widely used in ophthalmic formulations (contact lenses and artificial cornea) [167,168] due to their optimal biocompatibility, biodegradability, capacity to form films, transparency, proper viscosity, excellent ocular mucosa penetration, and eye contact [169,170,171]. Moreover, PVA exhibits an appropriate capacity to blend with other polymers such as collagen to develop a novel matrix for tissue regeneration. The matrix was embedded with indomethacin as an anti-inflammatory molecule, resulting in a stable biohybrid. Kinetic parameters indicated an initial burst release of indomethacin, followed by slow drug delivery. Thus, the proposed spongious delivery systems presented promising results for replacing classical formulations used for tissue recovery purposes [172]. Moreover, PVA can also be combined with BC, creating a new composite that controls drug-delivery rates for anti-inflammatory substances. The system included poly(vinyl alcohol), chitosan, and bacterial cellulose; experimental results proved their possible use as biocarriers [173]. A new nanocomposite was obtained by blending BC with PVA and magnetite nanoparticles, which exhibited proper characteristics for use in the development of smart electronic devices [174].
The major applications of bacterial cellulose-based composites are highlighted in Figure 4.

3.2. Applications of Bacterial Cellulose-Based Composites

Bacterial cellulose is considered an efficient substance carrier. Due to its architecture, which resembles a porous network, it acts as an intermediary between the wound and an antimicrobial agent embedded in its structure. It is also a physical barrier against infectious agents. Many composites were created starting with these structural advantages. One example is the BC-NBG (nano-bioactive glass) composite which proved the synergistic antibacterial action of the two components. Laboratory tests showed antibacterial properties against Escherichia coli, Salmonella typhymurium, Pseudomonas aeruginosa, Klebsiella pneumonia, Bacillus subtillis [175]. It is also important to emphasize that the field of biomedicine has much to gain from attributes of bacterial cellulose, including high compatibility with the human organism and cost-effectiveness. Antibacterial composites include, among others: MH-BC (from mulberry leaves fermented in an acid hydrolysate fermentation medium) with proven perspectives in regenerative medicine [176].
Experimental data also showed that thymol-enriched bacterial cellulose has in vitro antibacterial activity against infectious bacteria, with prevalent potential for use in burned skin therapies. Cell viability and fibroblast proliferation were analyzed and the results showed an increase in the protection and coverage of damaged and recovering tissues. In vivo results indicated that wound closure and re-epithelization were not only enabled, but also accelerated [177]. Other research data revealed that biocomposites containing BC and Bacillus subtilis were reported as efficient and promising wound dressings, enabling full-thickness wound healing [178].
Bacterial cellulose and its composites also have effective applications in reconstructive medicine. The literature data indicated that accelerative wound healing processes and urinary reconstruction using angiogenesis promoters were achieved by using a combination of bacterial cellulose and urine-derived stem cells [179]. The potential of this association is noteworthy because of its telomerase activity, revealing the markers that are found on the surface of mesenchymal stem cells. Advantages compared to traditional wound dressings include an absence of secondary damage to the tissue and the absence of exudate accumulation. Bacterial infections are prevented due to bacterial cellulose being a skin substitute with high water-retaining capability [180].
Bacterial cellulose-based biomaterials used for wound healing are of major importance, particularly for the treatment of open injuries, severe burns, basal carcinoma, dermal abrasions, and chronic ulcers. Extensive tissue destruction or even severe infections may be triggered if wounds are not treated properly. Numerous studies introduced bacterial cellulose as an excellent dressing material. Many such products are already on the market: Biofill®, XCell®, Bioprocess®, Nanoderm® [181], and also Cellumed® (veterinary use) [182]. The literature data include reports of patients with second-degree wounds who exhibit faster healing when using BC-derived dressings compared to conventional products [183]. Wound dressing moisture balance is maintained, whilst skin is allowed to breathe and pain is reduced. Moreover, recent studies carried out on animals proved that wound dressing materials containing BC reduced the inflammatory response and improved wound healing and regeneration [184]. Other investigations revealed that the BC wound dressing materials exhibited superior covering properties for all facial contours and increased the cohesion levels on the mouth and nose regions, compared to other dressings; they also promoted a high moisture level in the wound, pain reduction, re-epithelialization acceleration, and a reduction in the appearance of scars [184,185,186]. Moreover, histopathological experiments proved that thick-BC wound dressings induced superior capillary formation, tissue regeneration, and cell proliferation compared to thin ones [187]. Clinical applications of BC-based wound dressing materials have the potential to replace classical gauze materials, as shown by experiments carried out on rat models [188].
As presented previously, a wound, and especially a burn infection, is an important aspect to consider during therapy because of its many triggered limitations. Pathogenic microorganism adhesion and further proliferation in the wound should be reduced as much as possible. BC-based dressings comply with these demands, the only difference being that no antibacterial protection is offered unless special treatments using organic (Ag, CuO, ZnO) or inorganic (lysine) agents are applied to the microbial cellulose fibers [188]. Bacteria are killed by impairing their main metabolic processes (respiration, nutrition) or by changing their cell wall structure and its normal function [189,190,191].
BC composites have been proven to make important contributions in the wound healing process, with previously demonstrated beneficial effects on hemostasis, inflammation, proliferation and remodeling phases of injury recovering. Even so, controversial aspects remain to be clarified, as many tissue healing mechanisms, especially scar formation and full recovery, are discordant [192,193].
Along with reconstructive medicine where wound healing processes benefit from bacterial cellulose composites, tissue engineering represents another field in which BC gains terrain compared to other materials due to its structured and porous 3D-network, biocompatibility, biodegradability, mechanical properties, and high power to retain large amounts of biological fluids [194]. Tissue engineering represents a new and demanding field of scientific exploration; it is an extensive multidisciplinary domain because it requires information from biology, chemistry, medicine, physics, and especially from engineering. The main purpose of tissue engineering is to expand several biological substitutes that can contribute to the anatomical and functional restoration, reconstruction, and increase in any human body tissue [195]. A biomaterial offers support and proper growth conditions in collaboration with other specific factors, promoting regeneration itself. Thus, the scaffold’s performance depends on its biocompatibility in terms of cellular adhesion and surface development, making the biopolymer responsible for cellular behavior (adherence, proliferation, migration) [196].
The ideal material destined for bone tissue engineering should exhibit certain properties: mechanical characteristics analogous to bone tissues, the ability to support the proliferation and differentiation of cells, the tendency to establish the deposition of the extracellular matrix [197], biocompatibility to support cellular interactions and tissue growth, biodegradability, absorbability, and last but not least, innocuity. Cumulative properties such as crystallinity and purity, in comparison to commonly used materials, promote bacterial cellulose as a superior qualitative medical material [198]. BC represents an ideal biopolymer that has a high capacity to simulate natural collagen due to its excellent properties mentioned above. Therefore, it is a remarkable candidate for use in bone restoration. Such an example is the mixture between BC and hydroxyapatite, a natural polymer that participates in the process of bone ossification. A new biomineralized BC scaffold has been designed with CMC as an activator of the BC surface; this novel formulation is a promising bone scaffolding material that requires further investigation [199].
Studies reported numerous BC-hydroxyapatite composites obtained through different technological processes as having various stoichiometric values, depending on the location and tissue for which they are intended. Bacterial cellulose was included, along with Fe3O4 and hydroxyapatite, in the scaffolds. Properties of the composite resemble human trabecular and cancellous bones. Further, in vivo investigations regarding this material’s osteogenic properties are needed for the scaffold to gain terrain against other classical prosthetic materials frequently used in dentistry [200].
Classical polymeric scaffolds do not retain high strength stability over time. To overcome this disadvantage, many techniques are still developing, mostly to obtain quality scaffolds for bone- and cartilage-recovering therapies. Polysaccharide scaffolds are mentioned in the literature as being compliant with physiological conditions, showing superior osteoblast adhesion and progressive bone mineralization compared to other control scaffolds (poly(lactic acid-glycolic acid)) [201]. Studies on nanocellulose scaffolds with collagen showed a greater adhesion and phenotype maintenance of cultured human osteoblasts, reflected by increased levels of alkaline phosphatase and mineral deposition compared to the control polyester micro-nano structured scaffolds of identical pore properties. These scaffolds are competitors for other polyester-based scaffolds used in bone restoration [202].
As per the case of bone tissue grafting, cartilage surgical repair also consists of similar procedures: the autografting of chondrocytes or osteochondral plugs. BC and its composites have been tested as cartilage replacement materials. One example includes the BC-poly(vinyl alcohol) composite, which exhibits the potential for use in the orthopaedical field as cartilage or intervertebral disc-replacement material [193].
Nowadays, materials with extreme wetting properties gain scientists’ attention because these materials present multiple ascending applications. Following the trend of developing biomimetic materials with special surface properties (for example superhydrophobic materials structurally resembling the lotus leaf [203,204]), bacterial cellulose became of interest to scientists working on hybrid materials. Thus, biocomposites with improved interfacial wettability were created by chemical cross-linking with oligopeptides, promoting tissue repair, which is of high importance in regenerative medicine [205]. Recent investigations pointed out the advantages of using bacterial cellulose in tympanic grafts, which enhances the surgical procedure by improving the healing ability after the graft is accepted by the organism.
One of the most studied and yet not fully understood systems, the cardiovascular domain, also benefits from the emerging application of bacterial cellulose. Hypertension and other heart-related pathologies are mainly caused by clogged or pathologically destroyed blood vessels. Scientists have developed polyester or polytetrafluoroethylene artificial grafts. Their inconvenience includes thrombi formation and appropriate capillarity, which is hard to obtain [45]. Thanks to its biocompatibility, high porosity, incredible mechanical properties, strength, and elasticity, BC became very popular as a vascular graft component. BC embedded with graphene oxide nanosheets is another composite developed to replace even small-diameter vessels [206]. Bypass implants registered a notable quality improvement when the product BActerial SYnthesized Cellulose (BASYC) was developed. Among its advantages, there are a few that stand out: mechanical strength in a moist state, smoothness of the interior lumen, and increased moisture preservation. Animal testing proved its success in the replacement of blood vessels using this material [45,207]. Other experiments regarding hemodynamics and physiological phenomena at the implantation site were carried out and these underlined the efficacy of the BC biosynthetic blood vessel precursors [186,208].
Even though BC and its composites proved to possess superior qualitative properties compared to conventional materials when referring to biomedical applications, scientists are still searching for new strategies to develop better composites. Recently, Gengiflex® (based on BC) and Gore-Tex® (based on polytetrafluoroethylene) membranes were tested for applicability in the dental field. Both membranes stimulated the bone expansion. After three months, the efficacy of the two products was compared. Thus, Gore-Tex® showed higher efficacy than Gengiflex,® and it is a promising candidate for healing osseous deficiencies [186].
As was previously stated, industrial fields gravitate towards quality materials and rely on obtaining them by combining well-known and studied materials. Following this flow, new composites were designed using bacterial cellulose. These products allow quality improvement in terms of mechanical, optical, and water absorption properties over plain bacterial cellulose [209,210].
Along with the applications of many of the newly designed composites highlighted above, a structured presentation of other bacterial cellulose composites, combinations of bacterial cellulose and various bioactive agents, and their biomedical applications are illustrated in Table 1. The proposed classification of these biomaterials and their prospective uses in the medical field were designed depending on tissue type and organ class. The table includes composites ranging from simple to complex structures, along with a wider range in terms of applications.
Along with all the aforementioned applications of BC, this biopolymer is also a promising biomaterial for biological diagnosis. To this end, Qin et al. developed a ‘living membrane’ system, which comprises BC and Escherichia coli bacterial strains, whose main purpose was to identify biologically triggered molecules [250]. Moreover, BC has a great potential for applicability in personalized regenerative medicine. A double network, biphasic Janus BC-conducting polymer composite hydrogel showed a biocompatible and electroactive behavior allowing the growth, spread, and migration of normal fibroblasts [250]. In recent years, the attention of the researchers has been focused on the development of BC-based biosensors used as monitoring devices. Marques et al. designed a BC-Ag nanocomposite-based biosensor intended to analyze 1-phenylalanine, 1-glutamine, and 1-histidine using Surface-enhanced Raman Scattering lamellas and 2 analytes (thiosalicylic acid and 2,2-dithiodipyridine) [251]. An enzymatic biosensor used for the amperometric determination of glucose has been developed using BC nanofibers and gold nanoparticles [247,252]. For the amperometric determination of the glucose oxidase reactions, Eisele et al. have created an external lamella in glucose biosensor with an extended range based on the use of BC and polyamide [253].

4. Conclusions and Future Perspectives

This present review has focused on the exposure of the synthesis, fundamental properties of bacterial cellulose, and its multiple applications in diverse domains, from food packaging to biotechnological, biomedical, and pharmaceutical industries. Bacterial cellulose is a valuable polysaccharide synthesized by a wide range of non-pathogenic bacteria under special culture conditions. This fascinating biopolymer possesses particular physicochemical, mechanical, and biological properties, such as eco-friendliness, biocompatibility, biodegradability, non-toxicity, a 3D-porous structure, optimal viscoelasticity, and tensile strength, an adequate ability to retain a large amount of water, moldability, along with higher crystallinity and purity than pure cellulose. It has been shown that bacterial cellulose can manifest a therapeutic effect on different anatomical parts of the human body, alone or in combination with several biopolymers and bioactive agents. Consequently, bacterial cellulose can act as an excellent medical material for the development of new skin lesion and dental dressings, drug delivery devices, oral implants, bone restoration or replacement products, local chemotherapy treatments, and cardiovascular interventional therapies. Presently, the attention of the researchers is centered on other captivating biomedical applications of bacterial cellulose, such as the development of biosensors, biological diagnoses, contact lenses, and nerve and ophthalmic tissue engineering. Besides all of the engaging uses illustrated above, bacterial cellulose is a promising material with high relevance for future use in the food, paper, and textile industries; acoustic membranes, supercapacitors, optical, stimuli-responsive, and catalytic materials; energy storage, oil refining, pollution control; and aerogels (reusable polymer networks that trap and dispense metal nanoparticles (Cu, Ni) used as catalysts in the electronic fields, due to their optimized properties).

Author Contributions

The authors had equal contribution. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was financially supported by the Carol Davila University of Medicine and Pharmacy, Bucharest, through Contract No. CNFIS-FDI-2021-0300 and RDI Capability consolidation at the institutional level of the multidisciplinary research teams involved in the sustainability of the UMFCD priority research directions.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vasilevich, A.; de Boer, J. Robot-scientists will lead tomorrow’s biomaterials discovery. Curr. Opin. Biomed. Eng. 2018, 6, 74–80. [Google Scholar] [CrossRef]
  2. Chen, H.; Yao, Y. Progress of biomaterials for bone tumor therapy. J. Biomater. Appl. 2021, 36, 945–955. [Google Scholar] [CrossRef]
  3. Conticello, V.; Hughes, S.; Modlin, C. Biomaterials Made from Coiled-Coil Peptides. In Fibrous Proteins: Structures and Mechanisms; Parry, D.A.D., Squire, J.M., Eds.; Subcellular Biochemistry; Springer International Publishing: Cham, Switzerland, 2017; Volume 82, pp. 575–600. [Google Scholar]
  4. Diba, M.; Spaans, S.; Ning, K.; Ippel, B.D.; Yang, F.; Loomans, B.; Dankers, P.Y.W.; Leeuwenburgh, S.C.G. Self-Healing Biomaterials: From Molecular Concepts to Clinical Applications. Adv. Mater. Interfaces 2018, 5, 21. [Google Scholar] [CrossRef]
  5. Biswal, T.; BadJena, S.K.; Pradhan, D. Sustainable biomaterials and their applications: A short review. In Proceedings of the National Conference on Trends in Minerals & Materials Technology (MMT), Bhubaneswar, India, 30 October 2019; pp. 274–282. [Google Scholar]
  6. Sun, Q.; Qian, B.; Uto, K.; Chen, J.; Liu, X.; Minari, T. Functional biomaterials towards flexible electronics and sensors. Biosens. Bioelectron. 2018, 119, 237–251. [Google Scholar] [CrossRef] [PubMed]
  7. Kargozar, S.; Ramakrishna, S.; Mozafari, M. Chemistry of biomaterials: Future prospects. Curr. Opin. Biomed. Eng. 2019, 10, 181–190. [Google Scholar] [CrossRef]
  8. Datta, L.P.; Manchineella, S.; Govindaraju, T. Biomolecules-derived biomaterials. Biomaterials 2020, 230, 41. [Google Scholar] [CrossRef]
  9. Paunescu, C.; Pitigoi, G.; Cosma, G.; Pituru, S.M.; Grigore, V.; Petrescu, S.; Mircica, M.L.; Radulescu, M.; Cosma, A.; Rezaee, R.; et al. Increasing Endurance in Physical Effort by Administration Of Inosine. Farmacia 2021, 69, 148–154. [Google Scholar] [CrossRef]
  10. Fenton, O.S.; Olafson, K.N.; Pillai, P.S.; Mitchell, M.J.; Langer, R. Advances in Biomaterials for Drug Delivery. Adv. Mater. 2018, 30, 29. [Google Scholar] [CrossRef]
  11. Ammarullah, M.I.; Afif, I.Y.; Maula, M.I.; Winarni, T.I.; Tauviqirrahman, M.; Akbar, I.; Basri, H.; van der Heide, E.; Jamari, J. Tresca Stress Simulation of Metal-on-Metal Total Hip Arthroplasty during Normal Walking Activity. Materials 2021, 14, 7554. [Google Scholar] [CrossRef]
  12. Jamari, J.; Ammarullah, M.I.; Saad, A.P.M.; Syahrom, A.; Uddin, M.; van der Heide, E.; Basri, H. The Effect of Bottom Profile Dimples on the Femoral Head on Wear in Metal-on-Metal Total Hip Arthroplasty. J. Func. Biomater. 2021, 12, 38. [Google Scholar] [CrossRef]
  13. Zhang, Z.P.; Gupte, M.J.; Ma, P.X. Biomaterials and stem cells for tissue engineering. Expert Opin. Biol. Ther. 2013, 13, 527–540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Chen, F.M.; Liu, X.H. Advancing biomaterials of human origin for tissue engineering. Prog. Polym. Sci. 2016, 53, 86–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Triplett, R.G.; Budinskaya, O. New Frontiers in Biomaterials. Oral Maxillofac. Surg. Clin. N. Am. 2017, 29, 105–115. [Google Scholar] [CrossRef] [PubMed]
  16. Yadav, S.; Gangwar, S. An Overview on Recent progresses and future perspective of biomaterials. In Proceedings of the 1st International Conference on Contemporary Research in Mechanical Engineering with Focus on Materials and Manufacturing (ICCRME), Lucknow, India, 6–7 April 2018. [Google Scholar]
  17. Jahangirian, H.; Lemraski, E.G.; Rafiee-Moghaddam, R.; Webster, T.J. A review of using green chemistry methods for biomaterials in tissue engineering. Int. J. Nanomed. 2018, 13, 5953–5969. [Google Scholar] [CrossRef] [Green Version]
  18. Matsumura, K.; Rajan, R. Oxidized Polysaccharides as Green and Sustainable Biomaterials. Curr. Org. Chem. 2021, 25, 1483–1496. [Google Scholar] [CrossRef]
  19. Heinze, T. Cellulose: Structure and Properties. In Cellulose Chemistry and Properties: Fibers, Nanocelluloses and Advanced Materials; Rojas, O.J., Ed.; Advances in Polymer Science; Springer: Berlin/Heidelberg, Germany, 2016; Volume 271, pp. 1–52. [Google Scholar]
  20. Wang, Y.G.; Wang, X.J.; Xie, Y.J.; Zhang, K. Functional nanomaterials through esterification of cellulose: A review of chemistry and application. Cellulose 2018, 25, 3703–3731. [Google Scholar] [CrossRef] [Green Version]
  21. Miao, J.J.; Pangule, R.C.; Paskaleva, E.E.; Hwang, E.E.; Kane, R.S.; Linhardt, R.J.; Dordick, J.S. Lysostaphin-functionalized cellulose fibers with antistaphylococcal activity for wound healing applications. Biomaterials 2011, 32, 9557–9567. [Google Scholar] [CrossRef]
  22. Wei, L.; McDonald, A.G. A Review on Grafting of Biofibers for Biocomposites. Materials 2016, 9, 303. [Google Scholar] [CrossRef]
  23. Gallegos, A.M.A.; Carrera, S.H.; Parra, R.; Keshavarz, T.; Iqbal, H.M.N. Bacterial Cellulose: A Sustainable Source to Develop Value-Added Products—A Review. BioResources 2016, 11, 5641–5655. [Google Scholar] [CrossRef]
  24. Haldar, D.; Purkait, M.K. Micro and nanocrystalline cellulose derivatives of lignocellulosic biomass: A review on synthesis, applications and advancements. Carbohydr. Polym. 2020, 250, 116937. [Google Scholar] [CrossRef]
  25. Fillat, A.; Martinez, J.; Valls, C.; Cusola, O.; Roncero, M.B.; Vidal, T.; Valenzuela, S.V.; Diaz, P.; Pastor, F.I.J. Bacterial cellulose for increasing barrier properties of paper products. Cellulose 2018, 25, 6093–6105. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, X.C.; Tang, S.J.; Chai, S.L.; Wang, P.; Qin, J.H.; Pei, W.H.; Bian, H.Y.; Jiang, Q.; Huang, C.X. Preparing printable bacterial cellulose based gelatin gel to promote in vivo bone regeneration. Carbohydr. Polym. 2021, 270, 13. [Google Scholar] [CrossRef] [PubMed]
  27. Fernandes, I.D.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]
  28. Chen, S.Q.; Meldrum, O.W.; Liao, Q.D.; Li, Z.F.; Cao, X.; Guo, L.; Zhang, S.Y.; Zhu, J.; Li, L. The influence of alkaline treatment on the mechanical and structural properties of bacterial cellulose. Carbohydr. Polym. 2021, 271, 9. [Google Scholar] [CrossRef] [PubMed]
  29. Santos, S.M.; Carbajo, J.M.; Quintana, E.; Ibarra, D.; Gomez, N.; Ladero, M.; Eugenio, M.E.; Villar, J.C. Characterization of purified bacterial cellulose focused on its use on paper restoration. Carbohydr. Polym. 2015, 116, 173–181. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, J.; Wang, S.; Jiang, L.; Shao, W. Production and characterization of antimicrobial bacterial cellulose membranes with non-leaching activity. J. Ind. Eng. Chem. 2021, 103, 232–238. [Google Scholar] [CrossRef]
  31. Tsai, Y.-H.; Yang, Y.-N.; Ho, Y.-C.; Tsai, M.-L.; Mi, F.-L. Drug release and antioxidant/antibacterial activities of silymarin-zein nanoparticle/bacterial cellulose nanofiber composite films. Carbohydr. Polym. 2018, 180, 286–296. [Google Scholar] [CrossRef]
  32. Cazon, P.; Velazquez, G.; Vazquez, M. Characterization of bacterial cellulose films combined with chitosan and polyvinyl alcohol: Evaluation of mechanical and barrier properties. Carbohydr. Polym. 2019, 216, 72–85. [Google Scholar] [CrossRef]
  33. Betlej, I.; Zakaria, S.; Krajewski, K.J.; Boruszewski, P. Bacterial Cellulose–Properties and Its Potential Application. Sains Malays. 2021, 50, 493–505. [Google Scholar] [CrossRef]
  34. Andrade, F.K.; Morais, J.P.S.; Muniz, C.R.; Nascimento, J.H.O.; Vieira, R.S.; Gama, F.M.P.; Rosa, M.F. Stable microfluidized bacterial cellulose suspension. Cellulose 2019, 26, 5851–5864. [Google Scholar] [CrossRef] [Green Version]
  35. Quan, V.M.; Li, B.; Sukyai, P. Bacterial cellulose modification using static magnetic field. Cellulose 2020, 27, 5581–5596. [Google Scholar] [CrossRef]
  36. Sukhtezari, S.; Almasi, H.; Pirsa, S.; Zandi, M.; Pirouzifard, M. Development of bacterial cellulose based slow-release active films by incorporation of Scrophularia striata Boiss. extract. Carbohydr. Polym. 2017, 156, 340–350. [Google Scholar] [CrossRef] [PubMed]
  37. Wahid, F.; Huang, L.-H.; Zhao, X.-Q.; Li, W.-C.; Wang, Y.-Y.; Jia, S.-R.; Zhong, C. Bacterial cellulose and its potential for biomedical applications. Biotechnol. Adv. 2021, 53, 107856. [Google Scholar] [CrossRef] [PubMed]
  38. Barja, F. Bacterial nanocellulose production and biomedical applications. J. Biomed. Res. 2021, 35, 310–317. [Google Scholar] [CrossRef]
  39. Aritonang, H.F.; Kamea, O.E.; Koleangan, H.; Wuntu, A.D. Biotemplated synthesis of Ag-ZnO nanoparticles/bacterial cellulose nanocomposites for photocatalysis application. Polym.-Plast. Tech. Mater. 2020, 59, 1292–1299. [Google Scholar] [CrossRef]
  40. Kaminski, K.; Jarosz, M.; Grudzien, J.; Pawlik, J.; Zastawnik, F.; Pandyra, P.; Kolodziejczyk, A.M. Hydrogel bacterial cellulose: A path to improved materials for new eco-friendly textiles. Cellulose 2020, 27, 5353–5365. [Google Scholar] [CrossRef] [Green Version]
  41. Amin, M.; Abadi, A.G.; Katas, H. Purification, characterization and comparative studies of spray-dried bacterial cellulose microparticles. Carbohydr. Polym. 2014, 99, 180–189. [Google Scholar] [CrossRef]
  42. Shoda, M.; Sugano, Y. Recent advances in bacterial cellulose production. Biotechnol. Bioprocess Eng. 2005, 10, 1–8. [Google Scholar] [CrossRef]
  43. Torres, F.G.; Arroyo, J.J.; Troncoso, O.P. Bacterial cellulose nanocomposites: An all-nano type of material. Mater. Sci. Eng. C 2019, 98, 1277–1293. [Google Scholar] [CrossRef]
  44. Khan, H.; Kadam, A.; Dutt, D. Studies on bacterial cellulose produced by a novel strain of Lactobacillus genus. Carbohydr. Polym. 2019, 229, 115513. [Google Scholar] [CrossRef]
  45. Saxena, I.M.; Brown, R.M. Biosynthesis of bacterial cellulose. In Bacterial NanoCellulose: A Sophisticated Multifunctional Material; CRC Press: Boca Raton, FL, USA, 2012; pp. 1–18. [Google Scholar]
  46. Falcao, S.C.; Coelho, A.R.; Evencio Neto, J. Biomechanical evaluation of microbial cellulose (Zoogloea sp.) and expanded polytetrafluoroethylene membranes as implants in repair of produced abdominal wall defects in rats. Acta Cir. Bras. 2008, 23, 184–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Klemm, D.; Schumann, D.; Udhardt, U.; Marsch, S. Bacterial synthesized cellulose—artificial blood vessels for microsurgery. Prog. Polym. Sci. 2001, 26, 1561–1603. [Google Scholar] [CrossRef]
  48. Lin, S.P.; Calvar, I.L.; Catchmark, J.M.; Liu, J.R.; Demirci, A.; Cheng, K.C. Biosynthesis, production and applications of bacterial cellulose. Cellulose 2013, 20, 2191–2219. [Google Scholar] [CrossRef]
  49. McNamara, J.T.; Morgan, J.L.W.; Zimmer, J. A Molecular Description of Cellulose Biosynthesis. Annu. Rev. Biochem. 2015, 84, 895–921. [Google Scholar] [CrossRef] [Green Version]
  50. Morgan, J.L.W.; Strumillo, J.; Zimmer, J. Crystallographic snapshot of cellulose synthesis and membrane translocation. Nature 2013, 493, 181–186. [Google Scholar] [CrossRef] [Green Version]
  51. Römling, U.; Galperin, M.Y. Bacterial cellulose biosynthesis: Diversity of operons, subunits, products, and functions. Trends Microbiol. 2015, 23, 545–557. [Google Scholar] [CrossRef] [Green Version]
  52. Ross, P.; Mayer, R.; Benziman, M. Cellulose biosynthesis and function in bacteria. Microbiol. Rev. 1991, 55, 35–58. [Google Scholar] [CrossRef]
  53. Gea, S.; Reynolds, C.T.; Roohpour, N.; Wirjosentono, B.; Soykeabkaew, N.; Bilotti, E.; Peijs, T. Investigation into the structural, morphological, mechanical and thermal behaviour of bacterial cellulose after a two-step purification process. Bioresour. Technol. 2011, 102, 9105–9110. [Google Scholar] [CrossRef]
  54. 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. 2021. [Google Scholar] [CrossRef]
  55. Kongruang, S. Bacterial Cellulose Production by Acetobacter xylinum Strains from Agricultural Waste Products. Appl. Biochem. Biotechnol. 2008, 148, 245–256. [Google Scholar] [CrossRef]
  56. Jung, J.Y.; Park, J.K.; Chang, H.N. Bacterial cellulose production by Gluconacetobacter hansenii in an agitated culture without living non-cellulose producing cells. Enzym. Microb. Technol. 2005, 37, 347–354. [Google Scholar] [CrossRef]
  57. Vasconcelos, N.F.; Andrade, F.K.; Vieira, L.D.P.; Vieira, R.S.; Vaz, J.M.; Chevallier, P.; Mantovani, D.; Borges, M.D.; Rosa, M.D. Oxidized bacterial cellulose membrane as support for enzyme immobilization: Properties and morphological features. Cellulose 2020, 27, 3055–3083. [Google Scholar] [CrossRef]
  58. Martínez Ávila, H.; Schwarz, S.; Feldmann, E.-M.; Mantas, A.; von Bomhard, A.; Gatenholm, P.; Rotter, N. Biocompatibility evaluation of densified bacterial nanocellulose hydrogel as an implant material for auricular cartilage regeneration. Appl. Microbiol. Biotechnol. 2014, 98, 7423–7435. [Google Scholar] [CrossRef] [PubMed]
  59. Pigaleva, M.A.; Bulat, M.V.; Gromovykh, T.I.; Gavryushina, I.A.; Lutsenko, S.V.; Gallyamov, M.O.; Novikov, I.V.; Buyanovskaya, A.G.; Kiselyova, O.I. A new approach to purification of bacterial cellulose membranes: What happens to bacteria in supercritical media? J. Supercrit. Fluids 2019, 147, 59–69. [Google Scholar] [CrossRef]
  60. Gayathri, G.; Srinikethan, G. Bacterial Cellulose production by K. saccharivorans BC1 strain using crude distillery effluent as cheap and cost effective nutrient medium. Int. J. Biol. Macromol. 2019, 138, 950–957. [Google Scholar] [CrossRef] [PubMed]
  61. Quintana-Quirino, M.; Morales-Osorio, C.; Vigueras Ramírez, G.; Vázquez-Torres, H.; Shirai, K. Bacterial cellulose grows with a honeycomb geometry in a solid-state culture of Gluconacetobacter xylinus using polyurethane foam support. Process Biochem. 2019, 82, 1–9. [Google Scholar] [CrossRef]
  62. Ye, S.; Jiang, L.; Su, C.; Zhu, Z.; Wen, Y.; Shao, W. Development of gelatin/bacterial cellulose composite sponges as potential natural wound dressings. Int. J. Biol. Macromol. 2019, 133, 148–155. [Google Scholar] [CrossRef]
  63. Iguchi, M.; Yamanaka, S.; Budhiono, A. Bacterial cellulose—A masterpiece of nature’s arts. J. Mater. Sci. 2000, 35, 261–270. [Google Scholar] [CrossRef]
  64. Nakagaito, A.; Iwamoto, S.; Yano, H. Bacterial cellulose: The ultimate nano-scalar cellulose morphology for the production of high-strength composites. Appl. Phys. A 2005, 80, 93–97. [Google Scholar] [CrossRef]
  65. Hirai, A.; Tsuji, M.; Horii, F. TEM study of band-like cellulose assemblies produced by Acetobacter xylinum at 4 °C. Cellulose 2002, 9, 105–113. [Google Scholar] [CrossRef]
  66. Gindl, W.; Keckes, J. Tensile properties of cellulose acetate butyrate composites reinforced with bacterial cellulose. Compos. Sci. Technol. 2004, 64, 2407–2413. [Google Scholar] [CrossRef]
  67. Torres, F.G.; Commeaux, S.; Troncoso, O.P. Biocompatibility of Bacterial Cellulose Based Biomaterials. J. Funct. Biomater. 2012, 3, 864–878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Atalla, R.H.; Vanderhart, D.L. Native cellulose: A composite of two distinct crystalline forms. Science 1984, 223, 283–285. [Google Scholar] [CrossRef] [PubMed]
  69. Dahman, Y. Nanostructured Biomaterials and Biocomposites from Bacterial Cellulose Nanofibers. J. Nanosci. Nanotechnol. 2009, 9, 5105–5122. [Google Scholar] [CrossRef]
  70. Lee, K.Y.; Buldum, G.; Mantalaris, A.; Bismarck, A. More than meets the eye in bacterial cellulose: Biosynthesis, bioprocessing, and applications in advanced fiber composites. Macromol. Biosci. 2014, 14, 10–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Bäckdahl, H.; Helenius, G.; Bodin, A.; Nannmark, U.; Johansson, B.R.; Risberg, B.; Gatenholm, P. Mechanical properties of bacterial cellulose and interactions with smooth muscle cells. Biomaterials 2006, 27, 2141–2149. [Google Scholar] [CrossRef]
  72. Vandamme, E.J.; De Baets, S.; Vanbaelen, A.; Joris, K.; De Wulf, P. Improved production of bacterial cellulose and its application potential. Polym. Degrad. Stab. 1998, 59, 93–99. [Google Scholar] [CrossRef]
  73. Gregory, D.A.; Tripathi, L.; Fricker, A.T.R.; Asare, E.; Orlando, I.; Raghavendran, V.; Roy, I. Bacterial cellulose: A smart biomaterial with diverse applications. Mater. Sci. Eng. R Rep. 2021, 145, 100623. [Google Scholar] [CrossRef]
  74. Seifert, M.; Hesse, S.; Kabrelian, V.; Klemm, D. Controlling the water content of never dried and reswollen bacterial cellulose by the addition of water-soluble polymers to the culture medium. J. Polym. Sci. Part A Polym. Chem. 2004, 42, 463–470. [Google Scholar] [CrossRef]
  75. Ciechańska, D. Multifunctional bacterial cellulose/chitosan composite materials for medical applications. Fibres Text. East. Eur. 2004, 12, 69–72. [Google Scholar]
  76. Sharip, N.; Yasim-Anuar, T.; Norrrahim, M.; Shazleen, S.; Nurazzi, N.M.; Sapuan, S.; Ilyas, R. A review on nanocellulose composites in biomedical application. In Composites in Biomedical Applications; CRC Press: Boca Raton, FL, USA, 2020; pp. 161–190. [Google Scholar]
  77. Pang, M.; Huang, Y.; Meng, F.; Zhuang, Y.; Liu, H.; Du, M.; Ma, Q.; Wang, Q.; Chen, Z.; Chen, L.; et al. Application of Bacterial Cellulose in Skin and Bone Tissue Engineering. Eur. Polym. J. 2019, 122, 109365. [Google Scholar] [CrossRef]
  78. Potzinger, Y.; Kralisch, D.; Fischer, D. Bacterial nanocellulose: The future of controlled drug delivery? Ther. Deliv. 2017, 8, 753–761. [Google Scholar] [CrossRef] [PubMed]
  79. Lee, D.; Lee, S.; Moon, J.-H.; Kim, J.; Heo, D.N.; Bang, J.; Lim, H.-N.; Kwon, I.K. Preparation of antibacterial chitosan membranes containing silver nanoparticles for dental barrier membrane applications. J. Ind. Eng. Chem. 2018, 66, 196–202. [Google Scholar] [CrossRef]
  80. An, S.-J.; Lee, S.-H.; Huh, J.-B.; Jeong, S.I.; Park, J.-S.; Gwon, H.-J.; Kang, E.-S.; Jeong, C.-M.; Lim, Y.-M. Preparation and Characterization of Resorbable Bacterial Cellulose Membranes Treated by Electron Beam Irradiation for Guided Bone Regeneration. Int. J. Mol. Sci. 2017, 18, 2236. [Google Scholar] [CrossRef] [PubMed]
  81. Yoshino, A.; Tabuchi, M.; Uo, M.; Tatsumi, H.; Hideshima, K.; Kondo, S.; Sekine, J. Applicability of bacterial cellulose as an alternative to paper points in endodontic treatment. Acta Biomater. 2013, 9, 6116–6122. [Google Scholar] [CrossRef] [PubMed]
  82. Lv, X.; Yang, J.; Feng, C.; Li, Z.; Chen, S.; Xie, M.; Huang, J.; Li, H.; Wang, H.; Xu, Y. Bacterial Cellulose-Based Biomimetic Nanofibrous Scaffold with Muscle Cells for Hollow Organ Tissue Engineering. ACS Biomater. Sci. Eng. 2016, 2, 19–29. [Google Scholar] [CrossRef] [PubMed]
  83. Trache, D. Nanocellulose as a promising sustainable material for biomedical applications. AIMS Mater. Sci. 2018, 5, 201–205. [Google Scholar] [CrossRef]
  84. Augustine, R.; Dan, P.; Hasan, A.; Khalaf, I.M.; Prasad, P.; Ghosal, K.; Gentile, C.; McClements, L.; Maureira, P. Stem cell-based approaches in cardiac tissue engineering: Controlling the microenvironment for autologous cells. Biomed. Pharmacother. 2021, 138, 111425. [Google Scholar] [CrossRef]
  85. Sebastião, M.J.; Serra, M.; Gomes-Alves, P.; Alves, P.M. Stem cells characterization: OMICS reinforcing analytics. Curr. Opin. Biotechnol. 2021, 71, 175–181. [Google Scholar] [CrossRef]
  86. Su, X.; Wang, T.; Guo, S. Applications of 3D printed bone tissue engineering scaffolds in the stem cell field. Regen. Ther. 2021, 16, 63–72. [Google Scholar] [CrossRef]
  87. Tronser, T.; Laromaine, A.; Roig, A.; Levkin, P.A. Bacterial Cellulose Promotes Long-Term Stemness of mESC. ACS Appl. Mater. Interfaces 2018, 10, 16260–16269. [Google Scholar] [CrossRef] [PubMed]
  88. Favi, P.M.; Benson, R.S.; Neilsen, N.R.; Hammonds, R.L.; Bates, C.C.; Stephens, C.P.; Dhar, M.S. Cell proliferation, viability, and in vitro differentiation of equine mesenchymal stem cells seeded on bacterial cellulose hydrogel scaffolds. Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 1935–1944. [Google Scholar] [CrossRef] [PubMed]
  89. Silva, M.D.A.; Leite, Y.K.D.C.; de Carvalho, C.E.S.; Feitosa, M.L.T.; Alves, M.M.D.M.; Carvalho, F.A.D.A.; Neto, B.C.V.; Miglino, M.A.; Jozala, A.F.; de Carvalho, M.A.M. Behavior and biocompatibility of rabbit bone marrow mesenchymal stem cells with bacterial cellulose membrane. PeerJ 2018, 6, e4656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Innala, M.; Riebe, I.; Kuzmenko, V.; Sundberg, J.; Gatenholm, P.; Hanse, E.; Johannesson, S. 3D Culturing and differentiation of SH-SY5Y neuroblastoma cells on bacterial nanocellulose scaffolds. Artif. Cells Nanomed. Biotechnol. 2014, 42, 302–308. [Google Scholar] [CrossRef]
  91. Ludwicka, K.; Kolodziejczyk, M.; Gendaszewska-Darmach, E.; Chrzanowski, M.; Jedrzejczak-Krzepkowska, M.; Rytczak, P.; Bielecki, S. Stable composite of bacterial nanocellulose and perforated polypropylene mesh for biomedical applications. J. Biomed. Mater. Res. Part B 2019, 107, 978–987. [Google Scholar] [CrossRef]
  92. Sharma, C.; Bhardwaj, N.K. Bacterial nanocellulose: Present status, biomedical applications and future perspectives. Mater. Sci. Eng. C-Mater. Biol. Appl. 2019, 104, 18. [Google Scholar] [CrossRef]
  93. Picheth, G.F.; Pirich, C.L.; Sierakowski, M.R.; Woehl, M.A.; Sakakibara, C.N.; de Souza, C.F.; Martin, A.A.; da Silva, R.; de Freitas, R.A. Bacterial cellulose in biomedical applications: A review. Int. J. Biol. Macromol. 2017, 104, 97–106. [Google Scholar] [CrossRef] [PubMed]
  94. Abeer, M.M.; Amin, M.; Martin, C. A review of bacterial cellulose-based drug delivery systems: Their biochemistry, current approaches and future prospects. J. Pharm. Pharmacol. 2014, 66, 1047–1061. [Google Scholar] [CrossRef]
  95. Avrămescu, R.E.; Ghica, M.V.; Dinu-Pîrvu, C.; Udeanu, D.; Popa, L. Liquid Marbles: From Industrial to Medical Applications. Molecules 2018, 23, 1120. [Google Scholar] [CrossRef] [Green Version]
  96. Jia, Y.; Zheng, M.; Xu, Q.; Zhong, C. Rheological behaviors of Pickering emulsions stabilized by TEMPO-oxidized bacterial cellulose. Carbohydr. Polym. 2019, 215, 263–271. [Google Scholar] [CrossRef]
  97. Florea, M.; Hagemann, H.; Santosa, G.; Abbott, J.; Micklem, C.N.; Spencer-Milnes, X.; Garcia, L.D.; Paschou, D.; Lazenbatt, C.; Kong, D.Z.; et al. Engineering control of bacterial cellulose production using a genetic toolkit and a new cellulose-producing strain. Proc. Natl. Acad. Sci. USA 2016, 113, E3431–E3440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Trigona, C.; Graziani, S.; Di Pasquale, G.; Pollicino, A.; Nisi, R.; Licciulli, A. Green Energy Harvester from Vibrations Based on Bacterial Cellulose. Sensors 2020, 20, 136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Silviana, S.; Susanti, S. Bacterial Cellulose Based Biocomposite from Guava Fruit Reinforced with Bamboo Microfibrillated Cellulose Through Impregnation Method. Orient. J. Chem. 2019, 35, 1029–1036. [Google Scholar] [CrossRef] [Green Version]
  100. Lay, M.; Gonzalez, I.; Tarres, J.A.; Pellicer, N.; Bun, K.N.; Vilaseca, F. High electrical and electrochemical properties in bacterial cellulose/polypyrrole membranes. Eur. Polym. J. 2017, 91, 1–9. [Google Scholar] [CrossRef]
  101. Ullah, H.; Wahid, F.; Santos, H.A.; Khan, T. Advances in biomedical and pharmaceutical applications of functional bacterial cellulose-based nanocomposites. Carbohydr. Polym. 2016, 150, 330–352. [Google Scholar] [CrossRef]
  102. Shah, N.; Ul-Islam, M.; Khattak, W.A.; Park, J.K. Overview of bacterial cellulose composites: A multipurpose advanced material. Carbohydr. Polym. 2013, 98, 1585–1598. [Google Scholar] [CrossRef]
  103. Halib, N.; Ahmad, I.; Grassi, M.; Grassi, G. The remarkable three-dimensional network structure of bacterial cellulose for tissue engineering applications. Int. J. Pharm. 2019, 566, 631–640. [Google Scholar] [CrossRef]
  104. Awadhiya, A.; Kumar, D.; Rathore, K.; Fatma, B.; Verma, V. Synthesis and characterization of agarose-bacterial cellulose biodegradable composites. Polym. Bull. 2017, 74, 2887–2903. [Google Scholar] [CrossRef]
  105. Hu, W.L.; Chen, S.Y.; Yang, J.X.; Li, Z.; Wang, H.P. Functionalized bacterial cellulose derivatives and nanocomposites. Carbohydr. Polym. 2014, 101, 1043–1060. [Google Scholar] [CrossRef]
  106. White, J.R. Polymer ageing: Physics, chemistry or engineering? Time to reflect. C. R. Chim. 2006, 9, 1396–1408. [Google Scholar] [CrossRef]
  107. Park, J.; Kim, K.; Lee, T.; Kim, M. Tailings Storage Facilities (TSFs) Dust Control Using Biocompatible Polymers. Mining Metall. Explor. 2019, 36, 785–795. [Google Scholar] [CrossRef]
  108. Rujnic-Sokele, M.; Pilipovic, A. Challenges and opportunities of biodegradable plastics: A mini review. Waste Manag. Res. 2017, 35, 132–140. [Google Scholar] [CrossRef]
  109. Qin, M.; Chen, C.Y.; Song, B.; Shen, M.C.; Cao, W.C.; Yang, H.L.; Zeng, G.M.; Gong, J.L. A review of biodegradable plastics to biodegradable microplastics: Another ecological threat to soil environments? J. Clean Prod. 2021, 312, 15. [Google Scholar] [CrossRef]
  110. Vroman, I.; Tighzert, L. Biodegradable Polymers. Materials 2009, 2, 307–344. [Google Scholar] [CrossRef]
  111. Sid, S.; Mor, R.S.; Kishore, A.; Sharanagat, V.S. Bio-sourced polymers as alternatives to conventional food packaging materials: A review. Trends Food Sci. Technol. 2021, 115, 87–104. [Google Scholar] [CrossRef]
  112. Lehn, J.M. Supramolecular polymer chemistry-scope and perspectives. Polym. Int. 2002, 51, 825–839. [Google Scholar] [CrossRef]
  113. Bhatia, S. Natural Polymers vs. Synthetic Polymer. In Natural Polymer Drug Delivery Systems; Springer: Cham, Switzerland, 2016; pp. 95–118. [Google Scholar]
  114. Tudoroiu, E.-E.; Dinu-Pîrvu, C.-E.; Albu Kaya, M.G.; Popa, L.; Anuța, V.; Prisada, R.M.; Ghica, M.V. An Overview of Cellulose Derivatives-Based Dressings for Wound-Healing Management. Pharmaceuticals 2021, 14, 1215. [Google Scholar] [CrossRef] [PubMed]
  115. Maitz, M.F. Applications of synthetic polymers in clinical medicine. Biosurface Biotribol. 2015, 1, 161–176. [Google Scholar] [CrossRef] [Green Version]
  116. Tian, H.Y.; Tang, Z.H.; Zhuang, X.L.; Chen, X.S.; Jing, X.B. Biodegradable synthetic polymers: Preparation, functionalization and biomedical application. Prog. Polym. Sci. 2012, 37, 237–280. [Google Scholar] [CrossRef]
  117. Rebelo, R.; Fernandes, M.; Fangueiro, R. Biopolymers in Medical Implants: A Brief Review. In Proceedings of the 3rd International Conference on Natural Fibers–Advanced Materials for a Greener World (ICNF), Braga, Portugal, 21–23 June 2017; pp. 236–243. [Google Scholar]
  118. Rajwade, J.M.; Paknikar, K.M.; Kumbhar, J.V. Applications of bacterial cellulose and its composites in biomedicine. Appl. Microbiol. Biotechnol. 2015, 99, 2491–2511. [Google Scholar] [CrossRef]
  119. Yu, J.; Wang, C.P.; Wang, J.F.; Chu, F.X. Synthesis and Characterization of Ethyl Cellulose Based Acrylate. In Proceedings of the 4th International Conference on Manufacturing Science and Engineering (ICMSE 2013), Dalian, China, 30–31 March 2013; pp. 124–130. [Google Scholar]
  120. Petrauskaite, O.; Juodzbalys, G.; Viskelis, P.; Liesiene, J. Control of the Porous Structure of Cellulose-Based Tissue Engineering Scaffolds by Means of Lyophilization. Cell Chem. Technol. 2016, 50, 23–30. [Google Scholar]
  121. Mello, L.; Feltrin, Y.; Selbach, R.; Macedo, G.; Spautz, C.; Haas, L. Use of lyophilized cellulose in peripheral nerve lesions with loss of substance. Arq. De Neuro-Psiquiatr. 2001, 59, 372–379. [Google Scholar] [CrossRef] [Green Version]
  122. Menezes-Silva, R.; de Oliueira, B.M.B.; Fernandes, P.H.M.; Shimohara, L.Y.; Pereira, F.V.; Borges, A.F.S.; Buzalaf, M.A.R.; Pascotto, R.C.; Sidhu, S.K.; Navarro, M.F.D. Effects of the reinforced cellulose nanocrystals on glass-ionomer cements. Dent. Mater. 2019, 35, 564–573. [Google Scholar] [CrossRef] [PubMed]
  123. Bazghaleh, A.A.; Dogolsar, M.A. Preparation of Degradable Oxidized Regenerated Cellulose Gauze by Zinc Modification on HNO3/Cu Oxidized Viscose Fibers. Fiber. Polym. 2019, 20, 1125–1135. [Google Scholar] [CrossRef]
  124. Chowdhry, S.A. Use of oxidized regenerated cellulose (ORC)/collagen/silver-ORC dressings to help manage skin graft donor site wounds. JPRAS Open 2019, 22, 33–40. [Google Scholar] [CrossRef] [PubMed]
  125. Cornelia Nitipir, S.M.; Marin, M.M.; Kaya, M.A.; Ghica, M.V.; Mederle, N. Hybrid Collagen-NaCMC Matrices Loaded with Mefenamic Acid for Wound Healing. Rev. Chim. 2017, 68, 2605–2609. [Google Scholar] [CrossRef]
  126. Ghica, M.V.; Albu Kaya, M.G.; Dinu-Pirvu, C.E.; Lupuleasa, D.; Udeanu, D.I. Development, Optimization and In Vitro/In Vivo Characterization of Collagen-Dextran Spongious Wound Dressings Loaded with Flufenamic Acid. Molecules 2017, 22, 1552. [Google Scholar] [CrossRef] [PubMed]
  127. Albu, M.G.; Vuluga, Z.; Panaitescu, D.M.; Vuluga, D.M.; Căşărică, A.; Ghiurea, M. Morphology and thermal stability of bacterial cellulose/collagen composites. Cent. Eur. J. Chem. 2014, 12, 968–975. [Google Scholar] [CrossRef]
  128. Păunica-Panea, G.; Ficai, A.; Marin, M.M.; Marin, Ș.; Albu, M.G.; Constantin, V.D.; Dinu-Pîrvu, C.; Vuluga, Z.; Corobea, M.C.; Ghica, M.V. New Collagen-Dextran-Zinc Oxide Composites for Wound Dressing. J. Nanomater. 2016, 2016, 5805034. [Google Scholar] [CrossRef] [Green Version]
  129. Saska, S.; Teixeira, L.; Oliveira, P.; Marchetto, R.; Gaspar, A.; Ribeiro, S.; Messaddeq, Y. Bacterial cellulose-collagen nanocomposite for bone tissue engineering. J. Mater. Chem. 2012, 22, 22102–22112. [Google Scholar] [CrossRef]
  130. Noh, Y.K.; Dos Santos Da Costa, A.; Park, Y.S.; Du, P.; Kim, I.H.; Park, K. Fabrication of bacterial cellulose-collagen composite scaffolds and their osteogenic effect on human mesenchymal stem cells. Carbohydr. Polym. 2019, 219, 210–218. [Google Scholar] [CrossRef] [PubMed]
  131. Zaborowska, M.; Bodin, A.; Backdahl, H.; Popp, J.; Goldstein, A.; Gatenholm, P. Microporous bacterial cellulose as a potential scaffold for bone regeneration. Acta Biomater. 2010, 6, 2540–2547. [Google Scholar] [CrossRef] [PubMed]
  132. Saska, S.; Pigossi, S.C.; Oliveira, G.; Teixeira, L.N.; Capela, M.V.; Goncalves, A.; de Oliveira, P.T.; Messaddeq, Y.; Ribeiro, S.J.L.; Gaspar, A.M.M.; et al. Biopolymer-based membranes associated with osteogenic growth peptide for guided bone regeneration. Biomed. Mater. 2018, 13, 035009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Moraes, P.; Saska, S.; Barud, H.; de Lima, L.R.; Martins, V.D.A.; Plepis, A.M.D.; Ribeiro, S.J.L.; Gaspar, A.M.M. Bacterial Cellulose/Collagen Hydrogel for Wound Healing. Mater. Res.-Ibero-am. J. Mater. 2016, 19, 106–116. [Google Scholar] [CrossRef] [Green Version]
  134. Zhang, W.; Wang, X.C.; Wang, J.J.; Zhang, L.L. Drugs adsorption and release behavior of collagen/bacterial cellulose porous microspheres. Int. J. Biol. Macromol. 2019, 140, 196–205. [Google Scholar] [CrossRef]
  135. Jia, Y.Y.; Wang, X.H.; Huo, M.M.; Zhai, X.L.; Li, F.; Zhong, C. Preparation and characterization of a novel bacterial cellulose/chitosan bio-hydrogel. Nanomater. Nanotechnol. 2017, 7, 8. [Google Scholar] [CrossRef]
  136. Liu, X.; Wang, Y.; Cheng, Z.; Sheng, J.; Yang, R.D. Nano-sized fibrils dispersed from bacterial cellulose grafted with chitosan. Carbohydr. Polym. 2019, 214, 311–316. [Google Scholar] [CrossRef]
  137. Ren, J.M.; Li, Q.; Dong, F.; Feng, Y.; Guo, Z.Y. Phenolic antioxidants-functionalized quaternized chitosan: Synthesis and antioxidant properties. Int. J. Biol. Macromol. 2013, 53, 77–81. [Google Scholar] [CrossRef]
  138. Ai, H.; Wang, F.R.; Xia, Y.Q.; Chen, X.M.; Lei, C.L. Antioxidant, antifungal and antiviral activities of chitosan from the larvae of housefly, Musca domestica L. Food Chem. 2012, 132, 493–498. [Google Scholar] [CrossRef]
  139. Song, Y.H.; Nagai, N.; Saijo, S.; Kaji, H.; Nishizawa, M.; Abe, T. In situ formation of injectable chitosan-gelatin hydrogels through double crosslinking for sustained intraocular drug delivery. Mater. Sci. Eng. C-Mater. Biol. Appl. 2018, 88, 1–12. [Google Scholar] [CrossRef]
  140. Li, R.W.; Liu, Q.Y.; Wu, H.W.; Wang, K.; Li, L.H.; Zhou, C.R.; Ao, N.J. Preparation and characterization of in-situ formable liposome/chitosan composite hydrogels. Mater. Lett. 2018, 220, 289–292. [Google Scholar] [CrossRef]
  141. Starychova, L.; Zabka, M.; Spaglova, M.; Cuchorova, M.; Vitkova, M.; Cierna, M.; Bartonikova, K.; Gardavska, K. In Vitro Liberation of Indomethacin from Chitosan Gels Containing Microemulsion in Different Dissolution Mediums. J. Pharm. Sci. 2014, 103, 3977–3984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Parteni, O.; Radu, C.D.; Muresan, A.; Popa, M.; Ochiuz, L.; Sandu, A.V.; Agafitei, G.; Istrate, B.; Munteanu, C. Improving the Obtaining Factors of a Chitosan Hydrogel Based Biomaterial. Rev. Chim. 2015, 66, 1595–1599. [Google Scholar]
  143. Alnufaiy, B.M.; Lambarte, R.N.A.; Al-Hamdan, K.S. The Osteogenetic Potential of Chitosan Coated Implant: An In Vitro Study. J. Stem Cells Regen. Med. 2020, 16, 44–49. [Google Scholar] [CrossRef] [PubMed]
  144. Lăcrămioara Popa, M.V.G.; Dinu-Pîrvu, C.E. Periodontal chitosan-gels designed for improved local intra-pocket drug delivery. Farmacia 2013, 61, 240–250. [Google Scholar]
  145. Irimia, T.; Dinu-Pirvu, C.E.; Ghica, M.V.; Lupuleasa, D.; Muntean, D.L.; Udeanu, D.I.; Popa, L. Chitosan-Based In Situ Gels for Ocular Delivery of Therapeutics: A State-of-the-Art Review. Mar. Drugs 2018, 16, 373. [Google Scholar] [CrossRef] [Green Version]
  146. Irimia, T.; Ghica, M.V.; Popa, L.; Anuta, V.; Arsene, A.L.; Dinu-Pirvu, C.E. Strategies for Improving Ocular Drug Bioavailability and Corneal Wound Healing with Chitosan-Based Delivery Systems. Polymers 2018, 10, 1221. [Google Scholar] [CrossRef] [Green Version]
  147. Lăcrămioara Popa, M.V.G.; Dinu-Pîrvu, C.E.; Irimia, T. Chitosan: A Good Candidate for Sustained Release Ocular Drug Delivery Systems; IntechOpen: London, UK, 2018. [Google Scholar]
  148. Lin, W.C.; Lien, C.C.; Yeh, H.J.; Yu, C.M.; Hsu, S.H. Bacterial cellulose and bacterial cellulose-chitosan membranes for wound dressing applications. Carbohydr Polym 2013, 94, 603–611. [Google Scholar] [CrossRef]
  149. Cacicedo, M.L.; Pacheco, G.; Islan, G.A.; Alvarez, V.A.; Barud, H.S.; Castro, G.R. Chitosan-bacterial cellulose patch of ciprofloxacin for wound dressing: Preparation and characterization studies. Int. J. Biol. Macromol. 2019, 147, 1136–1145. [Google Scholar] [CrossRef]
  150. Indriyati; Dara, F.; Primadona, I.; Srikandace, Y.; Karina, M. Development of bacterial cellulose/chitosan films: Structural, physicochemical and antimicrobial properties. J. Polym. Res. 2021, 28, 8. [Google Scholar] [CrossRef]
  151. Kamel, S.; Khattab, T.A. Recent Advances in Cellulose-Based Biosensors for Medical Diagnosis. Biosensors 2020, 10, 67. [Google Scholar] [CrossRef] [PubMed]
  152. Liu, Y.W.; Ahmed, S.; Sameen, D.E.; Wang, Y.; Lu, R.; Dai, J.W.; Li, S.Q.; Qin, W. A review of cellulose and its derivatives in biopolymer-based for food packaging application. Trends Food Sci. Technol. 2021, 112, 532–546. [Google Scholar] [CrossRef]
  153. Correa, M.J.; Añón, M.C.; Pérez, G.T.; Ferrero, C. Effect of modified celluloses on dough rheology and microstructure. Food Res. Int. 2010, 43, 780–787. [Google Scholar] [CrossRef]
  154. Zhao, Q.; Zhao, M.; Li, J.; Yang, B.; Su, G.; Cui, C.; Jiang, Y. Effect of hydroxypropyl methylcellulose on the textural and whipping properties of whipped cream. Food Hydrocoll. 2009, 23, 2168–2173. [Google Scholar] [CrossRef]
  155. Suppakul, P.; Jutakorn, K.; Bangchokedee, Y. Efficacy of cellulose-based coating on enhancing the shelf life of fresh eggs. J. Food Eng. 2010, 98, 207–213. [Google Scholar] [CrossRef]
  156. Pérez, C.D.; Descalzo, A.M.; Rojas, A.M.; Gerschenson, L.N.; De’Nobili, M.D.; Rizzo, S.A. High methoxyl pectin–methyl cellulose films with antioxidant activity at a functional food interface. J. Food Eng. 2013, 116, 162–169. [Google Scholar] [CrossRef]
  157. Mallakpour, S.; Tukhani, M.; Hussain, C.M. Recent advancements in 3D bioprinting technology of carboxymethyl cellulose-based hydrogels: Utilization in tissue engineering. Adv. Colloid Interface Sci. 2021, 292, 102415. [Google Scholar] [CrossRef]
  158. Udeanu, D.I. Anti-Inflammatory Drug-Loaded Biopolymeric Spongious Matrices with Therapeutic Perspectives in Burns Treatment. Farmacia 2018, 66, 783–790. [Google Scholar] [CrossRef]
  159. Dinescu, S.; Ignat, S.R.; Lazar, A.D.; Marin, S.; Danila, E.; Marin, M.M.; Udeanu, D.I.; Ghica, M.V.; Albu-Kaya, M.G.; Costache, M. Efficiency of Multiparticulate Delivery Systems Loaded with Flufenamic Acid Designed for Burn Wound Healing Applications. J. Immunol. Res. 2019, 2019, 4513108. [Google Scholar] [CrossRef] [Green Version]
  160. Pavaloiu, R.D.; Stroescu, M.; Parvulescu, O.; Dobre, T. Composite Hydrogels of Bacterial Cellulose–carboxymethyl Cellulose for Drug Release. Rev. Chim. 2014, 65, 948–951. [Google Scholar]
  161. Juncu, G.; Stoica-Guzun, A.; Stroescu, M.; Isopencu, G.; Jinga, S.I. Drug release kinetics from carboxymethylcellulose-bacterial cellulose composite films. Int. J. Pharm. 2016, 510, 485–492. [Google Scholar] [CrossRef] [PubMed]
  162. Martinez Avila, H.; Feldmann, E.M.; Pleumeekers, M.M.; Nimeskern, L.; Kuo, W.; de Jong, W.C.; Schwarz, S.; Muller, R.; Hendriks, J.; Rotter, N.; et al. Novel bilayer bacterial nanocellulose scaffold supports neocartilage formation in vitro and in vivo. Biomaterials 2015, 44, 122–133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Chiaoprakobkij, N.; Sanchavanakit, N.; Subbalekha, K.; Pavasant, P.; Phisalaphong, M. Characterization and biocompatibility of bacterial cellulose/alginate composite sponges with human keratinocytes and gingival fibroblasts. Carbohydr. Polym. 2011, 85, 548–553. [Google Scholar] [CrossRef]
  164. Wichai, S.; Chuysinuan, P.; Chaiarwut, S.; Ekabutr, P.; Supaphol, P. Development of bacterial cellulose/alginate/chitosan composites incorporating copper (II) sulfate as an antibacterial wound dressing. J. Drug Deliv. Sci. Technol. 2019, 51, 662–671. [Google Scholar] [CrossRef]
  165. Shao, W.; Liu, H.; Liu, X.; Wang, S.; Wu, J.; Zhang, R.; Min, H.; Huang, M. Development of silver sulfadiazine loaded bacterial cellulose/sodium alginate composite films with enhanced antibacterial property. Carbohydr. Polym. 2015, 132, 351–358. [Google Scholar] [CrossRef]
  166. Kim, J.H.; Park, S.; Kim, H.; Kim, H.J.; Yang, Y.H.; Kim, Y.H.; Jung, S.K.; Kan, E.; Lee, S.H. Alginate/bacterial cellulose nanocomposite beads prepared using Gluconacetobacter xylinus and their application in lipase immobilization. Carbohydr. Polym. 2017, 157, 137–145. [Google Scholar] [CrossRef]
  167. Tummala, G.K.; Lopes, V.R.; Mihranyan, A.; Ferraz, N. Biocompatibility of Nanocellulose-Reinforced PVA Hydrogel with Human Corneal Epithelial Cells for Ophthalmic Applications. J. Func. Biomater. 2019, 10, 35. [Google Scholar] [CrossRef] [Green Version]
  168. Gade, S.K.; Shivshetty, N.; Sharma, N.; Bhatnagar, S.; Garg, P.; Venuganti, V.V.K. Effect of Mucoadhesive Polymeric Formulation on Corneal Permeation of Fluoroquinolones. J. Ocular Pharmacol. Ther. 2018, 34, 570–578. [Google Scholar] [CrossRef]
  169. Kodym, A.; Bilski, P.; Domanska, A.; Helminiak, L.; Jablonska, M.; Jachymska, A. Physical and Chemical Properties and Stability of Sodium Cefazolin in Buffered Eye Drops Determined with Hplc Method. Acta Pol. Pharm. 2012, 69, 95–105. [Google Scholar]
  170. Jain, D.; Carvalho, E.; Banthia, A.K.; Banerjee, R. Development of polyvinyl alcohol-gelatin membranes for antibiotic delivery in the eye. Drug Dev. Ind. Pharm. 2011, 37, 167–177. [Google Scholar] [CrossRef]
  171. Ran, W.Y.; Ma, H.D.; Li, M. In Vitro and In Vivo Studies of Polyvinyl Pyrrolidone-Coated Sparfloxacin-Loaded Ring Contact Lens to Treat Conjunctivitis. J. Pharm. Sci. 2020, 109, 1951–1957. [Google Scholar] [CrossRef] [PubMed]
  172. Marin, S.; Albu Kaya, M.G.; Ghica, M.V.; Dinu-Pirvu, C.; Popa, L.; Udeanu, D.I.; Mihai, G.; Enachescu, M. Collagen-Polyvinyl Alcohol-Indomethacin Biohybrid Matrices as Wound Dressings. Pharmaceutics 2018, 10, 224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Pavaloiu, R.D.; Stoica-Guzun, A.; Stroescu, M.; Jinga, S.I.; Dobre, T. Composite films of poly(vinyl alcohol)-chitosan-bacterial cellulose for drug controlled release. Int. J. Biol. Macromol. 2014, 68, 117–124. [Google Scholar] [CrossRef] [PubMed]
  174. Usawattanakul, N.; Torgbo, S.; Sukyai, P.; Khantayanuwong, S.; Puangsin, B.; Srichola, P. Development of Nanocomposite Film Comprising of Polyvinyl Alcohol (PVA) Incorporated with Bacterial Cellulose Nanocrystals and Magnetite Nanoparticles. Polymers 2021, 13, 1778. [Google Scholar] [CrossRef] [PubMed]
  175. Abdelraof, M.; Hasanin, M.S.; Farag, M.M.; Ahmed, H.Y. Green synthesis of bacterial cellulose/bioactive glass nanocomposites: Effect of glass nanoparticles on cellulose yield, biocompatibility and antimicrobial activity. Int. J. Biol. Macromol. 2019, 138, 975–985. [Google Scholar] [CrossRef]
  176. Chen, J.; Chen, C.; Liang, G.; Xu, X.; Hao, Q.; Sun, D. In situ preparation of bacterial cellulose with antimicrobial properties from bioconversion of mulberry leaves. Carbohydr. Polym. 2019, 220, 170–175. [Google Scholar] [CrossRef]
  177. Jiji, S.; Udhayakumar, S.; Rose, C.; Muralidharan, C.; Kadirvelu, K. Thymol enriched bacterial cellulose hydrogel as effective material for third degree burn wound repair. Int. J. Biol. Macromol. 2019, 122, 452–460. [Google Scholar] [CrossRef] [PubMed]
  178. Savitskaya, I.S.; Shokatayeva, D.H.; Kistaubayeva, A.S.; Ignatova, L.V.; Digel, I.E. Antimicrobial and wound healing properties of a bacterial cellulose based material containing B. subtilis cells. Heliyon 2019, 5, e02592. [Google Scholar] [CrossRef] [Green Version]
  179. Bodin, A.; Bharadwaj, S.; Wu, S.F.; Gatenholm, P.; Atala, A.; Zhang, Y.Y. Tissue-engineered conduit using urine-derived stem cells seeded bacterial cellulose polymer in urinary reconstruction and diversion. Biomaterials 2010, 31, 8889–8901. [Google Scholar] [CrossRef]
  180. Cao, Y.-M.; Liu, M.-Y.; Xue, Z.-W.; Qiu, Y.; Li, J.; Wang, Y.; Wu, Q.-K. Surface-structured bacterial cellulose loaded with hUSCs accelerate skin wound healing by promoting angiogenesis in rats. Biochem. Biophys. Res. Commun. 2019, 516, 1167–1174. [Google Scholar] [CrossRef]
  181. Petersen, N.; Gatenholm, P. Bacterial cellulose-based materials and medical devices: Current state and perspectives. Appl. Microbiol. Biotechnol. 2011, 91, 1277–1286. [Google Scholar] [CrossRef]
  182. Gorgieva, S.; Trcek, J. Bacterial Cellulose: Production, Modification and Perspectives in Biomedical Applications. Nanomaterials 2019, 9, 1352. [Google Scholar] [CrossRef] [Green Version]
  183. Czaja, W.K.; Young, D.J.; Kawecki, M.; Brown, R.M., Jr. The future prospects of microbial cellulose in biomedical applications. Biomacromolecules 2007, 8, 1–12. [Google Scholar] [CrossRef] [PubMed]
  184. Fu, L.; Zhou, P.; Zhang, S.; Yang, G. Evaluation of bacterial nanocellulose-based uniform wound dressing for large area skin transplantation. Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 2995–3000. [Google Scholar] [CrossRef] [PubMed]
  185. Czaja, W.; Krystynowicz, A.; Kawecki, M.; Wysota, K.; Sakiel, S.; Wróblewski, P.; Glik, J.; Nowak, M.; Bielecki, S. Biomedical Applications of Microbial Cellulose in Burn Wound Recovery. In Cellulose: Molecular and Structural Biology: Selected Articles on the Synthesis, Structure, and Applications of Cellulose; Brown, R.M., Saxena, I.M., Eds.; Springer: Dordrecht, The Netherlands, 2007; pp. 307–321. [Google Scholar]
  186. Moniri, M.; Boroumand Moghaddam, A.; Azizi, S.; Abdul Rahim, R.; Bin Ariff, A.; Zuhainis Saad, W.; Navaderi, M.; Mohamad, R. Production and Status of Bacterial Cellulose in Biomedical Engineering. Nanomaterials 2017, 7, 257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Fu, L.; Zhang, Y.; Li, C.; Wu, Z.; Zhuo, Q.; Huang, X.; Qiu, G.; Zhou, P.; Yang, G. Skin tissue repair materials from bacterial cellulose by a multilayer fermentation method. J. Mater. Chem. 2012, 22, 12349–12357. [Google Scholar] [CrossRef]
  188. Park, S.U.; Lee, B.K.; Kim, M.S.; Park, K.K.; Sung, W.J.; Kim, H.Y.; Han, D.G.; Shim, J.S.; Lee, Y.J.; Kim, S.H.; et al. The possibility of microbial cellulose for dressing and scaffold materials. Int. Wound J. 2014, 11, 35–43. [Google Scholar] [CrossRef]
  189. Cho, K.-H.; Park, J.-E.; Osaka, T.; Park, S.-G. The study of antimicrobial activity and preservative effects of nanosilver ingredient. Electrochim. Acta 2005, 51, 956–960. [Google Scholar] [CrossRef]
  190. Sondi, I.; Salopek-Sondi, B. Silver nanoparticles as antimicrobial agent: A case study on E. coli as a model for Gram-negative bacteria. J. Colloid Interface Sci. 2004, 275, 177–182. [Google Scholar] [CrossRef]
  191. Percival, S.L.; Bowler, P.G.; Russell, D. Bacterial resistance to silver in wound care. J. Hosp. Infect. 2005, 60, 1–7. [Google Scholar] [CrossRef]
  192. Napavichayanun, S.; Yamdech, R.; Aramwit, P. The safety and efficacy of bacterial nanocellulose wound dressing incorporating sericin and polyhexamethylene biguanide: In vitro, in vivo and clinical studies. Arch. Dermatol. Res. 2016, 308, 123–132. [Google Scholar] [CrossRef] [PubMed]
  193. Millon, L.E.; Oates, C.J.; Wan, W. Compression properties of polyvinyl alcohol-bacterial cellulose nanocomposite. J. Biomed. Mater. Res. Part B Appl. Biomater. 2009, 90, 922–929. [Google Scholar] [CrossRef] [PubMed]
  194. Dugan, J.M.; Gough, J.E.; Eichhorn, S.J. Bacterial cellulose scaffolds and cellulose nanowhiskers for tissue engineering. Nanomedicine 2013, 8, 287–298. [Google Scholar] [CrossRef] [PubMed]
  195. Kneser, U.; Schaefer, D.J.; Munder, B.; Klemt, C.; Andree, C.; Stark, G.B. Tissue engineering of bone. Minim. Invasive Ther. Allied Technol. 2002, 11, 107–116. [Google Scholar] [CrossRef]
  196. Kalia, S.; Dufresne, A.; Cherian, B.M.; Kaith, B.S.; Avérous, L.; Njuguna, J.; Nassiopoulos, E. Cellulose-Based Bio- and Nanocomposites: A Review. Int. J. Polym. Sci. 2011, 2011, 1–35. [Google Scholar] [CrossRef]
  197. Maia, F.R.; Bastos, A.R.; Oliveira, J.M.; Correlo, V.M.; Reis, R.L. Recent approaches towards bone tissue engineering. Bone 2022, 154, 116256. [Google Scholar] [CrossRef]
  198. Torgbo, S.; Sukyai, P. Bacterial cellulose-based scaffold materials for bone tissue engineering. Appl. Mater. Today 2018, 11, 34–49. [Google Scholar] [CrossRef]
  199. Zimmermann, K.A.; LeBlanc, J.M.; Sheets, K.T.; Fox, R.W.; Gatenholm, P. Biomimetic design of a bacterial cellulose/hydroxyapatite nanocomposite for bone healing applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2011, 31, 43–49. [Google Scholar] [CrossRef]
  200. Torgbo, S.; Sukyai, P. Fabrication of microporous bacterial cellulose embedded with magnetite and hydroxyapatite nanocomposite scaffold for bone tissue engineering. Mater. Chem. Phys. 2019, 237, 121868. [Google Scholar] [CrossRef]
  201. Kumbar, S.G.; Toti, U.S.; Deng, M.; James, R.; Laurencin, C.T.; Aravamudhan, A.; Harmon, M.; Ramos, D.M. Novel mechanically competent polysaccharide scaffolds for bone tissue engineering. Biomed. Mater. 2011, 6, 065005. [Google Scholar] [CrossRef]
  202. Aravamudhan, A.; Ramos, D.M.; Nip, J.; Harmon, M.D.; James, R.; Deng, M.; Laurencin, C.T.; Yu, X.; Kumbar, S.G. Cellulose and collagen derived micro-nano structured scaffolds for bone tissue engineering. J. Biomed. Nanotechnol. 2013, 9, 719–731. [Google Scholar] [CrossRef] [PubMed]
  203. Avrămescu, R.E.; Ghica, M.V.; Dinu-Pirvu, C.; Prisada, R.; Popa, L. Superhydrophobic Natural and Artificial Surfaces-A Structural Approach. Materials 2018, 11, 866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Dinu-Pîrvu, C.E.; Avrămescu, R.E.; Ghica, M.V.; Popa, L. Natural and Artificial Superwettable Surface-Superficial Phenomena: An Extreme Wettability Scenario. In Wettability and Interfacial Phenomena–Implications for Material Processing; IntechOpen: London, UK, 2019. [Google Scholar]
  205. Sun, B.; Wei, F.; Li, W.; Xu, X.; Zhang, H.; Liu, M.; Lin, J.; Ma, B.; Chen, C.; Sun, D. Macroporous bacterial cellulose grafted by oligopeptides induces biomimetic mineralization via interfacial wettability. Colloids Surf. B Biointerfaces 2019, 183, 110457. [Google Scholar] [CrossRef] [PubMed]
  206. Zhu, W.; Li, W.; He, Y.; Duan, T. In-situ biopreparation of biocompatible bacterial cellulose/graphene oxide composites pellets. Appl. Surf. Sci. 2015, 338, 22–26. [Google Scholar] [CrossRef]
  207. Schumann, D.A.; Wippermann, J.; Klemm, D.O.; Kramer, F.; Koth, D.; Kosmehl, H.; Wahlers, T.; Salehi-Gelani, S. Artificial vascular implants from bacterial cellulose: Preliminary results of small arterial substitutes. Cellulose 2009, 16, 877–885. [Google Scholar] [CrossRef]
  208. Esguerra, M.; Fink, H.; Laschke, M.W.; Jeppsson, A.; Delbro, D.; Gatenholm, P.; Menger, M.D.; Risberg, B. Intravital fluorescent microscopic evaluation of bacterial cellulose as scaffold for vascular grafts. J. Biomed. Mater. Res. Part A 2010, 93, 140–149. [Google Scholar] [CrossRef] [PubMed]
  209. Picolotto, A.; Pergher, D.; Pereira, G.P.; Machado, K.G.; da Silva Barud, H.; Roesch-Ely, M.; Gonzalez, M.H.; Tasso, L.; Figueiredo, J.G.; Moura, S. Bacterial cellulose membrane associated with red propolis as phytomodulator: Improved healing effects in experimental models of diabetes mellitus. Biomed. Pharmacother. 2019, 112, 108640. [Google Scholar] [CrossRef]
  210. Mohamad, N.; Mohd Amin, M.C.I.; Pandey, M.; Ahmad, N.; Rajab, N. Bacterial cellulose/acrylic acid hydrogel synthesized via electron beam irradiation: Accelerated burn wound healing in an animal model. Carbohydr. Polym. 2014, 114, 312–320. [Google Scholar] [CrossRef] [PubMed]
  211. Bottan, S.; Robotti, F.; Jayathissa, P.; Hegglin, A.; Bahamonde, N.; Heredia-Guerrero, J.A.; Bayer, I.S.; Scarpellini, A.; Merker, H.; Lindenblatt, N.; et al. Surface-structured bacterial cellulose with guided assembly-based biolithography (GAB). ACS Nano 2015, 9, 206–219. [Google Scholar] [CrossRef]
  212. Phutanon, N.; Motina, K.; Chang, Y.H.; Ummartyotin, S. Development of CuO particles onto bacterial cellulose sheets by forced hydrolysis: A synergistic approach for generating sheets with photocatalytic and antibiofouling properties. Int. J. Biol. Macromol. 2019, 136, 1142–1152. [Google Scholar] [CrossRef] [PubMed]
  213. Wu, C.N.; Fuh, S.C.; Lin, S.P.; Lin, Y.Y.; Chen, H.Y.; Liu, J.M.; Cheng, K.C. TEMPO-Oxidized Bacterial Cellulose Pellicle with Silver Nanoparticles for Wound Dressing. Biomacromolecules 2018, 19, 544–554. [Google Scholar] [CrossRef] [PubMed]
  214. Khalid, A.; Ullah, H.; Ul-Islam, M.; Khan, R.; Khan, S.; Ahmad, F.; Khan, T.; Wahid, F. Bacterial cellulose–TiO2 nanocomposites promote healing and tissue regeneration in burn mice model. RSC Adv. 2017, 7, 47662–47668. [Google Scholar] [CrossRef] [Green Version]
  215. Yang, G.; Xie, J.; Deng, Y.; Bian, Y.; Hong, F. Hydrothermal synthesis of bacterial cellulose/AgNPs composite: A “green” route for antibacterial application. Carbohydr. Polym. 2012, 87, 2482–2487. [Google Scholar] [CrossRef]
  216. Maneerung, T.; Tokura, S.; Rujiravanit, R. Impregnation of silver nanoparticles into bacterial cellulose for antimicrobial wound dressing. Carbohydr. Polym. 2008, 72, 43–51. [Google Scholar] [CrossRef]
  217. Wu, J.; Zheng, Y.D.; Song, W.H.; Luan, J.B.; Wen, X.X.; Wu, Z.G.; Chen, X.H.; Wang, Q.; Guo, S.L. In situ synthesis of silver-nanoparticles/bacterial cellulose composites for slow-released antimicrobial wound dressing. Carbohydr. Polym. 2014, 102, 762–771. [Google Scholar] [CrossRef] [PubMed]
  218. Barud, H.S. Development and Evaluation of Biocure Obtained from Bacterial Cellulose and Standardized Extract of Propolis (EPP-AF) for the Treatment of Burns and/or Skin Lesions; São Paulo Research Foundation—FAPESP: São Paulo, Brazil, 2009. [Google Scholar]
  219. Morais, E.S.; Silva, N.; Sintra, T.E.; Santos, S.A.O.; Neves, B.M.; Almeida, I.F.; Costa, P.C.; Correia-Sa, I.; Ventura, S.P.M.; Silvestre, A.J.D.; et al. Anti-inflammatory and antioxidant nanostructured cellulose membranes loaded with phenolic-based ionic liquids for cutaneous application. Carbohydr. Polym. 2019, 206, 187–197. [Google Scholar] [CrossRef]
  220. Laçin, M. Development of biodegradable antibacterial cellulose based hydrogel membranes for wound healing. Int. J. Biol. Macromol. 2014, 67, 22–27. [Google Scholar] [CrossRef]
  221. Qiu, Y.; Qiu, L.; Cui, J.; Wei, Q. Bacterial cellulose and bacterial cellulose-vaccarin membranes for wound healing. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 59, 303–309. [Google Scholar] [CrossRef]
  222. Celes, F.S.; Trovatti, E.; Khouri, R.; Van Weyenbergh, J.; Ribeiro, S.J.L.; Borges, V.M.; Barud, H.S.; de Oliveira, C.I. DETC-based bacterial cellulose bio-curatives for topical treatment of cutaneous leishmaniasis. Sci. Rep. 2016, 6, 38330. [Google Scholar] [CrossRef] [Green Version]
  223. Fursatz, M.; Skog, M.; Sivler, P.; Palm, E.; Aronsson, C.; Skallberg, A.; Greczynski, G.; Khalaf, H.; Bengtsson, T.; Aili, D. Functionalization of bacterial cellulose wound dressings with the antimicrobial peptide epsilon-poly-L-Lysine. Biomed. Mater. 2018, 13, 025014. [Google Scholar] [CrossRef]
  224. Mohamad, N.; Loh, E.Y.X.; Fauzi, M.B.; Ng, M.H.; Mohd Amin, M.C.I. In vivo evaluation of bacterial cellulose/acrylic acid wound dressing hydrogel containing keratinocytes and fibroblasts for burn wounds. Drug Deliv. Transl. Res. 2019, 9, 444–452. [Google Scholar] [CrossRef] [PubMed]
  225. Altun, E.; Aydogdu, M.O.; Koc, F.; Crabbe-Mann, M.; Brako, F.; Kaur-Matharu, R.; Ozen, G.; Kuruca, S.E.; Edirisinghe, U.; Gunduz, O.; et al. Novel Making of Bacterial Cellulose Blended Polymeric Fiber Bandages. Macromol. Mater. Eng. 2018, 303, 1700607. [Google Scholar] [CrossRef] [Green Version]
  226. Alkhatib, Y.; Dewaldt, M.; Moritz, S.; Nitzsche, R.; Kralisch, D.; Fischer, D. Controlled extended octenidine release from a bacterial nanocellulose/Poloxamer hybrid system. Eur. J. Pharm. Biopharm. Off. J. Arb. Fur Pharm. Verfahr. 2017, 112, 164–176. [Google Scholar] [CrossRef] [PubMed]
  227. De Lima Fontes, M.; Meneguin, A.B.; Tercjak, A.; Gutierrez, J.; Cury, B.S.F.; Dos Santos, A.M.; Ribeiro, S.J.L.; Barud, H.S. Effect of in situ modification of bacterial cellulose with carboxymethylcellulose on its nano/microstructure and methotrexate release properties. Carbohydr. Polym. 2018, 179, 126–134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Wu, H.; Williams, G.R.; Wu, J.; Wu, J.; Niu, S.; Li, H.; Wang, H.; Zhu, L. Regenerated chitin fibers reinforced with bacterial cellulose nanocrystals as suture biomaterials. Carbohydr. Polym. 2018, 180, 304–313. [Google Scholar] [CrossRef] [PubMed]
  229. Vasconcellos, P.K.F.M.; Nóia, M.P.; De Castro, I.C.V.; dos Santos, J.N.; Pinheiro, A.L.B.; Marques, A.M.C.; Ramos, E.A.G.; Rocha, C.G. Influence of laser therapy on the dynamic formation of extracellular matrix in standard second degree burns treated with bacterial cellulose membrane. J. Photochem. Photobiol. B Biol. 2018, 182, 1–8. [Google Scholar] [CrossRef] [PubMed]
  230. Hobzova, R.; Hrib, J.; Sirc, J.; Karpushkin, E.; Michalek, J.; Janouskova, O.; Gatenholm, P. Embedding of Bacterial Cellulose Nanofibers within PHEMA Hydrogel Matrices: Tunable Stiffness Composites with Potential for Biomedical Applications. J. Nanomater. 2018, 2018, 1–11. [Google Scholar] [CrossRef] [Green Version]
  231. Almeida, I.F.; Pereira, T.; Silva, N.H.; Gomes, F.P.; Silvestre, A.J.; Freire, C.S.; Sousa Lobo, J.M.; Costa, P.C. Bacterial cellulose membranes as drug delivery systems: An in vivo skin compatibility study. Eur. J. Pharm. Biopharm. Off. J. Arb. Fur Pharm. Verfahr. 2014, 86, 332–336. [Google Scholar] [CrossRef]
  232. Trovatti, E.; Freire, C.S.; Pinto, P.C.; Almeida, I.F.; Costa, P.; Silvestre, A.J.; Neto, C.P.; Rosado, C. Bacterial cellulose membranes applied in topical and transdermal delivery of lidocaine hydrochloride and ibuprofen: In vitro diffusion studies. Int. J. Pharm. 2012, 435, 83–87. [Google Scholar] [CrossRef]
  233. Trovatti, E.; Silva, N.H.C.S.; Duarte, I.F.; Rosado, C.F.; Almeida, I.F.; Costa, P.; Freire, C.S.R.; Silvestre, A.J.D.; Neto, C.P. Biocellulose Membranes as Supports for Dermal Release of Lidocaine. Biomacromolecules 2011, 12, 4162–4168. [Google Scholar] [CrossRef]
  234. Piasecka-Zelga, J.; Zelga, P.; Szulc, J.; Wietecha, J.; Ciechanska, D. An in vivo biocompatibility study of surgical meshes made from bacterial cellulose modified with chitosan. Int. J. Biol. Macromol. 2018, 116, 1119–1127. [Google Scholar] [CrossRef] [PubMed]
  235. Chiaoprakobkij, N.; Seetabhawang, S.; Sanchavanakit, N.; Phisalaphong, M. Fabrication and characterization of novel bacterial cellulose/alginate/gelatin biocomposite film. J. Biomater. Sci. Polym. Ed. 2019, 30, 961–982. [Google Scholar] [CrossRef] [PubMed]
  236. Wan, Y.; Gao, C.; Han, M.; Liang, H.; Ren, K.; Wang, Y.; Luo, H. Preparation and characterization of bacterial cellulose/heparin hybrid nanofiber for potential vascular tissue engineering scaffolds. Polym. Adv. Technol. 2011, 22, 2643–2648. [Google Scholar] [CrossRef]
  237. Veliz, D.S.; Alam, C.; Toivola, D.M.; Toivakka, M.; Alam, P. On the non-linear attachment characteristics of blood to bacterial cellulose/kaolin biomaterials. Colloids Surfaces. B Biointerfaces 2014, 116, 176–182. [Google Scholar] [CrossRef] [PubMed]
  238. Wu, L.; Zhou, H.; Sun, H.-J.; Zhao, Y.; Yang, X.; Cheng, S.Z.D.; Yang, G. Thermoresponsive Bacterial Cellulose Whisker/Poly(NIPAM-co-BMA) Nanogel Complexes: Synthesis, Characterization, and Biological Evaluation. Biomacromolecules 2013, 14, 1078–1084. [Google Scholar] [CrossRef] [PubMed]
  239. Mohammadi, H. Nanocomposite biomaterial mimicking aortic heart valve leaflet mechanical behaviour. Proc. Inst. Mech. Eng. Part H J. Eng. Med. 2011, 225, 718–722. [Google Scholar] [CrossRef] [PubMed]
  240. Azuma, C.; Yasuda, K.; Tanabe, Y.; Taniguro, H.; Kanaya, F.; Nakayama, A.; Chen, Y.M.; Gong, J.P.; Osada, Y. Biodegradation of high-toughness double network hydrogels as potential materials for artificial cartilage. J. Biomed. Mater. Res. Part A 2007, 81, 373–380. [Google Scholar] [CrossRef] [PubMed]
  241. Kumbhar, J.V.; Jadhav, S.H.; Bodas, D.S.; Barhanpurkar-Naik, A.; Wani, M.R.; Paknikar, K.M.; Rajwade, J.M. In vitro and in vivo studies of a novel bacterial cellulose-based acellular bilayer nanocomposite scaffold for the repair of osteochondral defects. Int. J. Nanomed. 2017, 12, 6437–6459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Shi, Q.; Li, Y.; Sun, J.; Zhang, H.; Chen, L.; Chen, B.; Yang, H.; Wang, Z. The osteogenesis of bacterial cellulose scaffold loaded with bone morphogenetic protein-2. Biomaterials 2012, 33, 6644–6649. [Google Scholar] [CrossRef]
  243. Vadaye Kheiry, E.; Parivar, K.; Baharara, J.; Fazly Bazzaz, B.S.; Iranbakhsh, A. The osteogenesis of bacterial cellulose scaffold loaded with fisetin. Iran J. Basic Med. Sci. 2018, 21, 965–971. [Google Scholar] [CrossRef] [PubMed]
  244. Pigossi, S.C.; de Oliveira, G.J.P.L.; Finoti, L.S.; Nepomuceno, R.; Spolidorio, L.C.; Rossa, C., Jr.; Ribeiro, S.J.L.; Saska, S.; Scarel-Caminaga, R.M. Bacterial cellulose-hydroxyapatite composites with osteogenic growth peptide (OGP) or pentapeptide OGP on bone regeneration in critical-size calvarial defect model. J. Biomed. Mater. Res. Part A 2015, 103, 3397–3406. [Google Scholar] [CrossRef] [PubMed]
  245. Weyell, P.; Beekmann, U.; Kupper, C.; Dederichs, M.; Thamm, J.; Fischer, D.; Kralisch, D. Tailor-made material characteristics of bacterial cellulose for drug delivery applications in dentistry. Carbohydr. Polym. 2019, 207, 1–10. [Google Scholar] [CrossRef] [PubMed]
  246. Novaes, A.B., Jr.; Novaes, A.B. Bone formation over a TiAl6V4 (IMZ) implant placed into an extraction socket in association with membrane therapy (Gengiflex). Clin. Oral Implant. Res. 1993, 4, 106–110. [Google Scholar] [CrossRef] [PubMed]
  247. Wang, J.; Gao, C.; Zhang, Y.; Wan, Y. Preparation and in vitro characterization of BC/PVA hydrogel composite for its potential use as artificial cornea biomaterial. Mater. Sci. Eng. C 2010, 30, 214–218. [Google Scholar] [CrossRef]
  248. Goncalves, S.; Rodrigues, I.P.; Padrao, J.; Silva, J.P.; Sencadas, V.; Lanceros-Mendez, S.; Girao, H.; Gama, F.M.; Dourado, F.; Rodrigues, L.R. Acetylated bacterial cellulose coated with urinary bladder matrix as a substrate for retinal pigment epithelium. Colloids Surf. B Biointerfaces 2016, 139, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  249. Mohd Amin, M.C.I.; Ahmad, N.; Halib, N.; Ahmad, I. Synthesis and characterization of thermo- and pH-responsive bacterial cellulose/acrylic acid hydrogels for drug delivery. Carbohydr. Polym. 2012, 88, 465–473. [Google Scholar] [CrossRef]
  250. Qin, G.; Panilaitis, B.J.; Kaplan, Z.S.D.L. A cellulosic responsive “living” membrane. Carbohydr. Polym. 2014, 100, 40–45. [Google Scholar] [CrossRef]
  251. Marques, P.A.A.P.; Nogueira, H.I.S.; Pinto, R.J.B.; Neto, C.P.; Trindade, T. Silver-bacterial cellulosic sponges as active SERS substrates. J. Raman Spectrosc. 2008, 39, 439–443. [Google Scholar] [CrossRef]
  252. Wang, W.; Li, H.-Y.; Zhang, D.-W.; Jiang, J.; Cui, Y.-R.; Qiu, S.; Zhou, Y.-L.; Zhang, X.-X. Fabrication of Bienzymatic Glucose Biosensor Based on Novel Gold Nanoparticles-Bacteria Cellulose Nanofibers Nanocomposite. Electroanalysis 2010, 22, 2543–2550. [Google Scholar] [CrossRef]
  253. Eisele, S.; Ammon, H.P.; Kindervater, R.; Grobe, A.; Gopel, W. Optimized biosensor for whole blood measurements using a new cellulose based membrane. Biosens. Bioelectron. 1994, 9, 119–124. [Google Scholar] [CrossRef]
Figure 1. The microscopic cytoplasm synthesis of bacterial cellulose.
Figure 1. The microscopic cytoplasm synthesis of bacterial cellulose.
Materials 15 01054 g001
Figure 2. The advantages of bacterial cellulose.
Figure 2. The advantages of bacterial cellulose.
Materials 15 01054 g002
Figure 3. A schematic illustration of a bacterial cellulose composite.
Figure 3. A schematic illustration of a bacterial cellulose composite.
Materials 15 01054 g003
Figure 4. The major applications of bacterial cellulose-based composites.
Figure 4. The major applications of bacterial cellulose-based composites.
Materials 15 01054 g004
Table 1. Biomedical applications of bacterial cellulose.
Table 1. Biomedical applications of bacterial cellulose.
Anatomical PartTissue TypeApplicationCompositionQualitative PropertiesReferences
SkinEpithelial tissue (soft tissue)Wound restorative therapyBC-modified topographyWound healing enhancement: collagen migration enabled at the wound site along with fibroblast infiltration[211]
BC-CuO membraneProper antimicrobial activity against Escherichia coli and Staphylococcus aureus. It may function as a prototype for other similar products exhibiting photocatalyst and antimicrobial characteristics[212]
TEMPO-oxidized BC-AgNPsAntimicrobial activity with 12% Ag release rates (37 °C).[213]
BC-TiO2Antibacterial activity against Staphylococcus aureus and Escherichia coli proven on mice[214]
BC-AgNPs nanocompositeAntibacterial activity against Escherichia coli, Staphylococcus aureus, and Pseudomonas aeruginosa due to the release of Ag; inflammation reduction[215,216,217]
BC-ZnO nanocompositeAntibacterial activity against Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, Citrobacter freundii[213]
BC-propolis extractAnti-inflammatory, antibacterial activity, and antioxidant functions on diabetic wounds[209,218]
BC-phenolic acids membranesSuitable anti-inflammatory and antioxidant effects; non-cytotoxicity[219]
Periodate oxidized BC-chloramphenicolAntibacterial spectrum, biodegradable, and biocompatible[220]
BC-vaccarinDe novo formation, neovascularization of tissues made of collagen, and fibrous connective tissue[221]
BC-diethyldithiocarbamateOH-slow releasing systems: parasitic-caused lesion size reduction, SOD inhibition[222]
BC-ε-poly-l-lysine nanocompositeExtended antimicrobial spectrum[223]
BC-acrylic acid hydrogelPromoter of complete healing of wounds: water absorption and retention with good mechanical properties.[210,224]
BC-poly-methyl methacrylateBiodegradable bandages, which support wound healing[225]
BC-Octenidine-Poloxamer
BC-CMC-Methotrexate
Ready to use topical drug delivery systems: controlled release of active substances, effective for infected wounds[226,227]
BC-acrylic acid-human keratocytes and dermal fibroblasts hydrogelSame wound healing properties as plain BC and a prolific cell carrier[224]
Enzymatic degradative biomaterials for surgical suturesBC nanocrystals-regenerated chitin fibersWound healing enhancer with adaptable degradation rate (chitin concentration), biodegradable, strong suture material[228]
Tissue restorationBC-tuned porosityMuscle cell growth enhanced due to pore diameter, but slight strength reduction[82]
BC membraneAppropriate nanomorphological properties, optimal control of infection, capacity to retain moisture; adequate drug delivery system[229]
BC-PHEMA hydrogel matricesMesenchymal stem cells proliferation proven in rats[230]
Connective tissue
(transdermal level)
Active ingredients for transdermal releaseBC-chloro-aluminum phthalocyanine membraneSkin cancer: delivery system for photodynamic therapy with adequate properties for topical administration[231]
BC-lidocaine/ibuprofen membranePossibility of drug bioavailability modulation-dermal administration of lidocaine and ibuprofen[232,233]
Dressing materialsModified BC-chitosanAbdominal hernia treatment-reduced chance of infections caused by the mesh, no irritation, no hypersensitivity at implant site[234]
BC-sericin-PHMB filmHealing acceleration: low inflammatory response, high degree of collagen formation, scar shrinkage[192]
BC-alginate-gelatin filmOptimal ductility, biocompatibility, increased flexibility, and capacity to absorb water.[235]
Blood vesselsConnective tissueRestoration replacementBC-Fe3O4NPs magnetic pellicleSmall capillarity blood vessels[230]
Biosynthetic blood vesselsBC-polyglycolic acid and expanded polytetrafluorethyleneBiocompatibility (absence of leukocyte activation), apoptotic cell absence, vascularized granulation tissue, and multiple proliferating cells[208]
Engineered vessels with anticoagulant propertyBC-heparin nanofibrous scaffoldAnticoagulant properties-sulphate groups-enriched BC-heparin hybrid[236]
Blood cloth controlBC from nata de coco-kaolinTopographical properties and malleability of the biomaterial exceed the attraction forces between clotted blood proteins[237]
Vascular embolization: interventional therapiesBC-poly-N-isopropyl acrylamide-co-butyl methacrylate nanogelThermosensitive injectable biomaterials: expanded to condensed gel state[238]
Aortic heart valveConnective tissueProspective replacement therapyBC-PVA hydrogelBiomimicry: non-linear mechanical properties[239]
CartilagesConnective tissueReplacement, reconstructionBC-poly(dimethyl acrylamide) double network gelMeets properties of artificial cartilage; no in vivo tests confirmation[240]
BC-PVA compositeProven elasticity and similar properties to native cartilages[193]
Osteochondral defect treatmentBilayer BC-hydroxyapatite and BC-glycosaminoglycan indiceAccelerated recovery of articular cartilage and subchondral bone in model rats with osteochondral defects[241]
BoneSkeletal tissueAdvanced regenerationBC-bone mesenchymal protein-2 scaffoldsOsteogenesis in rat ectopic models[242]
Regeneration, reconstructionBC-Fisetin scaffold indiceBone matrix induced biosynthesis[243]
Gums and
Teeth
Connective tissueEarly stages of regenerationBC-hydroxyapatite-osteogenic growth peptide nanocompositeOsteoblast differentiation[244]
Tooth extraction or transplantation of oral mucosaNative and oxidized BC-doxycyclineDental dressings with potential of biodegradability, antimicrobial activity against pathogenic oral bacteria, and suitable drug delivery system[245]
Periodontal tissue recovery after dental implantsInner membrane of BC and external alkali-cellulose (Gengiflex®)Osseo-deficiency treatment: inflammatory response diminished, reduced number of surgical steps, restoration of mouth functions, and aesthetic role[246]
EyeCorneal epithelial tissueArtificial corneal biomaterialBC/PVA hydrogelSuitable water content, high visible light transmittance, UV absorbance, proper strength, and thermal properties[247]
Retinal pigment epithelium (RPE)TransplantAcetylated BC-urinary bladder matrixAppropriate features as cell carriers in potential RPE transplantation[248]
Gastro-intestinal levelConnective and epithelial tissues
(Simulated gastric and intestinal fluid)
Drug delivery systemBC-polyacrylic acid-bovine albumin (various concentration) hydrogelOptimization of drug release rate: pH dependent (similar to plain BC membranes)[249]
Abbreviations: AgNPs—Silver (Ag) nanoparticles, CuO—Copper oxide, Fe3O4NPs—Iron oxide nanoparticles, PHEMA—Poly(2-hydroxyethyl methacrylate), PHMB—Polyhexamethylene biguanide, PVA—Poly(vinyl alcohol), SOD—Superoxide dismutase, TEMPO-oxidized (2,2,6,6-tetramethyl piperidine oxide), TiO2—Titanium Dioxide, ZnO—Zinc oxide.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Popa, L.; Ghica, M.V.; Tudoroiu, E.-E.; Ionescu, D.-G.; Dinu-Pîrvu, C.-E. Bacterial Cellulose—A Remarkable Polymer as a Source for Biomaterials Tailoring. Materials 2022, 15, 1054. https://doi.org/10.3390/ma15031054

AMA Style

Popa L, Ghica MV, Tudoroiu E-E, Ionescu D-G, Dinu-Pîrvu C-E. Bacterial Cellulose—A Remarkable Polymer as a Source for Biomaterials Tailoring. Materials. 2022; 15(3):1054. https://doi.org/10.3390/ma15031054

Chicago/Turabian Style

Popa, Lăcrămioara, Mihaela Violeta Ghica, Elena-Emilia Tudoroiu, Diana-Georgiana Ionescu, and Cristina-Elena Dinu-Pîrvu. 2022. "Bacterial Cellulose—A Remarkable Polymer as a Source for Biomaterials Tailoring" Materials 15, no. 3: 1054. https://doi.org/10.3390/ma15031054

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