Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Biodegradable graphene nanocomposites as functional biomaterials: a review of their role in controlled drug delivery and tissue engineering

Md. Mohiuddina, Md. Mahbubur Rahman*a, Md. Nizam Uddin*b, Rakib Hasanc and Ismail Rahman*d
aDepartment of Mechanical Engineering, Khulna University of Engineering & Technology, Khulna 9203, Bangladesh. E-mail: mahbub_rahman@me.kuet.ac.bd
bJames C. Morriss Division of Engineering, Texas A&M University-Texarkana, 7101 University Ave, Texarkana, TX 75503, USA. E-mail: muddin@tamut.edu
cDepartment of Mechanical, Aerospace, and Industrial Engineering, The University of Texas at San Antonio, 1 UTSA Circle, San Antonio, TX 78249, USA
dInstitute of Environmental Radioactivity, Fukushima University, 1 Kanayagawa, Fukushima City, Fukushima 960-1296, Japan. E-mail: immrahman@ipc.fukushima-u.ac.jp

Received 23rd August 2025 , Accepted 23rd October 2025

First published on 19th November 2025


Abstract

Biodegradable graphene nanocomposites (BGNs) have emerged as highly versatile platforms at the intersection of nanotechnology, materials science, and biomedicine. By combining the exceptional physicochemical properties of graphene-based materials with the biocompatibility and environmental sustainability of biodegradable polymers, BGNs constitute a unique class of materials for advanced biomedical applications. Key features of BGNs, such as high surface area, tunable surface chemistry, excellent mechanical strength, and the ability to interface effectively with biological systems, make them promising candidates for controlled drug delivery and tissue engineering. In drug delivery, BGNs facilitate high drug loading and enable spatially and temporally controlled release, which can be triggered by internal or external stimuli, thereby improving therapeutic efficiency while minimizing side effects. In tissue engineering, the mechanical robustness and customizable structure of BGNs support cellular attachment, proliferation, and differentiation, rendering them suitable as scaffolds for regenerating bone, cartilage, skin, and neural tissues. This review explores recent advancements in the fabrication techniques and biomedical applications of BGNs, emphasizing their role in achieving precise drug delivery and effective tissue regeneration.


1 Introduction

1.1 Background and rationale

Conventional drug administration systems present several limitations, including poor bioavailability, uncontrolled release kinetics, off-target effects, and systemic toxicity.1,2 Similarly, traditional materials used in tissue repair often lack the requisite biocompatibility and mechanical properties, leading to failed integration with native tissue, implant failure, and adverse immune responses.3 These significant clinical challenges have motivated researchers to develop advanced materials capable of enhancing therapeutic efficacy, promoting predictable tissue regeneration, and minimizing adverse effects.

In recent decades, biomaterials science has progressed significantly, leading to the introduction of biodegradable polymers for biomedical applications. These polymers offer favorable properties, including minimal toxicity and high biocompatibility.4 In drug delivery, they enable stimuli-responsive (e.g., to temperature, pH, light) controlled release of therapeutic agents,5 while in tissue engineering, they provide a temporary structural support that degrades in coordination with new tissue formation.6 However, the utility of many biodegradable polymers is constrained by inherent limitations in mechanical strength, electrical conductivity, or targeted functionality, properties which are essential for more demanding applications like load-bearing tissue repair or electro-active tissue stimulation.4,7

To overcome these shortcomings, research has shifted toward nanotechnology, which enables the engineering of materials at the molecular scale. Among various nanomaterials, graphene and its derivatives, graphene oxide (GO) and reduced graphene oxide (rGO), have gained substantial attention. Their exceptional characteristics, including an extremely high surface-area-to-volume ratio, low weight, excellent thermal and mechanical properties, and the capacity to adsorb drugs or bioactive molecules via non-covalent interactions, make them ideal reinforcing and functional agents.8 However, the standalone use of pristine graphene-based materials faces its own challenges, including potential cytotoxicity,9 poor physiological stability, and a tendency to aggregate in aqueous biological environments, which can limit their practical application.10

Consequently, a synergistic approach has emerged: incorporating these nanomaterials into biocompatible polymeric matrices to improve their biostability, biocompatibility, and overall functionality. This has led to the development of BGNs, a class of materials that strategically combines the functional advantages of graphene with the safety and biodegradability of polymers. In these composites, graphene provides significant mechanical reinforcement, high drug-loading capacity, and responsiveness to external stimuli. Simultaneously, the polymer matrix ensures biocompatibility and produces safe, absorbable degradation byproducts. Furthermore, BGNs are valued for their sustainable nature, especially when fabricated with natural polymers, addressing the growing need for environmentally conscious medical technologies.

1.2 Scope and methodology

Numerous reviews have addressed the biomedical applications of biodegradable polymers and graphene-based materials separately. Comprehensive analyses exist for chitosan-functionalized GO in cancer therapy,11 GO-based hydrogels for drug delivery,12 natural polymeric nanobiocomposites for anti-cancer therapeutics,13 and polysaccharide-based nanomedicines for cancer immunotherapy.14 Other reviews have covered topics such as plasma modification of drug delivery systems,15 bacterial cellulose for wound dressings,16 3D bioprinting with chitosan,17 and nanostructured composites for bone regeneration.18

Although these articles provide valuable insights, most focus on either drug delivery or tissue engineering, and many concentrate solely on either biodegradable polymers or graphene. Reviews that explore the synergy of polymer–graphene nanocomposites in both drug delivery and tissue engineering are limited. This highlights a gap in the literature: a comprehensive review that simultaneously addresses both applications using BGNs and explores advancements in their fabrication is lacking. As shown in Fig. 1, research on BGNs for drug delivery and tissue engineering has gained significant attention since 2014, indicating a rapidly emerging field that warrants an in-depth, integrated review.


image file: d5ra06280b-f1.tif
Fig. 1 Annual number of publications on BGNs since 2010. Data sourced from ‘https://scopus.com’.

This review bridges this gap by systematically examining both the fabrication techniques and the dual applications of BGNs in controlled drug delivery and tissue engineering. A comprehensive literature search was conducted using the Scopus database with the keywords: (“biopolymer” OR “natural polymer” OR “biodegradable polymer”) AND (“graphene” OR “graphene oxide” OR “rGO”) AND (“nanocomposite” OR “nanomaterial”) AND (“drug delivery” OR “controlled release”) OR (“tissue engineering” OR “regenerative medicine”). This query yielded 221 records published between 2010 and June 2025. Following the PRISMA guidelines (Fig. 2), records were screened, and 66 primary research articles meeting all inclusion criteria were selected for synthesis. These core studies, supplemented by additional references for background context, form the basis of this review.


image file: d5ra06280b-f2.tif
Fig. 2 PRISMA flow diagram illustrating the study selection process for the systematic review.

2 Foundational materials

2.1 Graphene and its oxidized derivatives

Graphene is a single layer of carbon atoms arranged in a two-dimensional honeycomb lattice (Fig. 3a). Each carbon atom is sp2 hybridized, forming strong covalent bonds approximately 0.142 nm in length.19 This unique structure imparts exceptionally high electrical conductivity, mechanical strength (with a Young's modulus of ∼1 TPa), thermal conductivity, and a large theoretical specific surface area (∼2630 m2 g−1). Each carbon atom in graphene is covalently bonded to three neighboring atoms, forming a robust and stable hexagonal framework. This sp2 bonding provides delocalized π-electrons across the basal plane, which are responsible for graphene's high conductivity and chemical inertness.20 While chemically inert, its surface can be functionalized through covalent or non-covalent methods to attach biomolecules or drugs, though its intrinsic hydrophobicity can be a challenge for biological dispersion.21,22 Such functionalization exploits graphene's π-system for noncovalent π–π stacking with aromatic drug molecules, while covalent grafting (e.g., carbodiimide-mediated amidation) allows linkage of –COOH groups on GO with –NH2 groups of polymers like chitosan. Similarly, hydroxyl and epoxy groups on GO readily form hydrogen bonds with hydrophilic polymers (e.g., PVA, cellulose), enabling stable polymer–graphene nanocomposites. These interfacial interactions dictate dispersion, mechanical reinforcement, and drug release behavior. Such functionalization strategies are crucial for biomedical applications because they allow the attachment of drugs, peptides, or targeting ligands, improving dispersibility and specificity.23 Common synthesis methods include mechanical exfoliation, which yields high-quality flakes but lacks scalability, and chemical vapor deposition, which allows for large-area film production.24 Other routes include epitaxial growth on SiC substrates,25 unzipping of carbon nanotubes,26 and chemical reduction of GO.27 Each method produces graphene with different levels of purity, layer control, and defect density, which in turn affect its chemical reactivity and biomedical performance.
image file: d5ra06280b-f3.tif
Fig. 3 Schematic representations of (a) pristine graphene, (b) graphene oxide, (c) reduced graphene oxide, and 3D models of (d) single-layer graphene, (e) bi-layer graphene, and (f) tri-layer graphene stack.

GO is derived from graphite via aggressive oxidation and subsequent exfoliation. Its structure contains a high density of oxygen-containing functional groups (hydroxyl, epoxy, and carboxyl groups) that disrupt the planar sp2 network, introducing sp3-hybridized carbon atoms and rendering it electrically insulating but highly hydrophilic and easily dispersible in water (Fig. 3b).22 This aqueous dispersibility and the abundance of functional groups for further chemical modification make GO particularly attractive for biomedical applications such as drug delivery, biosensing, and tissue engineering.28 The type and density of oxygen functionalities depend on the oxidation method used (Hummers', Brodie's, or Staudenmaier's), which not only influence dispersibility but also toxicity and stability.29,30 Hummers' method involves strong acids (typically sulfuric acid) and oxidants like potassium permanganate to rapidly oxidize graphite, producing GO with various oxygen functional groups.31 Brodie's method, one of the earliest, uses fuming nitric acid and potassium chlorate, resulting in slower oxidation and more defects.32 Staudenmaier's method improves on Brodie's by combining concentrated sulfuric and nitric acids with potassium chlorate for faster oxidation and higher oxygen content, but it produces hazardous chlorine dioxide gas.33 These oxygen functionalities (–OH, –COOH, –O–) provide chemical handles for bonding with polymers: –COOH groups undergo esterification or amidation with polymeric hydroxyl/amine groups, while hydroxyl and epoxy moieties participate in hydrogen bonding and ionic crosslinking. Such interactions govern composite stability, swelling, and enzymatic degradation in physiological environments.

rGO is produced by removing a significant portion of the oxygen-containing groups from GO through chemical, thermal, or electrochemical reduction (Fig. 3c). This process partially restores the sp2 carbon network, improving electrical conductivity to a level between that of pristine graphene and GO, although residual defects remain.19,34 Reduction routes include chemical reductants (e.g., hydrazine, sodium borohydride, or green alternatives such as ascorbic acid), high-temperature annealing, and electrochemical reduction. Each pathway affects the residual oxygen content and defect density, which determine conductivity, dispersibility, and biocompatibility.35 Chemical reduction of GO typically involves the use of strong reducing agents like hydrazine or sodium borohydride, which react with oxygen-containing functional groups (such as epoxides, hydroxyls, and carboxyls) to remove them from the GO surface. This significantly restores the conjugated sp2 carbon network but can introduce defects and leave toxic residues, negatively affecting electrical and structural properties.35 Green reductants like ascorbic acid operate through similar mechanisms but involve milder redox reactions, offering moderate deoxygenation while maintaining better biocompatibility and aqueous dispersibility.36 Thermal annealing, on the other hand, works by heating GO to high temperatures (typically 500–1100 °C) under inert or reducing atmospheres (e.g., argon or hydrogen), which causes decomposition of oxygen functional groups and reconstructs sp2 domains. This method yields highly conductive rGO but often at the expense of scalability and may cause layer restacking.37 Finally, electrochemical reduction involves applying a potential across GO films in an electrolyte solution, triggering electron transfer that selectively removes oxygen groups. This technique avoids harsh chemicals and allows fine-tuning of reduction levels, producing rGO with controlled surface chemistry and good dispersibility.38 The residual oxygen groups on rGO still permit limited hydrogen bonding or ionic interactions with polymers, while defect sites act as nucleation points for covalent grafting. Thus, tuning the reduction level enables control over degradation rate, electrical conductivity, and mechanical reinforcement of the resulting nanocomposites. The resulting material offers a tunable balance of conductivity and dispersibility, making rGO a cost-effective and versatile option for similar biomedical applications.19,24

Graphene materials can exist as monolayer, bilayer, trilayer, or multilayer sheets (Fig. 3d–f). As the number of layers increases, the specific surface area and some unique quantum electronic properties decrease, but the potential for functionalization for biomedical use remains a key feature.22,24 Specifically, one, two, and three layers are termed monolayer, bilayer, and trilayer graphene, respectively, while 5–10 layers are considered few-layer graphene and 20–30 layers as multilayer graphene (nanocrystalline thin graphite).

2.2 Common biodegradable polymers

Biodegradable polymers, derived from both natural and synthetic sources, are central to the development of BGNs. They provide essential biocompatibility and degrade into non-toxic byproducts that can be safely metabolized or excreted by the body, aligning with the need for sustainable and eco-friendly biomedical solutions.39 When combined with graphene, these polymers form composites with enhanced mechanical strength, electrical conductivity, and tailored functionality, making them ideal for drug delivery and tissue engineering.40

Table 1 provides an overview of common biodegradable polymers used in BGNs, detailing their origin, formulation possibilities, and key functional properties.

Table 1 Overview of biodegradable polymers for drug delivery and tissue engineering applications
Polymer Origin/type Formulation possibilities Key functional properties References
(A) Natural proteins
Silk fibroin (SF) Silkworm-derived (Bombyx mori) Fabricated as hydrogels, porous sponges/scaffolds, films, fibers, and nanoparticles High mechanical strength; tunable degradation; minimal immunogenicity 41 and 42
Gelatin Denatured collagen from animal bone/skin Used in hydrogels, porous scaffolds, films, microspheres, and electrospun nanofibers Non-toxic and non-immunogenic; promotes cell adhesion (collagen-mimetic RGD sequences) 43 and 44
Collagen (COL) Major ECM protein from connective tissues Sponges/fibrous scaffolds, hydrogels, sheets (via molding, electrospinning, 3D bioprinting) Low immunogenicity; contains natural cell-binding sites (RGD) and directs cell behavior 45
Zein Corn protein (GRAS excipient) Films/coatings, nanoparticles, electrospun fibers, scaffolds Biodegradable, biocompatible protein; hydrophobic (alcohol-soluble) enabling controlled drug release; good film-former with moderate mechanical strength; supports drug-loaded membranes and tissue scaffolds 46
[thin space (1/6-em)]
(B) Natural polysaccharides and derivatives
Chitosan (CS) Derived from chitin in crustacean shells Processed into hydrogels, nanoparticles, nanofibers, membranes, films, and 3D scaffolds Inherently mucoadhesive and hemostatic; antimicrobial; low immune rejection 47–49
Alginate From brown seaweed Formed into ionically crosslinked hydrogels/gels, porous foams/sponges, microcapsules/microspheres, fibers, and films Low-toxicity; undergoes mild Ca2+-induced gelation; yields high-porosity, tunable-stiffness networks 50
Cellulose nanofibers Plant-derived nanoscale fibers Fabricated into ionically crosslinked hydrogels, freeze-dried porous scaffolds/aerogels and composite matrices Extremely high surface area and tensile strength; forms interconnected porous networks that support cell growth and may confer antibacterial effects 51
Starch Plant polysaccharide (amylose/amylopectin) Hydrogels (drug-loaded matrices), electrospun nanofibrous scaffolds, films, microparticles Highly hydrophilic (water-absorbing); low-cost; promotes cell proliferation and wound healing 52 and 53
Agarose Red algae-derived galactose polymer Thermoresponsive hydrogels (injectable gels, bioinks), cryogels, sponges/scaffolds (e.g. 3D-printed) Reversible gelation (thermo-sensitive sol–gel transition); good mechanical strength and high water retention; inert 54 and 55
Carboxymethyl cellulose (CMC) Cellulose-derived anionic polysaccharide Hydrogels, films, 3D porous scaffolds (e.g. for bone or soft tissue), wound dressings Very hydrophilic (forms swellable gels); thixotropic (viscoelastic) with high viscosity 56
Hyaluronic acid ECM glycosaminoglycan Hydrogels (injectable gels, bioinks, cryogels), 3D-printed scaffolds, composite matrices; also viscous solutions/films (e.g. eye drops, dermal fillers) Mucoadhesive (interacts with tissues like cartilage, skin); highly hydrophilic (excellent water retention); tunable viscoelasticity and porosity 57
Arabinoxylan (ARX) Plant hemicellulose Hydrogel matrices (films, injectable gels, tablets, capsules) and particulate systems (micro/nanogels) Polysaccharide; forms gel networks for controlled release; shown to promote wound healing 58 and 59
β-Glucan (BG) Found in cereal grains, fungi, yeast cell walls Porous scaffolds (e.g. freeze-dried foams, nanocomposites with hydroxyapatite) and hydrogels Highly hydrophilic (water-adsorbing) polymer; supports cell attachment and proliferation 60
Guar gum (GG) From guar bean (Cyamopsis tetragonolobus) Injectable hydrogels, films/membranes, and freeze-dried scaffolds (often blended with other polymers) Mucoadhesive polymer; exhibits gel-forming ability, high swellability and controlled-release characteristics 61 and 62
Chondroitin sulfate (CS-MA) Cartilage ECM-derived sulfated polysaccharide Photocrosslinked hydrogels (often blended with HA) for cartilage TE Cartilage-mimetic biopolymer; provides hydration and growth-factor binding; tunable mechanics, swelling, and enzymatic degradability 63
Carboxymethyl arabinoxylan (CMARX) Psyllium-derived hemicellulose Hydrogels, polyelectrolyte nanoparticles (e.g. with CS), films, composite scaffolds Biocompatible, biodegradable polysaccharide; CM modification increases crystallinity and thermal stability; supports pH-responsive swelling and sustained drug release 64
Kappa-carrageenan (K-CG) Red algae-derived sulfated polysaccharide Ionic hydrogels (beads, gels, membranes), composite scaffolds, electrospun mats Forms strong ionically crosslinked gels (with K+/Ca2+); highly hydrophilic (swelling); biodegradable; excellent encapsulation and sustained release of growth factors/drugs 65
2-Hydroxyethyl cellulose (2-HEC) Cellulose ether Cryogels, hydrogels, films, microparticles Biocompatible and biodegradable; highly hydrophilic with large swelling; forms viscous, stable gels; demonstrated use in sustained-release cryogel systems 66
Konjac glucomannan (KGM) From konjac tuber Hydrogels (thermal or alkali-induced gelation), sponges/aerogels, films Biocompatible and biodegradable; extremely high viscosity and water uptake; strong gelation (acetyl-dependent); forms robust matrices for sustained drug release and wound dressings 67
[thin space (1/6-em)]
(C) Synthetic polyesters and copolymers
Poly(glycerol sebacate) (PGS) Synthetic thermoset elastomer Made into porous elastomeric scaffolds (foams), fibrous meshes, and composite grafts Elastomer; rubber-like elasticity; tunable mechanical properties and degradation rate (matched to native soft tissues) 68 and 69
Poly(3-hydroxybutyrate) P(3HB) Bacterial polyester Scaffolds (e.g. bone, tissue engineering), cardiovascular patches, biodegradable microspheres/carriers Intrinsically osteoinductive (promotes MSC osteogenic differentiation and bone regeneration) 70 and 71
Poly(ε-caprolactone) (PCL) Aliphatic polyester Processed into porous scaffolds (e.g. bone or vascular), electrospun fiber mats, and biodegradable microspheres/nanoparticles Thermoplastic; good mechanical strength and toughness; easily fabricated (melt, 3D-printing, electrospinning) for sustained-release applications 72
Poly-L-lactic acid (PLLA) Aliphatic polyester 3D porous scaffolds (electrospun fibers, meshes, foams), microspheres/nanoparticles (drug carriers), composites (e.g. with bioceramics for bone TE) High tensile strength; tunable mechanical stiffness and degradation rate (via crystallinity/porosity); FDA-approved for implants 73 and 74
Polylactic acid (PLA) Aliphatic polyester 3D-printed scaffolds, electrospun mats, films, microparticles, sutures Biodegradable (hydrolyzes to lactic acid); high tensile strength and stiffness (brittle); hydrophobic; tunable crystallinity; biocompatible but relatively inert (often blended or surface-modified for cell adhesion) 75
Poly(lactic-co-glycolic acid) (PLGA) Copolymer Microparticles, nanoparticles, 3D scaffolds, implants Biocompatible and biodegradable with tunable degradation rate (by LA/GA ratio); relatively hydrophobic; good mechanical strength; extensively used for controlled release of drugs, proteins and growth factors 76
Poly(propylene fumarate) (PPF) Unsaturated polyester Injectable/crosslinked hydrogels, 3D-printed scaffolds, composites Biodegradable (degrades to fumaric acid and propylene glycol); tunable mechanical properties and degradation rate; inherently crosslinkable (unsaturated sites) for high-strength bone/regenerative scaffolds 77
Poly(glycolic acid-co-propylene fumarate) (PGA-co-PF) Copolymer Electrospun nanofibrous scaffolds, composite fibers Biodegradable copolymer; high water uptake and accelerated degradation when formulated with nanofillers; reinforced scaffolds exhibit enhanced mechanical strength and osteoconductivity (e.g. with GO/hydroxyapatite) 78
[thin space (1/6-em)]
(D) Semi-synthetic methacrylated polymers
AESO (acrylated epoxidized soybean oil) Acrylated vegetable oil Photocurable 3D-printed resins and hydrogels (often composite with nano-hydroxyapatite for bone TE) Thermoset resin; UV-crosslinkable (photopolymerizable) with tunable mechanics; supports cell growth/differentiation 79 and 80
Alginate-methacrylate (AlgMA) Methacrylated alginate Photocrosslinked hydrogels and 3D-printed scaffolds Tunable stiffness and degradation (hydrolytic/enzymatic) 80
Gelatin-methacryloyl (GelMA) Methacrylated gelatin Photocurable hydrogels and bioinks for 3D bioprinting; microcarriers Retains natural RGD and matrix metalloproteinases-cleavage sequences 81
Chondroitin sulfate-methacrylate (CS-MA) Methacrylated cartilage ECM Photocrosslinked hydrogels (often blended with HA) for cartilage TE Cartilage-mimetic biopolymer; provides hydration and growth-factor binding; tunable mechanics, swelling, and enzymatic degradability 63
[thin space (1/6-em)]
(E) Natural or bioactive polymers
Peptides Natural or synthetic short proteins/polypeptides Self-assembling hydrogels and nanofiber networks (3D scaffolds), injectable gels, composite coatings; can form nanoparticles for drug delivery Often bioactive (e.g. signaling peptides); good mechanical stability and tissue-like elasticity in hydrogel form; tunable assembly (e.g. β-sheet fibers) 82


2.3 Polymer–graphene interfacial interactions

The interactions at the polymer–graphene interface are a key determinant of the performance of BGNs, and they have a significant impact on important parameters like dispersion, mechanical augmentation, degradation, and drug release behaviors. Usually, these interfacial interactions are divided into two categories: covalent bonds which involve the sharing of electron pairs between atoms and result in strong, stable chemical linkages such as amide, ester, or radical-mediated grafts; and non-covalent interactions which are weaker and reversible forces that do not involve sharing of electrons such as hydrogen bonding, electrostatic forces, π–π stacking, and hydrophobic interactions. Different classes of biodegradable polymers tend to favor different types of bonding depending on their functional groups.

Hydrogen bonding is a specific type of non-covalent interaction that occurs when a hydrogen atom, covalently attached to an electronegative atom (such as oxygen or nitrogen) within a molecule, experiences an attractive force to another electronegative atom with lone electron pairs. In GO, the abundant oxygenated functional groups including hydroxyl (–OH), carboxyl (–COOH), and epoxy (–C–O–C) act as both hydrogen bond donors and acceptors, allowing physiochemical interactions with complementary sites on surrounding polymers.83 Natural polysaccharides and their derivatives (such as alginate, cellulose, starch, hyaluronic acid, and carboxymethyl cellulose), as well as natural proteins (collagen, gelatin, silk fibroin), are especially suited for forming hydrogen bonds with GO. Their molecular structures are rich in hydroxyl, carboxyl, and amide groups, providing ample sites for hydrogen bonding interactions with the oxygenated functionalities on GO.83 It enhances water dispersibility by stabilizing the nanocomposite network, boosts swelling by increasing the matrix's hydrophilicity, and enables better control of drug release by slowing burst release and supporting sustained release. A dense hydrogen-bonding network also strengthens mechanical properties and elasticity while regulating degradation, which is important for biomedical hydrogels and drug delivery systems.84

Electrostatic interactions are fundamental to the interfacial binding between biodegradable polymers and graphene derivatives. Specifically, positively charged groups on cationic biodegradable polymers such as the amine groups found in chitosan and polycationic peptides are attracted to the negatively charged carboxyl groups on GO or other oxidized graphene surfaces. This electrostatic attraction enhances compatibility and mechanical integrity in composite materials, which is especially important in biomedical applications that require controlled interactions and stable matrices.85 Similarly, anionic biodegradable polymers including alginate, hyaluronic acid, and CMC can interact with amine functionalized or protonated graphene derivatives through complementary charge pairing. The carboxyl and sulfate groups in these polysaccharides form strong ionic bonds with positively charged sites on modified graphene. These electrostatic interactions often lead to pH responsive behavior, as the ionization states of both polymer and graphene surfaces change with pH, enabling tunable adhesion and controlled drug release.86,87

The basal plane of graphene and rGO provides an extended π-conjugated system, which facilitates π–π stacking and hydrophobic interactions with biodegradable polymers containing aromatic rings or hydrophobic backbones. Proteins such as gelatin and peptides enriched with aromatic residues such as phenylalanine, tyrosine, and tryptophan offer abundant sites for π–π interactions, while biodegradable polyesters like PLA, PLLA, PCL, and PLGA possess hydrophobic domains that participate in both π–π stacking and hydrophobic contacts with graphene surfaces. These interactions enhance compatibility and matrix integrity by stabilizing polymer chains along the graphene plane, promoting uniform dispersion and superior reinforcement.40,88 π–π stacking is particularly efficient at adsorbing aromatic drug molecules, increasing entrapment efficiency and allowing controlled release profiles due to strong molecular interactions at the interface. Furthermore, hydrophobic contacts between graphene and biodegradable polymer chains reduce water uptake and permeability, improving barrier properties and retarding degradation rates under aqueous conditions, an effect desirable for prolonged drug delivery and enhanced packaging performance.88

Amide bond formation typically utilizes the carboxyl (–COOH) groups abundant on GO surfaces. These can be activated using carbodiimide chemistry, most commonly with EDC (N-ethyl-N′-(3-dimethylaminopropyl)carbodiimide) and NHS (N-hydroxysuccinimide), to form amide bonds with amine containing biopolymers such as chitosan, collagen, or gelatin. This reaction proceeds by forming an active ester intermediate on GO that readily reacts with amino groups on the polymer, resulting in a stable amide linkage that covalently anchors the polymer chain to the graphene surface.89

Esterification reactions exploit the reaction between GO's carboxyl groups and hydroxyl (–OH) functional groups present in hydroxyl-rich biopolymers, such as polysaccharides and polyesters. This process, often catalyzed by acid or activation agents, leads to the formation of ester bonds, chemically grafting the polymer backbone onto the graphene architecture and yielding a network with improved mechanical and degradation resistance.89

Radical grafting is widely used for methacrylated biopolymers like GelMA, AlgMA, or CS-MA. Under UV irradiation, these methacryloyl groups form free radicals that can react with activated double bonds on the GO surface or pre-functionalized graphene, leading to the formation of robust covalent crosslinks. This method is especially effective in hydrogel synthesis, resulting in nanocomposite networks with high crosslinking density and mechanical stability suitable for scaffolds.90,91

3 Fabrication of BGNs

The fabrication of BGNs involves a variety of techniques aimed at achieving optimal dispersion of graphene, mechanical integrity, and functional performance. The choice of methods depends heavily on the desired final form (e.g., film, hydrogel, fiber, 3D scaffold) and application. Many studies employ hybrid approaches, combining multiple methods to leverage their respective advantages. For instance, sonication is often integrated with solvent casting or crosslinking to enhance nanoparticle dispersion. A comparative summary of the techniques discussed is provided in Table 2.
Table 2 Fabrication techniques for graphene–polymer composites: a comparative overview
Method Process steps Advantage/limitations References
Solvent casting Dissolve polymer and graphene (e.g. GO) in solvent; cast into mold; evaporate solvent to form film Simple, cheap, and scalable yields uniform thin films; enhances mechanical, barrier, and antibacterial properties 92–94
Limited 3D architecture control, potential nanoparticle aggregation, solvent residues
Solution casting with sonication As above but apply ultrasonic waves during mixing to disperse fillers; then cast and dry Improved nanoparticle dispersion compared to plain casting; more uniform composites 92, 95 and 96
Excessive sonication may degrade polymers, some aggregation and solvent residues may persist
Emulsification and crosslinking Create a (water-in-oil or double) emulsion of polymer solution containing graphene; solidify droplets by chemical crosslinking (e.g. GLA) Enables formation of micro/nanoparticles with high surface area and uniform graphene dispersion for drug delivery; crosslinking provides stability and tunable release 83 and 97
Use of surfactants/oil phases may affect biocompatibility; batch variability possible
Sonication-assisted (hybrid) synthesis Use ultrasonication to drive in situ reactions or mixing (e.g. solvothermal or co-precipitation) of graphene and other components Enables efficient exfoliation and uniform dispersion of graphene, scalable and simple, can be combined with hybrid synthesis for multifunctionality 98 and 99
Excessive sonication can cause defects and reduce flake size; risk of agglomeration or residual surfactants if not optimized
In situ polymerization Polymerize monomers in presence of graphene fillers (e.g. Michael addition, free-radical) to form network Enables uniform graphene dispersion and strong interfacial bonding, leading to enhanced mechanical, thermal, and drug release properties; allows covalent grafting and functionalization 99 and 100
Complex chemistry, need for initiators/catalysts, risk of aggregation at high graphene content, and potential cytotoxicity from residual chemicals
Free-radical polymerization (lyophilization) Initiate polymerization (e.g. acrylic monomers) in solution; freeze-dry the gel to form porous scaffold Creates porous 3D scaffolds after freeze-drying; controllable pore structure 96 and 101
Requires initiators and freeze-drying, which may affect drug stability and loading
Magnetic functionalization (co-precipitation) Synthesize magnetic nanoparticles (e.g. Fe3O4) onto graphene (GO) by co-precipitation; incorporate into polymer (e.g. via emulsification or casting) Enables magnetically targeted drug delivery and imaging; stable integration of Fe3O4 enhances multifunctionality 102 and 103
Potential agglomeration, requires precise chemical control
Electrospinning Apply high-voltage to polymer/graphene solution to draw nanofibers onto a collector; then (optionally) crosslink fibers (e.g. heat or ionic) Produces nanofibrous mats that mimic ECM, supporting cell growth and sustained drug release 104 and 105
Limited to thin mats and challenges in loading cells or drugs uniformly in 3D structures
UV-assisted crosslinking Mix photopolymerizable polymers with graphene and a photoinitiator; expose to UV light to induce covalent crosslinking into a network Rapid solidification; spatial control of crosslinking; GO reinforcement enhances stiffness 106
UV penetration depth limits thickness; photoinitiators may introduce toxicity
3D printing (additive manufacturing) Extrude thermoplastic melt or hydrogel ink layer-by-layer to build 3D scaffold; may include post-print coatings Enables precise, patient-specific scaffold geometry. Surface coatings (e.g. polydopamine) can aid GO deposition 107 and 108
High processing temperatures may limit cell/drug loading and can affect sensitive biomolecules


3.1 Solution-based and casting methods

3.1.1 Solvent casting. This widely used and straightforward method involves dissolving a polymer and dispersing graphene in a suitable solvent, casting the mixture onto a flat surface, and slowly evaporating the solvent to form a solid film (Fig. 4).109–111 During this process, solvent molecules act as mediators that control both polymer chain mobility and graphene dispersion within the solution. The viscosity of the casting solution and the polarity of the solvent influence the degree of mixing, with well-matched solvents promoting uniform dispersion and poor solvents increasing the risk of aggregation.112 As the solvent evaporates, the polymer chains reorganize and gradually entrap graphene sheets, locking in the noncovalent interactions (hydrogen bonding, electrostatic forces, or π–π stacking) formed in solution. The evaporation rate is a critical factor: slow solvent removal facilitates homogeneous film formation, whereas rapid evaporation can lead to nanoparticle migration or surface aggregation. In some systems, residual oxygen-containing groups on graphene derivatives may undergo limited condensation reactions with polymer hydroxyl or amine groups during drying, forming ester or amide linkages that further stabilize the composite structure.113 Solvent casting remains the simplest fabrication route, but its lack of 3D control makes it unsuitable for regenerative scaffolds compared to electrospinning or 3D printing. For instance, SF/GO nanocomposite films were fabricated by blending the components in solution, where the presence of GO induced a conformational transition in the SF matrix from a random coil to a more stable β-sheet structure, thereby enhancing mechanical integrity.114 This technique has also been used to create 2-HEC/graphene films with improved thermal stability115 and SA/hydroxyapatite (HA)/graphene nanoplatelets (GnP) bionanocomposite films for bone tissue engineering.116 A hybrid approach combining solvent casting with porogen leaching, using a sacrificial material like salt to create pores, has been used to create porous PPF/GO nanocomposites with a hierarchical architecture suitable for cell infiltration.117
image file: d5ra06280b-f4.tif
Fig. 4 Schematic representation of the solvent casting process for nanocomposite thin films.
3.1.2 Casting with sonication. To overcome the natural tendency of nanoparticles to agglomerate, high-frequency ultrasonic waves are applied to the polymer–graphene solution before casting (Fig. 5). The process, known as sonication, generates acoustic cavitation, which involves the formation and violent collapse of microscopic bubbles. This creates intense localized shear forces that effectively break apart particle aggregates, ensuring a uniform and stable dispersion.118–120 This method has been successfully used to fabricate PCL/PGS/GO tubular scaffolds with a uniform porous morphology,121 P(3HB)/GnP scaffolds for neuronal studies,122 and KGM/GO films with significantly improved mechanical strength (tensile strength of 183.3 MPa) and a well-organized, bioinspired brick-and-mortar structure.123
image file: d5ra06280b-f5.tif
Fig. 5 Schematic illustration of the casting process enhanced by ultrasonic sonication.
3.1.3 Emulsification. This process creates a stable dispersion of two immiscible liquids, such as oil and water, stabilized by a surfactant (Fig. 6). The emulsification step typically involves vigorous stirring or ultrasonication to generate fine droplets of one liquid phase dispersed within the other. Surfactant molecules orient at the interface, reducing interfacial tension and preventing droplet coalescence.124 In the context of polymer–graphene systems, the choice of surfactant and solvent pair strongly influences droplet size, stability, and the final morphology of the nanocomposite. Hydrophilic–lipophilic balance governs whether a water-in-oil or oil-in-water emulsion is formed, while the concentration of emulsifier controls droplet size distribution.125 Once the emulsion is stabilized, crosslinking or solvent evaporation solidifies the polymer phase, entrapping graphene sheets within spherical domains. The interaction between surfactant functional groups and graphene surfaces can also contribute to dispersion stability, as electrostatic repulsion or steric hindrance suppresses aggregation during processing.126 In nanocomposite fabrication, emulsification enables the uniform incorporation of graphene into a polymer matrix, often to form spherical micro- or nanoparticles.127,128 For example, a water-in-oil emulsification was used to fabricate starch/agarose/GO nanoparticles for 5-fluorouracil (5FU) delivery, achieving good colloidal stability and drug encapsulation.129 A more complex double emulsion (water-in-oil-in-water) method was employed to create a pH-sensitive CS/CMC/GQD/ZnO nanocomposite for quercetin delivery.130
image file: d5ra06280b-f6.tif
Fig. 6 Schematic illustration of emulsification.

3.2 Network formation and crosslinking

3.2.1 Crosslinking. This process forms covalent or non-covalent bonds between polymer chains or between polymers and nanofillers, creating a stable three-dimensional network. This network structure is fundamental to hydrogels and enhances the strength, thermal stability, and chemical resistance of the composite (Fig. 7). One common method is chemical crosslinking, which involves the formation of permanent covalent bonds using a chemical agent. For example, hydroxyl-functionalized polymers and polyurethane prepolymers can be used to form robust, crosslinked rGO composite films.97 Similarly, a folic acid (FA)–CS–GO quantum dot (GOQD) nanocomposite was synthesized by conjugating FA with CS via carbodiimide chemistry using 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide and N-hydroxysuccinimide, followed by mixing with GO quantum dots.131 In such carbodiimide-mediated reactions, the –COOH groups on GO or FA are activated to form O-acylisourea intermediates, which then react with –NH2 groups of chitosan, creating stable amide linkages. This covalent grafting not only improves dispersion and interfacial bonding but also alters degradation behavior, since the hydrolytic cleavage of amide bonds is slower compared to ester linkages. In contrast, physical crosslinking relies on weaker, reversible non-covalent interactions like hydrogen bonding, ionic interactions, or hydrophobic associations. GO/γ-poly(glutamic acid) films achieve a strong, nacre-like structure through a combination of hydrogen and ionic bonds (with Ca2+ ions).132 Similarly, GO–COL scaffolds can be formed via pH-dependent electrostatic self-assembly between the negatively charged GO and positively charged COL.133 These noncovalent systems are particularly relevant for drug delivery because protonation/deprotonation of functional groups (e.g., –COOH/–COO or –NH3+/–NH2) under physiological pH changes can reversibly weaken or strengthen the interactions, thereby triggering controlled drug release. Thermal crosslinking is another approach, achieved by heating the material to induce bond formation, often through carbon–carbon networks. In graphene-based biodegradable polymer composites, this method has been shown to significantly improve thermal conductivity and stability.134
image file: d5ra06280b-f7.tif
Fig. 7 Schematic illustration of polymer crosslinking mechanisms: (a) covalent bonding via cross-linkers and (b) non-covalent interactions via functional groups.135 Ramachandran et al., Cross-linking dots on metal oxides, NPG Asia Mater., 2019, 11, 19, under the terms of the Creative Commons Attribution 4.0 International Licensee (https://creativecommons.org/licenses/by/4.0/).

Another approach is UV-assisted crosslinking (photocrosslinking), which uses ultraviolet light to activate a photoinitiator, which then generates free radicals that initiate polymerization and crosslinking (Fig. 8). Upon exposure to UV light, photoinitiator molecules absorb photons and are promoted from their ground state to an excited triplet state. In this high-energy state, the photoinitiator can abstract a hydrogen atom from a neighboring polymer chain, generating a polymer radical and a hydrogenated photoinitiator radical. If an auxiliary crosslinking agent is present, the hydrogenated photoinitiator can also induce cleavage of carbon–carbon double bonds within the crosslinker, producing additional radicals. These transient radicals then recombine to form new covalent bonds between polymer chains and/or between polymer and crosslinker segments, resulting in a three-dimensional network.136,137 While photoinitiators and crosslinkers are commonly used, some modified polymers, such as methacrylated derivatives (e.g., GelMA, AlgMA), can undergo self-crosslinking under UV light with only a photoinitiator, eliminating the need for an external crosslinking agent.


image file: d5ra06280b-f8.tif
Fig. 8 Schematic representation of UV-assisted free radical crosslinking.

This technique offers rapid curing and spatial control and has been used to create AESO/PVDF/GO nanocomposites138 and to fabricate 3D bioprinted hydrogels from bioinks containing methacrylated polymers (AlgMA, GelMA) and GO for cartilage engineering.139 Finally, radiation-induced crosslinking using high-energy sources like γ-rays can generate free radicals on polymer chains and water molecules, initiating polymerization and network formation without the need for chemical initiators, as was done to fabricate sterile GO/(AAc-co-SA) interpenetrating network hydrogels.140

3.2.2 In situ polymerization. This elegant method involves synthesizing the polymer matrix directly in a solution containing pre-dispersed nanoparticles (Fig. 9). The process begins by uniformly dispersing nanoparticles in a liquid monomer or low-molecular-weight precursor. Polymerization is then initiated through heat, radiation, or chemical initiators, allowing polymer chains to grow around and chemically bond with the nanoparticles. This facilitates the integration of nanoparticles into the polymer network by enabling interactions during chain propagation and network formation.112 For instance, when GO or rGO is incorporated, its surface functional groups can form covalent or non-covalent bonds with the polymerizing matrix. In many systems, these groups can act as anchoring sites, leading to interfacial polymerization where the growing polymer chains either initiate from or become physically entangled with the nanoparticle surface.141 This results in improved filler–matrix compatibility, reduced agglomeration, and enhanced load transfer efficiency.142 Furthermore, in biodegradable polymer systems such as PLA, PCL, or chitosan, in situ polymerization can promote uniform dispersion and interfacial adhesion due to hydrogen bonding and polar interactions formed during the growth of the polymer chains. This approach promotes homogeneous filler distribution and allows for the formation of strong interfacial bonds between the growing polymer chains and the nanoparticle surface.143
image file: d5ra06280b-f9.tif
Fig. 9 Schematic representation of polymer nanocomposite fabrication via in situ polymerization.144 Adopted from Basavegowda et al., Advances in functional biopolymer-based nanocomposites for active food packaging applications, Polymers, 2021, 13(23), 4198, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

It has been used to create hybrid rGO–multi-walled carbon nanotubes (MWCNT) nanocarriers functionalized with CoNi2S4 and ZnO, coated with CS and alginate, nanoparticles for co-delivery of drugs and genes,145 and to fabricate thiol–maleimide hydrogels incorporating rGO for chemo-photothermal therapy via a Michael addition reaction, where hyaluronic acid was functionalized with thiol groups and CS was modified with maleimide groups.146 A solvent-free approach using in situ polycondensation has also been developed for PGS/gelatin/GO nanocomposites, avoiding the use of potentially toxic solvents.147

3.2.3 Free radical polymerization. This common chain-growth process involves initiation, propagation, and termination steps to form polymer chains from monomer precursors (Fig. 10). The process begins with the thermal or photochemical decomposition of initiators such as benzoyl peroxide, AIBN, or ammonium persulfate, generating free radicals that attack the π-bonds of vinyl monomers like acrylamide or methyl methacrylate. This initiates chain growth, where monomers are sequentially added during the propagation phase. Termination occurs through radical recombination or disproportionation, yielding stable polymer chains.148 When applied in the synthesis of graphene-based biodegradable nanocomposites, this method enables the polymer matrix to form in direct contact with dispersed graphene derivatives. The presence of oxygen-containing functional groups on materials like GO can influence radical polymerization kinetics, affect the local polymerization environment, and improve filler compatibility by modulating surface energy and dispersion stability.149 It is a versatile method for creating hydrogels and is often followed by lyophilization (freeze-drying) to produce highly porous scaffolds. This technique has been used to synthesize multifunctional porous scaffolds by grafting sodium alginate (SA) with acrylic acid (AAc) in the presence of nHA, SiO2, and GO,150 and to create hybrid scaffolds from GG, AAc, GO, and other nanoparticles for bone tissue engineering.151
image file: d5ra06280b-f10.tif
Fig. 10 Steps of free radical polymerization.152 Adopted from Ribas-Massonis et al., Free-radical photopolymerization for curing products for refinish coatings market, Polymers, 2022, 14(14), 2856, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

3.3 Advanced and hybrid fabrication techniques

3.3.1 Electrospinning. This technique uses a high-voltage electric field to draw exceedingly fine nanocomposite fibers (typically nanometer to micrometer scale) from a polymer–graphene solution (Fig. 11). Under high voltage, a polymer–nanofiller solution is ejected as a fine jet from a syringe, stretching and solidifying into nanocomposite fibers with embedded nanofillers as the solvent evaporates. The resulting non-woven nanofibrous mats have a high surface-area-to-volume ratio and a structure that can effectively mimic the native extracellular matrix (ECM).153,154 However, the resulting thin mats limit applications to skin or vascular grafts rather than bulk bone regeneration. The success of electrospinning depends on a delicate balance of solution parameters such as viscosity, surface tension, and electrical conductivity,155 all of which are strongly influenced by the presence of graphene derivatives. GO, for instance, increases the conductivity of the spinning solution due to its surface charges and polar groups, which can lead to finer fiber diameters and more uniform fiber morphology.156 Additionally, uniform dispersion of GO within the polymer matrix is crucial to avoid defects during fiber formation; this often requires prior functionalization or surfactant stabilization.157 The interactions between the polymer chains and the nanofillers during electrospinning can influence chain alignment and packing density, ultimately affecting mechanical properties and porosity of the final scaffold. Moreover, residual solvent–filler interactions and rapid solvent evaporation can induce local phase separation or filler aggregation if chemical compatibility is poor.155 Electrospinning has been used to fabricate SA/PVA/GnP composite wound dressings,158 CS-based nanofibrous mats with GO and carbon quantum dot (CQD)-doped TiO2 for accelerated wound healing,159 and plasmonic rGO@AuNP–PCL composite scaffolds for synergistic cancer therapy and nerve regeneration.160
image file: d5ra06280b-f11.tif
Fig. 11 Electrospinning setup for nanofiber fabrication.161 Adopted from Antonios Keirouz, Zhe Wang, Vundrala Sumedha Reddy, Zsombor Kristóf Nagy, Panna Vass, Matej Buzgo, Seeram Ramakrishna, and Norbert Radacsi, The History of Electrospinning: Past, Present, and Future Developments, Adv. Mater. Technol., 2023, 8, 2201723, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).
3.3.2 Wet spinning. In this process, a viscous polymer–graphene solution is extruded through a spinneret directly into a liquid coagulation bath. The solvent exchange between the extruded jet and the bath causes the polymer and any embedded nanofillers to solidify into continuous filaments (Fig. 12).162 Wet spinning relies on phase inversion, where the solvent in the polymer solution diffuses into the coagulation bath while nonsolvent molecules simultaneously diffuse into the jet. This induces polymer precipitation, driven by a decrease in solubility and thermodynamic instability of the polymer–solvent system. The presence of graphene-based nanofillers can significantly alter this process by influencing local viscosity, diffusion rates, and nucleation during solidification.163 If well-dispersed, graphene derivatives may serve as nucleating agents that guide polymer chain alignment and promote the formation of more crystalline or oriented domains along the fiber axis.164 However, poor dispersion or interfacial incompatibility can lead to phase separation or filler aggregation, resulting in heterogeneous fiber morphology.163 Surface-functionalized graphene can mitigate these issues by improving interfacial affinity with the polymer matrix, enhancing mechanical integrity.165 It has been used to fabricate flexible and electrically conductive alginate–graphene hydrogel biofibers with enhanced mechanical properties and thermal stability.166
image file: d5ra06280b-f12.tif
Fig. 12 Schematic representation of wet spinning procedure.
3.3.3 Freeze-drying (lyophilization). This dehydration process involves freezing a material (typically a hydrogel) and then reducing the surrounding pressure to allow the frozen water to sublimate directly from a solid to a gas phase.167 This gentle removal of solvent preserves the material's delicate porous structure. It is widely used as a crucial post-processing step after hydrogel formation to create stable, highly porous scaffolds suitable for tissue engineering, as demonstrated in the fabrication of CS–GO nanocomposites.168 GO–COL composite aerogels further exemplify this approach and were validated through in vivo studies.169
3.3.4 3D printing and bioprinting. This transformative additive manufacturing technology enables the layer-by-layer fabrication of complex, patient-specific scaffolds with precise control over geometry and porosity. In 3D printing, a thermoplastic polymer–graphene composite is typically melted and extruded as a filament.170,171 This has been used to create PCL/GO scaffolds with mussel-inspired coatings172 and PLA/GO scaffolds with controlled porosity.173,174 3D printing consistently outperforms traditional methods in shape fidelity and porosity control, though processing temperatures may denature bioactive molecules. 3D bioprinting advances this concept by directly printing “bioinks” containing living cells, biomaterials, and growth factors to construct living tissue structures (Fig. 13).175 In bioprinting, patient-specific images (CT/MRI) are converted into 3D models; cells are cultured and mixed with biomaterials to form bioink, which is then printed layer-by-layer and stabilized, followed by maturation to promote tissue formation.176 This has been successfully used to create cell-laden hydrogels for cartilage regeneration139 and functional, spontaneously beating cardiac rings from rGO-containing bioinks.177
image file: d5ra06280b-f13.tif
Fig. 13 Schematic workflow of the 3D bioprinting process for tissue engineering applications.176 Reproduced from Lima et al., 3D bioprinting technology and hydrogels used in the process, J. Funct. Biomater., 2022, 13(4), 214, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).
3.3.5 Magnetic functionalization. This involves the integration of magnetic nanoparticles, typically iron oxide (Fe3O4), into the composite to impart magnetic responsiveness. This is commonly achieved via an in situ co-precipitation method, where ferrous (Fe2+) and ferric (Fe3+) salts are mixed with GO in an aqueous medium under alkaline conditions. As the pH increases, usually above 10, iron ions react with hydroxide ions to form Fe(OH)2 and Fe(OH)3, which subsequently undergo nucleation and crystallization into magnetite (Fe3O4) nanoparticles. The oxygen-containing functional groups on GO act as nucleation and anchoring sites, enabling uniform deposition of nanoparticles onto the graphene surface (Fig. 14). The resulting magnetic BGNs can be guided by external magnetic fields for targeted drug delivery, manipulated for remote-controlled release, or used as contrast agents in magnetic resonance imaging.178,179 This method was used to create a multi-responsive SA-g-poly(2-hydroxypropyl methacrylamide) (PHPM)/mGO nanocomposite for etoposide delivery.180 And a CS-based composite with magnetic GQDs for transdermal microneedle arrays.181
image file: d5ra06280b-f14.tif
Fig. 14 Synthesis of magnetic nanoparticles.182 Adopted from Stiufiuc et al., Magnetic nanoparticles: synthesis, characterization, and their use in biomedical field, Appl. Sci., 2024, 14(4), 1623, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

4 Applications in controlled drug delivery

BGNs are highly effective platforms for controlled drug delivery due to their high drug-loading capacity, enhanced stability, and responsiveness to a range of biological and external stimuli. The large surface area and π-electron system of graphene facilitate high drug loading via π–π stacking and hydrogen bonding, while the polymer matrix ensures biocompatibility and biodegradability.183,184

4.1 Mechanisms of stimuli-responsive release

The “smart” behavior of BGNs stems from their ability to release therapeutic payloads in response to specific triggers. This enables on-demand, site-specific delivery, which is crucial for maximizing therapeutic efficacy and minimizing systemic toxicity.

One of the most widely exploited mechanisms is pH-sensitive release, which leverages the pH gradients that exist in the body and in pathological tissues. The mechanism typically relies on the ionization of functional groups within the polymer matrix or on the GO surface. For polymers with acidic (e.g., –COOH) or basic (e.g., –NH2) groups, a change in pH alters their charge state, leading to changes in swelling behavior, as illustrated in Fig. 15a.185 In an acidic environment, for instance, amine groups on CS become protonated (–NH3+), leading to electrostatic repulsion between polymer chains. This causes the hydrogel network to swell, increasing the mesh size and facilitating drug diffusion. At the same time, protonation weakens π–π stacking interactions and hydrogen bonding between chitosan chains and GO nanosheets, thereby loosening the carrier network. The protonation/deprotonation of GO's functional groups also plays a key role (Fig. 15b).186


image file: d5ra06280b-f15.tif
Fig. 15 pH-Responsive drug delivery system: (a) swelling and deswelling mechanism, (b) protonation and deprotonation mechanism.

At low pH, carboxyl groups are protonated, reducing electrostatic repulsion and weakening the hydrogen bonding interactions that hold the drug, thereby promoting its release. Conversely, at neutral or basic pH, deprotonation enhances interactions (e.g., hydrogen bonding or π–π stacking), stabilizing drug loading and slowing release.186,187 Additionally, hydrolysis of ester linkages in polyesters such as PLA and PLGA is accelerated in acidic environments, producing lactic and glycolic acid that further decreases local pH and autocatalyzes matrix degradation.188,189 In some systems, pH-sensitive bonds such as Schiff bases and acetals are commonly used in drug delivery systems because they selectively cleave in acidic environments. Under low pH, Schiff bases hydrolyze due to protonation of the imine nitrogen, allowing water to attack and cleave the bond (e.g., R–CH[double bond, length as m-dash]N–R′ + H2O → R–CHO + R′–NH2), leading to controlled drug release in acidic environments.190 For instance, imine (C[double bond, length as m-dash]N) bonds in Schiff base linkages undergo protonation followed by nucleophilic attack by water, whereas acetals (C–O–C) are cleaved through acid-catalyzed oxonium ion intermediates, both providing predictable release profiles in tumor-like acidic conditions.191,192

Another important mechanism is thermo-responsive release, which uses temperature as a trigger (Fig. 16). It is often designed using polymers like poly(N-isopropylacrylamide) (PNIPAAm) that exhibit a lower critical solution temperature (LCST). Below the LCST, the polymer is hydrophilic and swollen, retaining the drug. Above the LCST, the polymer undergoes a phase transition, this transition arises from disruption of hydrogen bonding between PNIPAAm's amide groups and water, leading to hydrophobic association within the network, which squeezes out the encapsulated drug.193,194 Graphene's excellent photothermal properties can be harnessed here; when incorporated into a thermosensitive polymer matrix, near-infrared (NIR) irradiation can be used to remotely heat the BGN above its LCST, triggering on-demand drug release.195


image file: d5ra06280b-f16.tif
Fig. 16 Schematic illustration of thermo-responsive drug release: at temperatures below the LCST, the nanocarrier remains swollen and retains the drug; upon heating above the LCST, the polymer network collapses, triggering.

Enzyme-responsive release offers high specificity by exploiting enzymes that are overexpressed in pathological tissues, such as matrix metalloproteinases (MMPs) in the tumor microenvironment. In the context of BGN, this responsiveness can be engineered by incorporating enzyme-cleavable peptide linkers either between the therapeutic agent and the nanocarrier or within the polymer backbone itself. Upon exposure to the target enzyme, specific peptide bonds (–CO–NH–) are hydrolyzed (R–CO–NH–R′ → R–COOH + H2N–R′), resulting in cleavage of the linker or degradation of the polymer matrix. This cleavage destabilizes the nanocomposite structure, triggering the controlled release of the encapsulated or conjugated drug. In particular, hydrophilic oxygen-containing groups on GO and surface functionalities of the polymer can enhance enzyme accessibility and interaction, facilitating more efficient degradation and payload release payload.196–198

Photo-responsive release, illustrated in Fig. 17, can be achieved through several pathways beyond the photothermal effect. Photo-cleavable linkers can be incorporated into the BGN structure; upon irradiation with a specific wavelength of light, these bonds break, releasing the drug (Fig. 17b).199 Alternatively, photoisomerization (Fig. 17a) uses molecules like azobenzene that change conformation under light, altering the carrier's structure and porosity to allow drug diffusion.200 Additionally, photosensitization involves a photosensitizer that, upon light exposure, produces reactive oxygen species (ROS) causing cell damage or drug release, forming the basis of photodynamic therapy (Fig. 17c).201 Similarly, photoactivation uses light to convert inactive prodrugs into active therapeutic forms at the target site (Fig. 17d).


image file: d5ra06280b-f17.tif
Fig. 17 Schematic representation of different photosensitive drug release mechanisms: (a) photoisomerization, (b) photocleavage, (c) photosensitization, and (d) photoactivation.202 Adopted from Fernández et al., Advances in functionalized photosensitive polymeric nanocarriers, Polymers, 2021, 13(15), 2464, under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

Finally, electro- and magnetically-triggered release offer remote control. Graphene's conductivity allows for electro-responsive systems where an applied electric field can induce conformational changes or electrophoretic movement to trigger release (Fig. 18).203 For magnetic release, incorporating magnetic nanoparticles allows an external oscillating magnetic field to induce localized hyperthermia or mechanical deformation of the scaffold, both of which can drive drug release.204,205


image file: d5ra06280b-f18.tif
Fig. 18 Electro-responsive release mechanism.

A comparative summary of these release mechanisms is provided in Table 3.

Table 3 Comparison of drug release mechanisms in biodegradable polymer–graphene nanocomposites
Release mechanism Advantages Disadvantages Usage frequency in BGN
pH-Sensitive • Targeted release in acidic microenvironment • Off-target release in any acidic site (e.g. stomach) Common
• Minimal release at normal pH • Requires precise tuning of polymer pKa
• Polymer swelling at low pH enhances release • Possible drug degradation in strong acid
Thermo-responsive • On-demand release via external heating • Risk of tissue damage if overheated Moderately used
• Synergistic with hyperthermia-based therapy • Requires effective heat delivery to target site (depth/penetration issues)
Enzyme-responsive • High specificity if target enzyme is unique • Complex design (requires enzyme-cleavable linkers) Less common
• Activated under mild physiological conditions • Variability in enzyme expression levels (patient-to-patient)
Photo-responsive • Precise spatiotemporal control • Phototoxicity/heating of healthy tissue is possible Less common
• Deep NIR penetration enables efficient photothermal ablation • Limited penetration for visible/UV light
Electroresponsive • Fine control over drug release by modulating electrical signal • Requires conductive materials Less common
• Can be easily integrated into implantable devices • Risk of electrochemical reactions and tissue irritation
• Limited penetration depth in tissue
Magnetically triggered • Remote, non-invasive control • Requires incorporation of magnetic nanoparticles Moderately used
• Enables localized hyperthermia and mechanical actuation • Risk of local overheating or unintended tissue exposure
• Magnetic targeting enhances accumulation at desired site


4.2 Drug diffusion mechanisms

In addition to stimuli-responsive release, the passive diffusion of drugs from the BGN matrix is a fundamental process. The rate of release is governed by the concentration gradient and the physical properties of the matrix.206 In erodible matrices, the drug is released as the polymer matrix gradually degrades or dissolves from the surface inward, exposing new layers of drug over time (Fig. 19a). Enzymatically degradable polymers (e.g., collagen, gelatin) coupled with GO generally degrade faster than synthetic polyesters, highlighting their suitability for soft tissue repair rather than long-term implants.207 In contrast, hydrophilic matrices swell upon contact with aqueous fluids, forming a gel layer (Fig. 19b). The drug must then diffuse through this viscous gel layer, and the release rate is controlled by the rate of swelling and the thickness of the gel.208 Polyester-based BGNs degrade primarily through hydrolysis, which is slowed by graphene reinforcement due to reduced water uptake and chain mobility.209 Interestingly, electrostatic polymer–graphene interactions accelerate degradation in acidic microenvironments by destabilizing the network, a feature exploited in tumor-targeted release systems.210 In reservoir-type systems, the drug is contained within a core that is encapsulated by a rate-controlling polymer membrane. The drug diffuses through this membrane to the external environment (Fig. 19c).211 Comparisons show that covalently bonded systems maintain mechanical stability during degradation, while non-covalent systems show early-stage swelling and burst release.212
image file: d5ra06280b-f19.tif
Fig. 19 Illustration of drug release from diffusion-controlled delivery systems, where the drug is uniformly distributed in: (a) an erodible polymer matrix, (b) a hydrophilic, swellable polymer matrix, and (c) a reservoir-type system.211,213 Adapted from Joseph et al., Emerging Bio-Based Polymers from Lab to Market: Current Strategies, Market Dynamics and Research Trends, C, 2023, 9(1), 30, and Hossain et al., Scope of bio-based nanoparticle targeted through the cancer zone to deactivate cancer affected cells, Chem. Phys. Impact, 2023, 6, 100180, both under the terms of the Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/).

The release kinetics in these systems are often described by mathematical models, such as Fickian diffusion, where release is proportional to the square root of time, or more complex non-Fickian (anomalous) models when polymer relaxation and drug diffusion rates are comparable.214 This evidence suggests that degradation can be fine-tuned by adjusting graphene loading and interfacial chemistry rather than polymer choice alone.

4.3 Applications in cancer therapy

BGNs have shown significant promise in cancer therapy by enabling targeted delivery and controlled release, thereby enhancing efficacy and minimizing side effects. In the context of breast cancer, multiple studies have targeted MCF-7 and MDA-MB-231 cells. A GO–CS–FA nanobiocomposite co-loaded with camptothecin and diindolylmethane achieved 95.7% inhibition of MCF-7 cells and showed enhanced in vivo bioavailability.215 A starch/agarose/GO system for 5FU delivery demonstrated accelerated release at pH 5.4.129 Combining chemo- and photothermal therapy, a dopamine–rGO hydrogel loaded with doxorubicin (DOX) reduced MCF-7 viability to 21% under NIR irradiation.146 For lung cancer, GO-based systems have been developed to target A549 and H1299 cells. An folic acid–CS–GOQD nanocomposite showed pH-responsive DOX release and folate receptor-mediated uptake, inducing apoptosis selectively in cancer cells.131 A multi-responsive SA-g-PHPM/mGO system for etoposide delivery was triggered by pH, NIR light, and magnetic fields, enabling precise, multi-modal therapy that led to a sharp reduction in H1299 cell viability.180 BGNs have also been applied to other cancers. For colon cancer, a CS/GO/TiO2/escin nanocomposite demonstrated significant cytotoxicity against COLO 205 cells through TiO2-induced reactive oxygen species (ROS) generation.216 For cervical cancer, an FA-functionalized gelatin-coated GO nanocarrier delivered chlorambucil with pH-responsive release kinetics.217 For melanoma, an rGO–5FU–CMARX hydrogel showed pH-dependent 5FU release and strong antibacterial activity.218

4.4 Treatment of infectious and other diseases

Beyond cancer, BGNs are being developed for a range of therapeutic areas. For infectious diseases, nanoengineered self-assembling peptide-based hydrogels incorporating GO have been used for the sustained release of drugs against tuberculosis (isoniazid), fungal infections (amphotericin B), and bacterial infections (ciprofloxacin).219 In the treatment of neurological disorders, double-network hydrogels composed of amino-acid-based networks reinforced with polyacrylamide (PAAm), PNIPAAm, agarose, or low-gelling agarose, containing GO or carbon nanotubes (CNTs) as photothermal agents, have been used for the on-demand, NIR light-triggered release of baclofen to treat severe spasticity.220 For gastrointestinal conditions, a CS–GQD/sodium salicylate@CMC hydrogel bead system was designed for oral delivery, protecting the drug from the acidic stomach and enabling controlled release in the intestines.221 Finally, in ocular drug delivery, a CS–GQD nanocomposite demonstrated enzyme-triggered release of latanoprost for glaucoma treatment, responding to lysozyme in tear fluid.222

A summary of selected studies on BGNs for controlled drug delivery is presented in Table 4.

Table 4 Applications of BGN systems in controlled drug delivery
Polymer matrix Graphene type and additives Application (target, drug, mechanism) Observed outcomes (e.g. release%, viability, responsiveness) References
CS rGO–5FU blended with CMARX (from Plantago ovata) Melanoma therapy and wound healing; pH-responsive 5FU delivery 93.1% release at pH 7.4; sustained release at pH 6.4; promotes skin cell proliferation; antibacterial vs. S. aureus, P. aeruginosa 218
FA-functionalized CS GOQDs; DOX Cancer therapy (A549, SH-SY5Y), folate-receptor targeting 57% DOX release at pH 5.5 vs. 12% at pH 7.4; nuclear fragmentation; less than 5% hemolysis; selective cytotoxicity 131
CS/CMC GQD beads in CMC, sodium salicylate Oral delivery, inflammation therapy (pH-responsive) Minimal release at pH 1.2; enhanced at pH 6.8 and 7.4; more than 70% HT29 cell viability 221
Alg/AAc GO Colon-specific drug delivery; pH-responsive cefadroxil release Minimal release at pH 1 (stomach); sustained release at pH 7 (intestine); GO reduces burst release and regulates swelling 140
SA/K-CG rGO Amoxicillin delivery for wound treatment 94% loading efficiency; 26% release at pH 5.5 vs. 34% at pH 7.4 over 95 h; Fickian release; strong antibacterial activity 223
SF GO (0.5–3 wt%) Controlled drug delivery (rhodamine B) and tissue engineering scaffolds β-Sheet content peaked at 1.0% GO then declined; reduced burst release and sustained RhB release; faster degradation with higher β-sheet 114
2-HEC HCl/HNO3-modified graphene Nanocomposite films for potential biomedical applications (drug delivery, biosensors) Enhanced thermal stability (11 °C increase in Tmax); hydrophilic transformation of graphene 115
CS GO–CS–FA decorated with camptothecin and diindolylmethane Breast cancer (MCF-7); targeted co-delivery system 95.67% inhibition (MCF-7); enhanced bioavailability (AUC ∼ 33[thin space (1/6-em)]858 ng mL−1 h−1); low renal/liver toxicity 215
Gelatin-coated GO FA-functionalized rGO Cervical cancer (Siha), chlorambucil delivery 82% release at pH 1.2, 62.1% at pH 5.4, 43.7% at pH 7.4; IC50: 125.9 µg mL−1 (vs. 86 µg mL−1 for free drug) 217
Starch/agarose GO Breast cancer (MCF-7), 5FU delivery High 5FU encapsulation (∼87.3%); significant MCF-7 growth inhibition; acidic pH triggers release (targeted tumor delivery) 129
SA grafted with PHPM mGO, Fe3O4-decorated GO Lung cancer (H1299), etoposide delivery Triple stimuli-responsive release (pH 5.5, NIR, magnetic); cell viability drops to ∼19.1% (H1299); acid/NIR/magnet enhance release 180
CS/CMC GQDs; ZnO nanoparticles Brain cancer (U-87 MG), quercetin delivery 49% inhibition of U-87 MG (cancer cells); 85% viability in L929 (normal); pH-sensitive release (t1/2 ≈ 72 h) 130
CS (folate-functionalized) rGO; NiO nanoparticles; FA targeting ligand Lung (A549) and breast (MCF-7), DOX delivery DOX release ∼98.6% at pH 5 vs. 9.6% at pH 7.4; A549 viability 12.3%, MCF-7 7.1%; low zebrafish toxicity 224
CS rGO, Pd nanoparticles Colon cancer (HT-29), dual-drug system (5FU + another) ∼95–98% encapsulation; pH-sensitive release; IC50: 9.87 µg mL; Fickian diffusion kinetics 225
CS (folic acid-modified) GO nanoscrolls; FA-functionalization; DOX + caffeic acid Lung carcinoma (A549), co-delivery of DOX and caffeic acid (CA) DOX release ∼83% at pH 5 vs. 71% at pH 7.4; selective apoptosis in A549, high viability in HEK293 (normal) 226
CS GO; TiO2 nanoparticles; escin Colon cancer (COLO 205) IC50 ≈ 22.7 µg mL−1 (COLO 205); induces ROS-mediated apoptosis; minimal toxicity to normal cells 216
CS GQDs Ocular drug delivery; latanoprost for glaucoma; real-time tracking via photoluminescence “On–Off–On” photoluminescence response with lysozyme; more than 80% cell viability; protective effect against H2O2-induced oxidative damage in human corneal epithelial cells 222
CS/alginate rGO; MWCNTs; CoNi2S4 and ZnO nanoparticles Co-delivery (HeLa/HEK-293), DOX and pCRISPR pH-Responsive sustained DOX release (more in acidic environment); enhanced cellular uptake; improved therapeutic efficacy (co-delivery) 145
Thiol–maleimide hydrogel Dopamine–rGO; DOX Breast cancer (MCF-7), chemo-photothermal therapy MCF-7 viability ∼21% under NIR; pH-sensitive DOX release accelerated by NIR (ΔT ≈ 22 °C); targeted chemo/photothermal effect 146
Self-assembling peptide hydrogel GO; (with antibiotics: isoniazid, amphotericin B, ciprofloxacin) Sustained delivery of TB/antifungal/antibacterial drugs Sustained, controlled release of each loaded drug (isoniazid, amphotericin B, ciprofloxacin); good biocompatibility 219
Double-network hydrogel (PAAm/agarose or PNIPAAm/agarose) GO, CNT (photothermal agents); Fmoc-protected amino acids Neurological (spasticity), baclofen release (NIR-triggered) NIR-triggered baclofen release; inclusion of GO/CNT enables photothermal actuation (CNT-containing gel shows highest release efficiency) 220
CS MrGO (Fe3O4@RGO) + Pluronic F127; α-mangosteen Breast cancer (MCF-7), α-mangosteen delivery Faster release at pH 5.5 (tumor-like) vs. 7.4; magnetic field enables targeting; inhibited MCF-7 proliferation 227
CS mGQDs Transdermal microneedle arrays; electrically triggered drug delivery 96.4% drug release under electrical stimulation vs. 25.7% passive diffusion; detachable design with PEG base for rapid fluid response 181
PCL GO, Fe3O4, TRAIL, DOX Magnetically-triggered cancer therapy (scaffold system) Magnetic stimulation enhances drug release; dual action therapy; localized delivery; in vivo validation not provided 228
Aminated CMC GO; Fe3O4 magnetic nanoparticles; curcumin Breast cancer (MDA-MB-231), curcumin delivery Enhanced cytotoxicity vs. free curcumin; rapid curcumin release at pH 5.5 vs. slower at 7.4; magnetic targeting improves localization 186
CMC N-doped graphene, imatinib mesylate pH-Responsive cancer therapy (simulated tumor microenvironment) ∼74% loading at pH 7.0 in 3 h; ∼58% release at pH 4.0; reduced release at neutral/basic pH 229


5 Applications in tissue engineering

In tissue engineering, scaffolds provide a three-dimensional framework that mimics the native ECM, offering physical support and biological cues to guide cell adhesion, proliferation, and differentiation for the regeneration of new tissue.230,231 BGNs are excellent scaffold materials because the biopolymer component provides essential biocompatibility and biodegradability, while the graphene component significantly enhances mechanical strength, stability, and bioactivity.232–235

5.1 Scaffold design principles and tissue engineering strategies

The success of a tissue engineering strategy depends on the careful design of scaffolds that can orchestrate a complex series of biological events. This can be achieved through two primary approaches: in vitro and in vivo tissue engineering.

The classic approach is in vitro tissue engineering, which involves creating tissues outside the body in a controlled laboratory setting (Fig. 20). The process begins with isolating cells from a patient or donor and seeding them onto a pre-fabricated 3D scaffold. This cell-seeded construct is then cultured in a bioreactor, which provides a dynamic environment with nutrients, oxygen, and mechanical or electrical stimuli to promote cell proliferation, differentiation, and ECM production. Once the engineered tissue reaches a desired level of maturity, it is implanted into the patient.236–238 However, implantation into the human body is not always the ultimate or mandatory step, as in vitro tissue constructs are also crucial for evaluating scaffold performance, studying cell–material interactions, modeling diseases, and testing drugs in a physiologically relevant environment.


image file: d5ra06280b-f20.tif
Fig. 20 Steps of in vitro tissue engineering.

In contrast, in vivo tissue engineering leverages the body's own regenerative capacity by using it as a natural bioreactor (Fig. 21). A biomaterial scaffold, often loaded with growth factors or other signaling molecules, is implanted directly into the site of injury. The scaffold is designed to recruit endogenous stem or progenitor cells from surrounding tissues. These recruited cells then populate the scaffold, differentiate into the appropriate cell types, and regenerate the damaged tissue in situ. As the new tissue forms, the scaffold gradually degrades and is replaced.239,240 However, studies in this area remain comparatively limited, as this approach requires significant prior research and validation to ensure safety and efficacy.


image file: d5ra06280b-f21.tif
Fig. 21 Steps of in vivo tissue engineering.

Most of the studies discussed in the following sections are thereby focused only up to various stages of the in vitro process. Some involve scaffold fabrication alone, others extend to cell seeding and bioreactor-based culture, while only a few report implantation into a living body. Understandably, in vivo studies are far less common than in vitro ones due to the extensive safety and regulatory requirements involved. Regardless of the approach, the scaffold's properties are paramount. It must be biocompatible, biodegradable at an appropriate rate, and possess sufficient mechanical integrity. Furthermore, its architecture must feature an interconnected porous network to allow for cell infiltration and nutrient transport.230,231,241 The incorporation of graphene into biopolymer scaffolds can enhance all these properties, improving mechanical strength while providing a surface that promotes cell adhesion and guides differentiation.234,235

5.2 Bone and cartilage regeneration

BGNs are particularly well-suited for bone and cartilage tissue engineering due to their ability to provide robust mechanical support and promote chondro/osteogenesis. In terms of osteogenic potential, numerous studies have shown that GO-based composites enhance osteoblastic differentiation. For example, PGA-co-PF/GO/HA electrospun scaffolds significantly increased alkaline phosphatase activity in MG63 cells,78 and SA/HA/GnP films promoted apatite-like mineral formation in simulated body fluid.116 For load-bearing applications, the reinforcing effect of graphene is critical. PCL scaffolds reinforced with 3% graphene achieved a compressive modulus of 136.74 MPa,242 while PLLA scaffolds incorporating 12% GO–hydroxyapatite reached a compressive strength of 21.52 MPa.243 In the challenging area of cartilage regeneration, biomimetic hydrogels composed of alginate, gelatin, chondroitin sulfate, and GO have been 3D bioprinted with mesenchymal stem cells, inducing intrinsic chondrogenic differentiation without exogenous growth factors.139

5.3 Neural and cardiac tissue engineering

For neural tissue, scaffolds must support cell viability and provide electrical cues to guide regeneration. Graphene's conductivity is a major advantage. GO–COL and rGO–COL scaffolds supported extensive Schwann cell spreading and attachment,133 while P(3HB) scaffolds containing GnP restored physiological firing patterns in cultured neurons.122 For cardiac tissue, which relies on electrical signal propagation for synchronized contractions, a 3D bioprinted cardiac “BioRing” fabricated from a GelMA/AlgMA/rGO bioink successfully mimicked key aspects of native heart tissue, including spontaneous and synchronous beating.177

5.4 Skin and wound healing

BGNs can accelerate wound healing by acting as both protective wound dressings and pro-regenerative templates. Electrospun mats of SA/PVA/GnP loaded with curcumin provided sustained antimicrobial and antioxidant effects,158 while GO–CS/PVP membranes enhanced skin wound repair in rat models, showing 33% faster wound closure than sterile gauze.244 For skin regeneration, CS–GO scaffolds implanted subdermally in rats supported tissue encapsulation and vascularization.168 Furthermore, PLLA nanofiber scaffolds coated with polydopamine and carbon nanomaterials such as GO and CNTs have been shown to exhibit piezoelectric behavior, suggesting they could generate therapeutic electrical cues to accelerate healing in response to body movement.245

A detailed summary of selected studies on BGNs for tissue engineering is presented in Table 5.

Table 5 Applications of BGN systems in tissue engineering
Polymer matrix Graphene type and additives Applications (target tissue and role) Observed outcomes References
PGA-co-PF GO, HA Bone – electrospun scaffold Increased alkaline phosphatase activity; osteoblastic differentiation; more than 5 times protein adsorption; greater than 98% metabolic activity 78
COL GO Neural – scaffold for nerve tissue engineering Schwann cell adhesion/spreading; 3D porous (20–100 µm); mechanically stable; cell infiltration 133
AESO/PVDF blend (UV-crosslinked) GO; curcumin Bone – bone tissue scaffold Semi-crystalline matrix with enhanced stiffness; antibacterial against common pathogens; supports osteoblast viability 138
Alginate/gelatin/chondroitin sulfate bioink GO Cartilage – 3D bioprinted scaffold Promoted intrinsic chondrogenic differentiation of stem cells; high cell viability; improved ECM synthesis 139
GelMA/AlgMA bioink rGO Cardiac – 3D bioprinted cardiac tissue scaffold Supported cardiomyocyte viability, alignment, and beating; enhanced electrical conductivity for synchronized activity 177
Type I COL GO (0.05–0.2% w/v) Bone – aerogel scaffold for cranial defects Enhanced compressive modulus (0.20–0.51 MPa); superior biomineralization (Ca/P = 1.67); 1.5× increased bone volume in vivo; improved rat bone marrow mesenchymal stem cell proliferation 169
SA GnP; HA Bone – scaffold for bone regeneration Nearly doubled tensile strength at 0.5% GP; promoted apatite-like mineralization; excellent biocompatibility and biodegradability 116
CS (anisotropic membrane) GO; HA Bone – bone-mimicking membrane scaffold Improved early osteoblast viability and long-term growth; enhanced cell spreading and adhesion; favorable microenvironment via GO/HA 246
PPF GO nanoribbons/nanoplatelets Bone – porous scaffold for bone regeneration 26% modulus increase; cell infiltration; ECM formation; enhanced osteoblast viability 117
PCL rGO–Ag nanoparticles Bone – film scaffold Increased mechanical/electrical properties; promotes stem cell differentiation; antimicrobial 247
CS beads GO, TiO2 nanoparticles Bone – injectable beads Less than 1% resorption over 90 days; osteoconductivity; COL type I formation; enhanced crystallinity 248
PCL/PGS tubular scaffold GO Nervous/vascular/renal – nerve conduit About 84% fibroblast viability; enhanced compressive modulus and thermal stability; channel-like porosity aiding regeneration 121
P(3HB) GnP Neural – conductive neuronal scaffold Restored physiological neuronal firing patterns; increased responsiveness at low stimulus; dense neuronal network formation 122
CS (GLA-crosslinked beads) GO; TiO2 nanoparticles; blackberry extract Bone – injectable bead scaffold Strong in vivo biocompatibility; stimulated new bone formation and mineral deposition; COL fiber development 249
KGM GO Scaffold through solvent casting Young's modulus: 16.8 GPa; tensile strength: 183.3 MPa; over 90% cell viability; strong cell adherence; bioactive and biocompatible 123
PLGA GO, MoS2 nanoplatelets Bone – porous scaffold 20–27% early bone regeneration; minimal inflammation; enhanced mechanical strength 250
CS GO Skin – scaffold Around 78 µm porosity; improved vascularization and healing; mild inflammation 168
SA/PVA GnP; curcumin Skin – electrospun wound dressing Controlled curcumin release (∼80% in 24 h); combined antimicrobial and antioxidant effects; supports tissue regeneration 158
PVA + CS nanofibers GO; CQD-doped TiO2 Skin – nanofiber wound healing scaffold Accelerated wound closure (greater than 93% in 14 days); promoted fibroblast migration; antibacterial against gram-positive and negative bacteria 159
CS/PVP nanofibers GO Wound healing – nanofibrous mat In vitro study: more cell viability (40%); faster wound closure (33%); mimics ECM; water-permeable 244
Zein nanofibers GO, curcumin Wound healing – dressing In vitro study: biphasic drug release; fibroblast proliferation; low swelling, controlled CUR release 251
PLGA/gelatin GO Bone – electrospun scaffold Increased ALP, RUNX2, calcium deposition; ECM-like morphology; supports osteogenic differentiation 252
PCL rGO, AuNPs Neural – aligned nanofiber scaffold Improved neurite outgrowth (2.5×); 90% Schwann cell viability; NIR-induced tumor cell ablation 160
Gelatin GO Bone – electrospun scaffold Improved Young's modulus (70%) and tensile strength (200%); supports osteoblast proliferation 253
Alginate GO Muscle – biofiber scaffold In vitro study: increased tensile strength/modulus; C2C12 myoblast viability and differentiation; myogenic morphology 166
PLLA GO–HA hybrid composite Bone – load-bearing scaffold Increased compressive strength (∼21.5 MPa) and modulus (∼5 GPa); apatite layer formation in simulated body fluid; excellent osteoblast compatibility 243
CS–PPPOEMA GO–Ag Wound healing – scaffold Improved wound closure (89.81%); antimicrobial activity; selective cytotoxicity 254
PGS/gelatin GO; Clay Scaffold through in situ polymerization Enhanced thermal and mechanical properties; controlled degradation; homogeneous filler dispersion 147
SA grafted AAc GO; nHA@SiO2 Bone – bioactive porous scaffold In vitro study: enhanced osteoblast proliferation and adhesion; increased mineralization; improved hydrophilicity and porosity 150
ARX and BG GO; nHA Bone – porous hydrogel scaffold High compressive strength and modulus; optimal porosity (∼55%) for cell migration; promoted osteogenic differentiation 255
Poly(acrylic acid) hybrid GO; nHA; TiO2; Ag-sulfadiazine Bone – antibacterial fracture scaffold Sustained silver drug release; increased mechanical strength; enhanced osteoblast adhesion and proliferation 151
PPF GO nanoplatelets, MoS2 nanoplatelets Bone – porous scaffold Enhanced compressive modulus (up to 108%); high cytocompatibility; ECM deposition 256
PPF CNTs, GO nanoribbons/nanoplatelets Bone – tissue scaffold High cell viability; mild degradation cytotoxicity; enhanced spreading and attachment 257
PCL (3D-printed) GO (mussel-inspired coating) Bone – surface-modified scaffold Improved osteoblast proliferation and differentiation; increased alkaline phosphatase activity and calcium deposition 172
PCL Graphene/GO Bone – 3D-printed scaffold Enhanced modulus (136.74 MPa); cell proliferation; trabecular bone mimicry 242
PLA GO Bone – 3D printed scaffold Improved biocompatibility; 2× mineralization; 30% increase in Young's modulus 173 and 174
GG hydrogel film rGO; TiO2 nanowires Skin – wound healing hydrogel In vitro study: stimulated fibroblast migration and wound closure; bioactive composite aiding skin regeneration 258


6 Future outlook

The integration of biodegradable polymers with graphene-based nanomaterials has produced a formidable class of BGNs with multifunctional advantages for biomedicine. This review has detailed their significant advancements in both controlled drug delivery and tissue engineering. Looking forward, the primary goal is to translate these promising materials from the laboratory to clinical settings. This transition requires comprehensive long-term in vivo studies focusing on BGN biocompatibility, degradation kinetics, biodistribution, and immunogenicity. Standardized testing protocols and adherence to good manufacturing practice guidelines are critical hurdles for achieving regulatory approval and enabling scalable, reproducible production.

Future research will likely focus on several key areas. One is the development of smart, multifunctional platforms (theranostics) capable of simultaneous diagnosis and therapy. At the chemical level, this will require more precise tailoring of interfacial chemistry between graphene derivatives and polymers. Another is 4D bioprinting, which can create scaffolds that change their shape or function in response to physiological stimuli after implantation. Advances here will depend not only on printing resolution but also on chemical innovation in photo-crosslinkable or reversible-bonding polymers, which dictate scaffold adaptability and long-term stability. BGNs are also exceptionally well-suited for personalized and AI-driven medicine, where therapies could be tailored based on a patient's specific needs. Mechanistic studies on hydrolytic, enzymatic, and oxidative degradation pathways will be critical to integrate predictive modeling with clinical translation. Finally, a growing focus on bioinspired and sustainable design, using green synthesis methods and mimicking natural tissue structures, will continue to drive the field forward. In particular, greener chemical synthesis routes for graphene and biodegradable polymers can help reduce cytotoxic byproducts while maintaining functional surface chemistry, aligning with both regulatory and sustainability goals.

7 Conclusions

BGNs represent a powerful convergence of nanotechnology, materials science, and biomedicine. These materials synergistically combine the exceptional mechanical, electrical, and responsive properties of graphene with the biocompatibility and sustainability of biodegradable polymers. In controlled drug delivery, BGNs enhance drug-loading capacity and enable targeted, stimuli-triggered release, improving therapeutic efficacy while minimizing side effects. In tissue engineering, they function as robust scaffolds that promote cellular growth and guide the regeneration of diverse tissues, including bone, skin, and cartilage. The fabrication of BGNs has progressed from simple solution-based methods to advanced techniques like 3D bioprinting, where innovations in crosslinking chemistry, surface functionalization, and reversible bonding are crucial for scaffold adaptability and stability. While BGNs hold immense promise for sustainable and effective biomedical solutions, further research, particularly rigorous in vivo experiments and clinical trials, is essential to advance these innovative materials toward practical medical applications.

Author contributions

M. Mohiuddin: writing – original draft, writing – review and editing, methodology. M. M. Rahman: writing – original draft, supervision, project administration, formal analysis, conceptualization. M. N. Uddin: writing – review and editing, supervision, project administration, conceptualization. R. Hasan: writing – review and editing, methodology, investigation. I. Rahman: writing – review and editing, investigation, formal analysis, visualization.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

All data used to prepare the manuscript are included. Additional explanations will be available upon request.

Acknowledgements

While preparing this work, the author(s) used Gemini Advanced, QuillBot, and Grammarly to paraphrase and edit the language. After using those tools, the author(s) reviewed and revised the content as needed and take(s) full responsibility for the publication's content.

References

  1. S. Adepu and S. Ramakrishna, Molecules, 2021, 26, 5905 CrossRef CAS PubMed.
  2. M. S. Rao, R. Gupta, M. J. Liguori, M. Hu, X. Huang, S. R. Mantena, S. W. Mittelstadt, E. A. Blomme and T. R. Van Vleet, Frontiers in Big Data, 2019, 2, 25 CrossRef PubMed.
  3. D. F. Williams, Front. Bioeng. Biotechnol., 2019, 7, 127 CrossRef PubMed.
  4. J. Kurowiak, T. Klekiel and R. Będziński, Int. J. Mol. Sci., 2023, 24, 16952 CrossRef CAS PubMed.
  5. C. M. Wells, M. Harris, L. Choi, V. P. Murali, F. D. Guerra and J. A. Jennings, J. Funct. Biomater., 2019, 10, 34 CrossRef CAS PubMed.
  6. M. Modrák, M. Trebuňová, A. F. Balogová, R. Hudák and J. Živčák, J. Funct. Biomater., 2023, 14, 159 CrossRef PubMed.
  7. V. Gayathri, T. Khan, M. Gowtham, R. Balan and T. A. Sebaey, Front. Bioeng. Biotechnol., 2025, 13, 1533944 CrossRef CAS PubMed.
  8. H. Zhao, R. Ding, X. Zhao, Y. Li, L. Qu, H. Pei, L. Yildirimer, Z. Wu and W. Zhang, Drug Discovery Today, 2017, 22, 1302–1317 CrossRef CAS PubMed.
  9. J. Zhang, H.-Y. Cao, J.-Q. Wang, G.-D. Wu and L. Wang, Front. Cell Dev. Biol., 2021, 9, 616888 CrossRef PubMed.
  10. H. Tang, S. Zhang, T. Huang, J. Zhang and B. Xing, Environ. Sci. Technol., 2021, 55, 14639–14648 CrossRef CAS PubMed.
  11. P. Rajput and S. Khanchandani, Int. J. Biol. Macromol., 2025, 311, 143999 CrossRef CAS PubMed.
  12. R. Saharan, S. K. Paliwal, A. Tiwari, M. A. Babu, V. Tiwari, R. Singh, S. K. Beniwal, M. Kumar, A. Sharma and W. H. Almalki, J. Drug Delivery Sci. Technol., 2024, 94, 105506 CrossRef CAS.
  13. A. Mondal, A. K. Nayak, P. Chakraborty, S. Banerjee and B. C. Nandy, Pharmaceutics, 2023, 15, 2064 CrossRef CAS PubMed.
  14. Y. Zeng, Y. Xiang, R. Sheng, H. Tomás, J. Rodrigues, Z. Gu, H. Zhang, Q. Gong and K. Luo, Bioact. Mater., 2021, 6, 3358–3382 CAS.
  15. P. Bhatt, V. Kumar, V. Subramaniyan, K. Nagarajan, M. Sekar, S. V. Chinni and G. Ramachawolran, Pharmaceutics, 2023, 15, 2066 CrossRef CAS PubMed.
  16. M. Horue, J. M. Silva, I. R. Berti, L. R. Brandão, H. d. S. Barud and G. R. Castro, Pharmaceutics, 2023, 15, 424 CrossRef CAS PubMed.
  17. S. Mallakpour, F. Sirous and C. M. Hussain, New J. Chem., 2021, 45, 10565–10576 RSC.
  18. E. V. Prathyusha, S. S. Gomte, H. Ahmed, A. Prabakaran, M. Agrawal, N. Chella and A. Alexander, Int. J. Biol. Macromol., 2024, 137834 Search PubMed.
  19. M. Azizi-Lalabadi and S. M. Jafari, Adv. Colloid Interface Sci., 2021, 292, 102416 CrossRef CAS PubMed.
  20. A. D. Ghuge, A. R. Shirode and V. J. Kadam, Curr. Drug Targets, 2017, 18, 724–733 CrossRef CAS PubMed.
  21. H. Shen, L. Zhang, M. Liu and Z. Zhang, Theranostics, 2012, 2, 283–294 CrossRef CAS PubMed.
  22. Y. Xiao, Y. X. Pang, Y. Yan, P. Qian, H. Zhao, S. Manickam, T. Wu and C. H. Pang, Adv. Sci., 2023, 10, 2205292 CrossRef CAS PubMed.
  23. A. Gostaviceanu, S. Gavrilaş, L. Copolovici and D. M. Copolovici, Int. J. Mol. Sci., 2024, 25, 10174 CrossRef CAS PubMed.
  24. M. S. A. Bhuyan, M. N. Uddin, M. M. Islam, F. A. Bipasha and S. S. Hossain, Int. Nano Lett., 2016, 6, 65–83 CrossRef CAS.
  25. H. I. Røst, R. K. Chellappan, F. S. Strand, A. Grubišić-Čabo, B. P. Reed, M. J. Prieto, L. C. Tănase, L. de Souza Caldas, T. Wongpinij and C. Euaruksakul, J. Phys. Chem. C, 2021, 125, 4243–4252 CrossRef.
  26. Y. Liu, J. Zheng, X. Zhang, K. Li, Y. Du, G. Yu, Y. Jia and Y. Zhang, J. Appl. Polym. Sci., 2021, 138, 50474 CrossRef CAS.
  27. J. Narayan and K. Bezborah, RSC Adv., 2024, 14, 13413–13444 RSC.
  28. X.-M. Han, K.-W. Zheng, R.-L. Wang, S.-F. Yue, J. Chen, Z.-W. Zhao, F. Song, Y. Su and Q. Ma, Am. J. Transl. Res., 2020, 12, 1515 CAS.
  29. V. Marsala, Y. Gerasymchuk, M. L. Saladino, E. Paluch, M. Wawrzyńska, V. Boiko, X. Li, C. Giordano, D. Hreniak and B. Sobieszczańska, Molecules, 2025, 30, 240 CrossRef CAS PubMed.
  30. J. V. Lim, S. T. Bee, L. Tin Sin, C. T. Ratnam and Z. A. Abdul Hamid, Polymers, 2021, 13, 3547 CrossRef CAS PubMed.
  31. W. S. Hummers Jr and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef.
  32. P. Feicht, J. Biskupek, T. E. Gorelik, J. Renner, C. E. Halbig, M. Maranska, F. Puchtler, U. Kaiser and S. Eigler, Chem.–Eur. J., 2019, 25, 8955–8959 CrossRef CAS PubMed.
  33. H. L. Poh, F. Šaněk, A. Ambrosi, G. Zhao, Z. Sofer and M. Pumera, Nanoscale, 2012, 4, 3515–3522 RSC.
  34. G. Utkan, G. Yumusak, B. C. Tunali, T. Ozturk and M. Turk, ACS Omega, 2023, 8, 31188–31200 CrossRef CAS PubMed.
  35. B. Lesiak, G. Trykowski, J. Tóth, S. Biniak, L. Kövér, N. Rangam, L. Stobinski and A. Malolepszy, J. Mater. Sci., 2021, 56, 3738–3754 CrossRef CAS.
  36. K. K. H. De Silva, H.-H. Huang and M. Yoshimura, Appl. Surf. Sci., 2018, 447, 338–346 CrossRef CAS.
  37. P. Zhang, Z. Li, S. Zhang and G. Shao, Energy Environ. Mater., 2018, 1, 5–12 CrossRef CAS.
  38. N.-F. Chiu, C.-D. Yang, C.-C. Chen and C.-T. Kuo, Sens. Actuators, B, 2018, 258, 981–990 CrossRef CAS.
  39. A. Samir, F. H. Ashour, A. A. A. Hakim and M. Bassyouni, npj Mater. Degrad., 2022, 6, 68 CrossRef CAS.
  40. E. Avcu, F. E. Bastan, M. Guney, Y. Yildiran Avcu, M. A. Ur Rehman and A. R. Boccaccini, Acta Biomater., 2022, 151, 1–44 CrossRef CAS PubMed.
  41. Z.-H. Li, S.-C. Ji, Y.-Z. Wang, X.-C. Shen and H. Liang, Front. Mater. Sci., 2013, 7, 237–247 CrossRef.
  42. Y. Lyu, Y. Liu, H. He and H. Wang, Gels, 2023, 9, 431 CrossRef CAS PubMed.
  43. R. Andreazza, A. Morales, S. Pieniz and J. Labidi, Polymers, 2023, 15, 1026 CrossRef CAS PubMed.
  44. X. Jiang, Z. Du, X. Zhang, F. Zaman, Z. Song, Y. Guan, T. Yu and Y. Huang, Front. Bioeng. Biotechnol., 2023, 11, 1158749 CrossRef PubMed.
  45. Y. Isobe, T. Kosaka, G. Kuwahara, H. Mikami, T. Saku and S. Kodama, Materials, 2012, 5, 501–511 CrossRef CAS PubMed.
  46. S. Tortorella, M. Maturi, V. V. Buratti, G. Vozzolo, E. Locatelli, L. Sambri and M. C. Franchini, RSC Adv., 2021, 11, 39004–39026 RSC.
  47. V. B. Sindhusha and J. N. Doraiswamy, Cureus, 2023, 15, e48667 Search PubMed.
  48. Z. Arabpour, F. Abedi, M. Salehi, S. M. Baharnoori, M. Soleimani and A. R. Djalilian, Int. J. Mol. Sci., 2024, 25, 1982 CrossRef CAS PubMed.
  49. T. M. M. Ways, W. M. Lau and V. V. Khutoryanskiy, Polymers, 2018, 10, 267 CrossRef PubMed.
  50. N. Farshidfar, S. Iravani and R. S. Varma, Mar. Drugs, 2023, 21, 189 CrossRef CAS PubMed.
  51. P. Zou, J. Yao, Y.-N. Cui, T. Zhao, J. Che, M. Yang, Z. Li and C. Gao, Gels, 2022, 8, 364 CrossRef CAS PubMed.
  52. C.-S. Lee and H. S. Hwang, Gels, 2023, 9, 951 CrossRef CAS PubMed.
  53. V. S. Waghmare, P. R. Wadke, S. Dyawanapelly, A. Deshpande, R. Jain and P. Dandekar, Bioact. Mater., 2018, 3, 255–266 Search PubMed.
  54. F. Jiang, X.-W. Xu, F.-Q. Chen, H.-F. Weng, J. Chen, Y. Ru, Q. Xiao and A.-F. Xiao, Mar. Drugs, 2023, 21, 299 CrossRef CAS PubMed.
  55. Y. Huang, S. Peng, Y. Chen and B. Chu, Gels, 2025, 11, 255 CrossRef CAS PubMed.
  56. M. S. Rahman, M. S. Hasan, A. S. Nitai, S. Nam, A. K. Karmakar, M. S. Ahsan, M. J. Shiddiky and M. B. Ahmed, Polymers, 2021, 13, 1345 CrossRef CAS PubMed.
  57. A. Di Mola, M. R. Landi, A. Massa, U. D'Amora and V. Guarino, Int. J. Mol. Sci., 2022, 23, 14372 CrossRef CAS PubMed.
  58. J. G. Pérez-Flores, L. García-Curiel, E. Pérez-Escalante, E. Contreras-López and E. J. Olloqui, Heliyon, 2024, 10, e25445 CrossRef PubMed.
  59. M. U. A. Khan, M. A. Raza, S. I. A. Razak, M. R. Abdul Kadir, A. Haider, S. A. Shah, A. H. Mohd Yusof, S. Haider, I. Shakir and S. Aftab, J. Tissue Eng. Regen. Med., 2020, 14, 1488–1501 CrossRef CAS PubMed.
  60. M. U. A. Khan, M. A. Al-Thebaiti, M. U. Hashmi, S. Aftab, S. I. Abd Razak, S. Abu Hassan, M. R. Abdul Kadir and R. Amin, Materials, 2020, 13, 971 CrossRef CAS PubMed.
  61. C. M. Robles-Kanafany, M. L. Del Prado-Audelo, M. González-Torres, D. M. Giraldo-Gomez, I. H. Caballero-Florán, M. González-Del Carmen, J. Sharifi-Rad, G. Figueroa-González, O. D. Reyes-Hernández and H. Cortés, Cell. Mol. Biol., 2021, 67, 89–95 CrossRef PubMed.
  62. H. Poudel, A. B. RanguMagar, P. Singh, A. Oluremi, N. Ali, F. Watanabe, J. Batta-Mpouma, J. W. Kim, A. Ghosh and A. Ghosh, Bioengineering, 2023, 10, 1088 CrossRef CAS PubMed.
  63. T. Hardingham, Osteoarthr. Cartil., 1998, 6, 3–5 CrossRef PubMed.
  64. M. Bhatia and M. Ahuja, Int. J. Biol. Macromol., 2015, 72, 495–501 CrossRef CAS PubMed.
  65. V. E. Santo, A. M. Frias, M. Carida, R. Cancedda, M. E. Gomes, J. F. Mano and R. L. Reis, Biomacromolecules, 2009, 10, 1392–1401 CrossRef PubMed.
  66. D. Momekova, E. Ivanov, S. Konstantinov, F. Ublekov and P. D. Petrov, Polymers, 2020, 12, 1172 CrossRef CAS PubMed.
  67. Y. Sun, X. Xu, Q. Zhang, D. Zhang, X. Xie, H. Zhou, Z. Wu, R. Liu and J. Pang, Polymers, 2023, 15, 1852 CrossRef CAS PubMed.
  68. M. Rosalia, D. Rubes, M. Serra, I. Genta, R. Dorati and B. Conti, Polymers, 2024, 16, 1405 CrossRef CAS PubMed.
  69. P. Piszko, M. Włodarczyk, S. Zielińska, M. Gazińska, P. Płociński, K. Rudnicka, A. Szwed, A. Krupa, M. Grzymajło, A. Sobczak-Kupiec, D. Słota, M. Kobielarz, M. Wojtków and K. Szustakiewicz, Int. J. Mol. Sci., 2021, 22, 8587 CrossRef CAS PubMed.
  70. B. Dariš and Ž. Knez, Acta Pharm., 2020, 70, 1–15 CrossRef PubMed.
  71. A. P. Bonartsev, V. V. Voinova, A. V. Volkov, A. A. Muraev, E. M. Boyko, A. A. Venediktov, N. N. Didenko and A. A. Dolgalev, Sovremennye Tehnologii v Medicine, 2022, 14, 78–90 CAS.
  72. R. Dwivedi, S. Kumar, R. Pandey, A. Mahajan, D. Nandana, D. S. Katti and D. Mehrotra, J. Oral Biol. Craniofac. Res., 2020, 10, 381–388 CrossRef PubMed.
  73. E. Capuana, F. Lopresti, M. Ceraulo and V. La Carrubba, Polymers, 2022, 14, 1153 CrossRef CAS PubMed.
  74. Y. Dai, T. Lu, M. Shao and F. Lyu, Front. Bioeng. Biotechnol., 2022, 10, 1011783 CrossRef PubMed.
  75. S. Castañeda-Rodríguez, M. González-Torres, R. M. Ribas-Aparicio, M. L. Del Prado-Audelo, G. Leyva-Gómez, E. S. Gürer and J. Sharifi-Rad, J. Biol. Eng., 2023, 17, 21 CrossRef PubMed.
  76. H. K. Makadia and S. J. Siegel, Polymers, 2011, 3, 1377–1397 CrossRef CAS PubMed.
  77. Z. Cai, Y. Wan, M. L. Becker, Y.-Z. Long and D. Dean, Biomaterials, 2019, 208, 45–71 CrossRef CAS PubMed.
  78. A. M. Díez-Pascual and A. L. Díez-Vicente, J. Mater. Chem. B, 2017, 5, 4084–4096 RSC.
  79. D. Mondal, A. Srinivasan, P. Comeau, Y.-C. Toh and T. L. Willett, Mater. Sci. Eng., C, 2021, 118, 111400 CrossRef CAS PubMed.
  80. S. C. Mauck, S. Wang, W. Ding, B. J. Rohde, C. K. Fortune, G. Yang, S.-K. Ahn and M. L. Robertson, Macromolecules, 2016, 49, 1605–1615 CrossRef CAS.
  81. H. Yin, M. Zhu, Y. Wang, L. Luo, Q. Ye and B. H. Lee, Frontiers in Soft Matter, 2023, 2, 1101680 CrossRef.
  82. R. Binaymotlagh, L. Chronopoulou and C. Palocci, J. Funct. Biomater., 2023, 14, 233 CrossRef CAS PubMed.
  83. A. M. Díez-Pascual, Polymers, 2021, 13, 2978 CrossRef PubMed.
  84. H. Xue, P. Wang, L. Ji, K. Zhang, S. Ge and J. Tan, J. Agric. Food Res., 2025, 24, 102395 CAS.
  85. W. Feng and Z. Wang, iScience, 2022, 25, 103629 CrossRef CAS PubMed.
  86. S. Zhang, J. Dong, R. Pan, Z. Xu, M. Li and R. Zang, Polymers, 2023, 15, 2149 CrossRef CAS PubMed.
  87. N. Sittisanguanphan, N. Paradee and A. Sirivat, J. Pharm. Sci., 2022, 111, 1633–1641 CrossRef CAS PubMed.
  88. R. Binaymotlagh, L. Chronopoulou and C. Palocci, Materials, 2025, 18, 2978 CrossRef CAS PubMed.
  89. H. J. Salavagione, G. Martínez and G. Ellis, Macromol. Rapid Commun., 2011, 32, 1771–1789 CrossRef CAS PubMed.
  90. R. Muñoz, L. León-Boigues, E. López-Elvira, C. Munuera, L. Vázquez, F. Mompeán, J. Á. Martín-Gago, I. Palacio and M. García-Hernández, ACS Appl. Mater. Interfaces, 2023, 15, 46171–46180 CrossRef PubMed.
  91. X. Liu, A. L. Miller II, B. E. Waletzki and L. Lu, J. Biomed. Mater. Res., Part A, 2018, 106, 1247–1257 CrossRef CAS PubMed.
  92. A. Kausar, I. Ahmad, M. H. Eisa, M. Maaza and H. Khan, Nanomanufacturing, 2023, 3, 1–19 CrossRef.
  93. S. J. Lee, S. J. Yoon and I.-Y. Jeon, Polymers, 2022, 14, 4733 CrossRef CAS PubMed.
  94. A. M. Díez-Pascual, Polymers, 2022, 14, 2102 CrossRef PubMed.
  95. M. Dong, Y. Sun, D. J. Dunstan, R. J. Young and D. G. Papageorgiou, Nanoscale, 2024, 16, 13247–13299 RSC.
  96. M. Silva, N. M. Alves and M. C. Paiva, Polym. Adv. Technol., 2018, 29, 687–700 CrossRef CAS.
  97. D. Spasevska, G. Leal, M. Fernández, J. B. Gilev, M. Paulis and R. Tomovska, RSC Adv., 2015, 5, 16414–16421 RSC.
  98. A. F. Ghanem and M. H. Abdel Rehim, Biomedicines, 2018, 6, 63 CrossRef PubMed.
  99. S. Qamar, N. Ramzan and W. Aleem, Synth. Met., 2024, 307, 117697 CrossRef CAS.
  100. L. Sisti, G. Totaro, A. Celli, L. Giorgini, S. Ligi and M. Vannini, Polymers, 2021, 13, 1377 CrossRef CAS PubMed.
  101. A. T. Lawal, Cogent Chem., 2020, 6, 1833476 CrossRef.
  102. S. Uten, P. Boonbanjong, Y. Prueksathaporn, K. Treerattrakoon, N. Sathirapongsasuti, N. Chanlek, S. Pinitsoontorn, P. Luksirikul and D. Japrung, ACS Omega, 2024, 9, 2263–2271 CrossRef CAS PubMed.
  103. F. Li, Y. Huang, K. Huang, J. Lin and P. Huang, Int. J. Mol. Sci., 2020, 21, 390 CrossRef CAS PubMed.
  104. R. Jiffrin, S. I. A. Razak, M. I. Jamaludin, A. S. A. Hamzah, M. A. Mazian, M. A. T. Jaya, M. Z. Nasrullah, M. Majrashi, A. Theyab, A. A. Aldarmahi, Z. Awan, M. M. Abdel-Daim and A. K. Azad, Polymers, 2022, 14, 3725 CrossRef CAS PubMed.
  105. A. M. Muñoz-Gonzalez, S. Leal-Marin, D. Clavijo-Grimaldo and B. Glasmacher, Int. J. Artif. Organs, 2024, 47, 633–641 CrossRef PubMed.
  106. W. Xu, X. Zhang, P. Yang, O. Långvik, X. Wang, Y. Zhang, F. Cheng, M. Österberg, S. Willför and C. Xu, ACS Appl. Mater. Interfaces, 2019, 11, 12389–12400 CrossRef CAS PubMed.
  107. P. H. N. Cardoso and E. S. Araújo, Compounds, 2024, 4, 71–105 CrossRef CAS.
  108. R. Patil and S. Alimperti, J. Funct. Biomater., 2024, 15, 82 CrossRef CAS PubMed.
  109. R. Mohan, A. Jineesh and N. M. Prabu, Environ. Eng. Res., 2022, 27, 210161 CrossRef.
  110. L. Di Muzio, P. Simonetti, V. C. Carriero, C. Brandelli, J. Trilli, C. Sergi, J. Tirillò, F. Cairone, S. Cesa, G. Radocchia, S. Schippa, S. Petralito, P. Paolicelli and M. A. Casadei, Molecules, 2022, 27, 2959 CrossRef CAS PubMed.
  111. N. Jouault, D. Zhao and S. K. Kumar, Macromolecules, 2014, 47, 5246–5255 CrossRef CAS.
  112. S. N. Tripathi, G. S. Rao, A. B. Mathur and R. Jasra, RSC Adv., 2017, 7, 23615–23632 RSC.
  113. S. Cheng and G. S. Grest, ACS Macro Lett., 2016, 5, 694–698 CrossRef CAS PubMed.
  114. X. Xu, Z. Ren, M. Zhang and L. Ma, J. Appl. Polym. Sci., 2021, 138, 51173 CrossRef CAS.
  115. E. Kılıç and N. P. Bayramgil, Results Chem., 2022, 4, 100544 CrossRef.
  116. S. Iswarya, T. Theivasanthi, K. Chinnaiah and S. C. B. Gopinath, Appl. Nanosci., 2024, 14, 109–122 CrossRef CAS.
  117. B. Farshid, G. Lalwani, M. S. Mohammadi, J. S. Sankaran, S. Patel, S. Judex, J. Simonsen and B. Sitharaman, J. Biomed. Mater. Res., Part A, 2019, 107, 1143–1153 CrossRef CAS PubMed.
  118. S. Sinha Ray and M. Okamoto, Prog. Polym. Sci., 2003, 28, 1539–1641 CrossRef.
  119. M. M. Rahman, K. H. Khan, M. M. H. Parvez, N. Irizarry and M. N. Uddin, Processes, 2025, 13, 994 CrossRef CAS.
  120. H. M. Shanshool, M. Yahaya, W. M. M. Yunus and I. Y. Abdullah, J. Mater. Sci.: Mater. Electron., 2016, 27, 9804–9811 CrossRef CAS.
  121. B. Davoodi, V. Goodarzi, H. Hosseini, M. Tirgar, S. Shojaei, A. Asefnejad, A. Saeidi, F. Oroojalian and S. Zamanlui, J. Polym. Res., 2022, 29, 54 CrossRef CAS.
  122. M. Moschetta, M. Chiacchiaretta, F. Cesca, I. Roy, A. Athanassiou, F. Benfenati, E. L. Papadopoulou and M. Bramini, Front. Neurosci., 2021, 15, 731198 CrossRef PubMed.
  123. W. K. Zhu, H. P. Cong, H. B. Yao, L. B. Mao, A. M. Asiri, K. A. Alamry, H. M. Marwani and S. H. Yu, Small, 2015, 11, 4298–4302 CrossRef CAS PubMed.
  124. S. P. Ogilvie, M. J. Large, M. A. O’mara, A. C. Sehnal, A. Amorim Graf, P. J. Lynch, A. J. Cass, J. P. Salvage, M. Alfonso and P. Poulin, ACS Nano, 2022, 16, 1963–1973 CrossRef CAS PubMed.
  125. T. D. Gamot, A. R. Bhattacharyya, T. Sridhar, F. Beach, R. F. Tabor and M. Majumder, Langmuir, 2017, 33, 10311–10321 CrossRef CAS PubMed.
  126. J. Luo, L. Yang, D. Sun, Z. Gao, K. Jiao and J. Zhang, Small, 2020, 16, 2003426 CrossRef CAS PubMed.
  127. N. M. Farokhi, J. M. Milani and Z. R. Amiri, J. Food Eng., 2024, 367, 111885 CrossRef CAS.
  128. Z. Chen, G. Zhang, Y. Luo and Z. Suo, Proc. Natl. Acad. Sci. U. S. A., 2024, 121, e2322684121 CrossRef CAS PubMed.
  129. M. Ghotbi, M. Pourmadadi, F. Yazdian and A. Hallajsani, Inorg. Chem. Commun., 2024, 170, 113119 CrossRef CAS.
  130. S. Ostovar, M. Pourmadadi and M. A. Zaker, Int. J. Biol. Macromol., 2023, 253, 127091 CrossRef CAS PubMed.
  131. S. De, K. Patra, D. Ghosh, K. Dutta, A. Dey, G. Sarkar, J. Maiti, A. Basu, D. Rana and D. Chattopadhyay, ACS Biomater. Sci. Eng., 2018, 4, 514–531 CrossRef CAS PubMed.
  132. K. Liang, E. M. Spiesz, D. T. Schmieden, A.-W. Xu, A. S. Meyer and M.-E. Aubin-Tam, ACS Nano, 2020, 14, 14731–14739 CrossRef CAS PubMed.
  133. A. F. Girão, G. Gonçalves, K. S. Bhangra, J. B. Phillips, J. Knowles, G. Irurueta, M. K. Singh, I. Bdkin, A. Completo and P. A. Marques, RSC Adv., 2016, 6, 49039–49051 RSC.
  134. J. Chen, B. Liu and X. Gao, Results Phys., 2020, 16, 102974 CrossRef.
  135. R. Ramachandran, D. Jung and A. M. Spokoyny, NPG Asia Mater., 2019, 11, 19 CrossRef.
  136. Y.-W. Fu, W.-F. Sun and X. Wang, Polymers, 2020, 12, 420 CrossRef CAS PubMed.
  137. S. H. Moon, H. J. Hwang, H. R. Jeon, S. J. Park, I. S. Bae and Y. J. Yang, Front. Bioeng. Biotechnol., 2023, 11, 1127757 CrossRef PubMed.
  138. M. Rahimi, V. Pirouzfar and H. Sakhaeinia, J. Polym. Environ., 2024, 32, 2762–2779 CrossRef CAS.
  139. F. Olate-Moya, G. Rubí-Sans, E. Engel, M. Á. Mateos-Timoneda and H. Palza, Biomacromolecules, 2024, 25, 3312–3324 CrossRef CAS PubMed.
  140. A. I. Raafat and A. E.-H. Ali, Polym. Bull., 2017, 74, 2045–2062 CrossRef CAS.
  141. K. Mylvaganam and L. Zhang, J. Phys. Chem. C, 2013, 117, 2817–2823 CrossRef CAS.
  142. S. Das, A. S. Wajid, J. L. Shelburne, Y.-C. Liao and M. J. Green, ACS Appl. Mater. Interfaces, 2011, 3, 1844–1851 CrossRef CAS PubMed.
  143. J. Fawaz and V. Mittal, in Synthesis Techniques for Polymer Nanocomposites, ed. V. Mittal, Wiley-VCH GmbH, Weinheim, Germany, 2014, pp. 1–30. Search PubMed.
  144. N. Basavegowda and K.-H. Baek, Polymers, 2021, 13, 4198 CrossRef CAS PubMed.
  145. N. Rabiee, M. Bagherzadeh, A. M. Ghadiri, Y. Fatahi, A. Aldhaher, P. Makvandi, R. Dinarvand, M. Jouyandeh, M. R. Saeb, M. Mozafari, M. Shokouhimehr, M. R. Hamblin and R. S. Varma, ACS Appl. Bio Mater., 2021, 4, 5336–5351 CrossRef CAS PubMed.
  146. F. J. P. Costa, M. Nave, R. Lima-Sousa, C. G. Alves, B. L. Melo, I. J. Correia and D. de Melo-Diogo, Int. J. Pharm., 2023, 635, 122713 CrossRef CAS PubMed.
  147. M. H. Aghajan, M. Panahi-Sarmad, N. Alikarami, S. Shojaei, A. Saeidi, H. A. Khonakdar, M. Shahrousvan and V. Goodarzi, Eur. Polym. J., 2020, 131, 109720 CrossRef CAS.
  148. Z. Wang, F. Cui, Y. Sui and J. Yan, Beilstein J. Org. Chem., 2023, 19, 1580–1603 CrossRef CAS PubMed.
  149. I. S. Tsagkalias, T. K. Manios and D. S. Achilias, Polymers, 2017, 9, 432 CrossRef PubMed.
  150. M. U. A. Khan, S. I. A. Razak, S. Rehman, A. Hasan, S. Qureshi and G. M. Stojanović, Int. J. Biol. Macromol., 2022, 222, 462–472 CrossRef CAS PubMed.
  151. M. U. Aslam Khan, W. S. Al-Arjan, M. S. Binkadem, H. Mehboob, A. Haider, M. A. Raza, S. I. Abd Razak, A. Hasan and R. Amin, Nanomaterials, 2021, 11, 1319 CrossRef PubMed.
  152. A. Ribas-Massonis, M. Cicujano, J. Duran, E. Besalú and A. Poater, Polymers, 2022, 14, 2856 CrossRef CAS PubMed.
  153. A. Kausar and I. Ahmad, J. Compos. Sci., 2023, 7, 290 CrossRef CAS.
  154. M. Toriello, M. Afsari, H. K. Shon and L. D. Tijing, Membranes, 2020, 10, 204 CrossRef CAS PubMed.
  155. A. Al-Abduljabbar and I. Farooq, Polymers, 2023, 15, 65 CrossRef CAS PubMed.
  156. R. Nayak, R. Padhye, I. L. Kyratzis, Y. B. Truong and L. Arnold, Text. Res. J., 2013, 83, 606–617 CrossRef.
  157. Y. Sun, S. Cheng, W. Lu, Y. Wang, P. Zhang and Q. Yao, RSC Adv., 2019, 9, 25712–25729 RSC.
  158. M. Rezaei, M. Nikkhah, S. Mohammadi, S. H. Bahrami and M. Sadeghizadeh, J. Appl. Polym. Sci., 2021, 138, 50884 CrossRef CAS.
  159. F. Norouzi, M. Pourmadadi, F. Yazdian, K. Khoshmaram, J. Mohammadnejad, M. H. Sanati, F. Chogan, A. Rahdar and F. Baino, J. Funct. Biomater., 2022, 13, 300 CrossRef CAS PubMed.
  160. R. Jaswal, D. Kumar, A. I. Rezk, V. K. Kaliannagounder, C. H. Park and K. H. Min, Colloids Surf., B, 2024, 237, 113820 CrossRef CAS PubMed.
  161. A. Keirouz, Z. Wang, V. S. Reddy, Z. K. Nagy, P. Vass, M. Buzgo, S. Ramakrishna and N. Radacsi, Adv. Mater. Technol., 2023, 8, 2201723 CrossRef CAS.
  162. N. Ren, A. Qiao, M. Cui, R. Huang, W. Qi and R. Su, Chem. Eng. Sci., 2023, 282, 119320 CrossRef CAS.
  163. M. A. Al Faruque, R. Remadevi, A. Guirguis, A. Kiziltas, D. Mielewski and M. Naebe, Sci. Rep., 2021, 11, 12068 CrossRef CAS PubMed.
  164. H. He, L. Guan and H. Le Ferrand, J. Mater. Chem. A, 2022, 10, 19129–19168 RSC.
  165. M. Oksuz and H. Y. Erbil, RSC Adv., 2018, 8, 17443–17452 RSC.
  166. S. Talebian, M. Mehrali, R. Raad, F. Safaei, J. Xi, Z. Liu and J. Foroughi, Front. Chem., 2020, 8, 88 CrossRef CAS PubMed.
  167. C. Katrilaka, N. Karipidou, N. Petrou, C. Manglaris, G. Katrilakas, A. N. Tzavellas, M. Pitou, E. E. Tsiridis, T. Choli-Papadopoulou and A. Aggeli, Materials, 2023, 16, 4425 CrossRef CAS PubMed.
  168. C. Valencia, C. H. Valencia, F. Zuluaga, M. E. Valencia, J. H. Mina and C. D. Grande-Tovar, Molecules, 2018, 23, 2651 CrossRef PubMed.
  169. S. Liu, C. Zhou, S. Mou, J. Li, M. Zhou, Y. Zeng, C. Luo, J. Sun, Z. Wang and W. Xu, Mater. Sci. Eng., C, 2019, 105, 110137 CrossRef CAS PubMed.
  170. A.-V. Do, B. Khorsand, S. M. Geary and A. K. Salem, Adv. Healthcare Mater., 2015, 4, 1742–1762 CrossRef CAS PubMed.
  171. A. Seyedsalehi, L. Daneshmandi, M. Barajaa, J. Riordan and C. T. Laurencin, Sci. Rep., 2020, 10, 22210 CrossRef CAS PubMed.
  172. J. M. Seok, G. Choe, S. J. Lee, M.-A. Yoon, K.-S. Kim, J. H. Lee, W. D. Kim, J. Y. Lee, K. Lee and S. A. Park, Mater. Des., 2021, 209, 109941 CrossRef CAS.
  173. M. Mashhadi Keshtiban, H. Taghvaei, R. Noroozi, V. Eskandari, Z. U. Arif, M. Bodaghi, H. Bardania and A. Hadi, Adv. Eng. Mater., 2024, 26, 2301260 CrossRef CAS.
  174. H. Belaid, S. Nagarajan, C. Teyssier, C. Barou, J. Barés, S. Balme, H. Garay, V. Huon, D. Cornu, V. Cavaillès and M. Bechelany, Mater. Sci. Eng., C, 2020, 110, 110595 CrossRef CAS PubMed.
  175. S. Agarwal, S. Saha, V. K. Balla, A. Pal, A. Barui and S. Bodhak, Frontiers in Mechanical Engineering, 2020, 6, 589171 CrossRef.
  176. T. d. P. L. Lima, C. A. d. A. Canelas, V. O. C. Concha, F. A. M. d. Costa and M. F. Passos, J. Funct. Biomater., 2022, 13, 214 CrossRef CAS PubMed.
  177. A. Mousavi, A. Hedayatnia, P. P. van Vliet, D. R. Dartora, N. Wong, N. Rafatian, A. M. Nuyt, C. Moraes, A. Ajji, G. Andelfinger and H. Savoji, Appl. Mater. Today, 2024, 36, 102035 CrossRef.
  178. S. Sharma, H. Sharma and R. Sharma, Chemistry of Inorganic Materials, 2024, 2, 100035 CrossRef.
  179. C. Comanescu, Coatings, 2023, 13, 1772 CrossRef CAS.
  180. N. Işıklan, G. Geyik and E. Güncüm, Mater. Today Chem., 2024, 41, 102323 CrossRef.
  181. R. Justin and B. Chen, Interface Focus, 2018, 8, 20170055 CrossRef PubMed.
  182. G. F. Stiufiuc and R. I. Stiufiuc, Appl. Sci., 2024, 14, 1623 CrossRef CAS.
  183. Q. Zhang, Z. Wu, N. Li, Y. Pu, B. Wang, T. Zhang and J. Tao, Mater. Sci. Eng., C, 2017, 77, 1363–1375 CrossRef CAS PubMed.
  184. E. Khakpour, S. Salehi, S. M. Naghib, S. Ghorbanzadeh and W. Zhang, Front. Bioeng. Biotechnol., 2023, 11, 1129768 CrossRef PubMed.
  185. J. Singh and P. Nayak, J. Polym. Sci., 2023, 61, 2828–2850 CrossRef CAS.
  186. E. Daneshmoghanlou, M. Miralinaghi, E. Moniri and S. K. Sadjady, J. Polym. Environ., 2022, 30, 3718–3736 CrossRef CAS.
  187. M. U. A. Khan, Z. Yaqoob, M. N. M. Ansari, S. I. A. Razak, M. A. Raza, A. Sajjad, S. Haider and F. M. Busra, Polymers, 2021, 13, 3124 CrossRef CAS PubMed.
  188. J. Siepmann, K. Elkharraz, F. Siepmann and D. Klose, Biomacromolecules, 2005, 6, 2312–2319 CrossRef CAS PubMed.
  189. S. Li and S. McCarthy, Biomaterials, 1999, 20, 35–44 CrossRef CAS PubMed.
  190. Y. Liu, S. Longqing, J. Yuxuan, J. Shaojing, Z. Xiaokang, L. Song, C. Jing and J. Hu, Int. J. Nanomed., 2025, 20, 705–721 CrossRef CAS PubMed.
  191. P. Misra, B. K. Mishra and G. B. Behera, Int. J. Chem. Kinet., 1991, 23, 639–654 CrossRef CAS.
  192. A. C. Dash, B. Dash and S. Praharaj, J. Chem. Soc., Dalton Trans., 1981,(10), 2063–2069 RSC.
  193. M. Karimi, P. Sahandi Zangabad, A. Ghasemi, M. Amiri, M. Bahrami, H. Malekzad, H. Ghahramanzadeh Asl, Z. Mahdieh, M. Bozorgomid, A. Ghasemi, M. R. Rahmani Taji Boyuk and M. R. Hamblin, ACS Appl. Mater. Interfaces, 2016, 8, 21107–21133 CrossRef CAS PubMed.
  194. M. Bikram and J. L. West, Expet Opin. Drug Deliv., 2008, 5, 1077–1091 CrossRef CAS PubMed.
  195. D. N. Céspedes-Valenzuela, S. Sánchez-Rentería, J. Cifuentes, S. C. Gómez, J. A. Serna, L. Rueda-Gensini, C. Ostos, C. Muñoz-Camargo and J. C. Cruz, Front. Bioeng. Biotechnol., 2022, 10, 947616 CrossRef PubMed.
  196. Q. Hu, P. S. Katti and Z. Gu, Nanoscale, 2014, 6, 12273–12286 RSC.
  197. M. Li, G. Zhao, W.-K. Su and Q. Shuai, Front. Chem., 2020, 8, 647 CrossRef CAS PubMed.
  198. R. L. Minehan and M. P. Del Borgo, Frontiers in Biomaterials Science, 2022, 1, 916985 CrossRef.
  199. J. Liu, W. Kang and W. Wang, Photochem. Photobiol., 2022, 98, 288–302 CrossRef CAS PubMed.
  200. Y. Xing, B. Zeng and W. Yang, Front. Bioeng. Biotechnol., 2022, 10, 1075670 CrossRef PubMed.
  201. R. J. Mitchell, D. Havrylyuk, A. C. Hachey, D. K. Heidary and E. C. Glazer, Chem. Sci., 2025, 16, 721–734 RSC.
  202. M. Fernández and J. Orozco, Polymers, 2021, 13, 2464 CrossRef PubMed.
  203. C. L. Weaver, J. M. LaRosa, X. Luo and X. T. Cui, ACS Nano, 2014, 8, 1834–1843 CrossRef CAS PubMed.
  204. D. Zahn, A. Weidner, K. Saatchi, U. O. Häfeli and S. Dutz, Current Directions in Biomedical Engineering, 2019, 5, 161–164 CrossRef.
  205. Z. Wang, C. Liu, B. Chen and Y. Luo, Int. J. Biol. Macromol., 2021, 168, 38–45 CrossRef CAS PubMed.
  206. F. V. Lavrentev, V. V. Shilovskikh, V. S. Alabusheva, V. Y. Yurova, A. A. Nikitina, S. A. Ulasevich and E. V. Skorb, Molecules, 2023, 28, 5931 CrossRef CAS PubMed.
  207. M. S. B. Reddy, D. Ponnamma, R. Choudhary and K. K. Sadasivuni, Polymers, 2021, 13, 1105 CrossRef CAS PubMed.
  208. M. V. S. Varma, A. M. Kaushal, A. Garg and S. Garg, Am. J. Drug Delivery, 2004, 2, 43–57 CrossRef CAS.
  209. D. Hazarika, K. Gupta, M. Mandal and N. Karak, ACS Omega, 2018, 3, 2292–2303 CrossRef CAS PubMed.
  210. V. J. Sawant, B. V. Tawade, V. M. Desai, B. B. Dongare and S. V. Nipane, Prog. Biomater., 2022, 11, 193–205 CrossRef CAS PubMed.
  211. T. M. Joseph, A. B. Unni, K. S. Joshy, D. Kar Mahapatra, J. Haponiuk and S. Thomas, C-J. Carbon Res., 2023, 9, 30 CrossRef CAS.
  212. L. Moradi, L. Witek, V. Vivekanand Nayak, A. Cabrera Pereira, E. Kim, J. Good and C. J. Liu, Biomaterials, 2023, 301, 122289 CrossRef CAS PubMed.
  213. N. Hossain, M. A. Chowdhury, M. Kchaou, A. Alam and M. M. Rahman, Chem. Phys. Impact, 2023, 6, 100180 CrossRef.
  214. E. Gudiño, C. M. Oishi and A. Sequeira, Int. J. Numer. Methods Eng., 2018, 114, 292–320 CrossRef.
  215. A. Deb, N. G. Andrews and V. Raghavan, Int. J. Biol. Macromol., 2018, 113, 515–525 CrossRef CAS PubMed.
  216. I. A. A. Ibrahim, A. R. Alzahrani, I. M. Alanazi, N. Shahzad, I. Shahid, A. H. Falemban, M. F. N. Azlina and P. Arulselvan, Int. J. Biol. Macromol., 2023, 253, 127334 CrossRef CAS PubMed.
  217. G. Singh, B. P. Nenavathu, K. Imtiyaz and M. Moshahid A Rizvi, Biomed. Pharmacother., 2020, 129, 110443 CrossRef CAS PubMed.
  218. S. Nazir, M. Umar Aslam Khan, W. Shamsan Al-Arjan, S. Izwan Abd Razak, A. Javed and M. Rafiq Abdul Kadir, Arab. J. Chem., 2021, 14, 103120 CrossRef CAS.
  219. I. Sahu and P. Chakraborty, Colloids Surf., B, 2024, 233, 113654 CrossRef CAS PubMed.
  220. S. Xiang, C. Guilbaud-Chéreau, P. Hoschtettler, L. Stefan, A. Bianco and C. Ménard-Moyon, Int. J. Biol. Macromol., 2024, 255, 127919 CrossRef CAS PubMed.
  221. S. Javanbakht and A. Shaabani, Int. J. Biol. Macromol., 2019, 123, 389–397 CrossRef CAS PubMed.
  222. B. N. Kumara, R. Shambhu, A. Prabhu and K. S. Prasad, Carbohydr. Polym., 2022, 289, 119426 CrossRef CAS PubMed.
  223. Z. Karimzadeh and H. Namazi, J. Polym. Res., 2022, 29, 37 CrossRef CAS.
  224. S. R. Obireddy, A. Ayyakannu, G. Kasi, B. Sridharan, W.-F. Lai and K. Viswanathan, Int. J. Nanomed., 2025, 20, 1965–1981 CrossRef CAS PubMed.
  225. S. Dhanavel, P. Praveena, V. Narayanan and A. Stephen, Polym. Bull., 2020, 77, 5681–5696 CrossRef CAS.
  226. A. D. Sontakke, P. Gupta, S. K. Banerjee and M. K. Purkait, Int. J. Biol. Macromol., 2024, 271, 132621 CrossRef CAS PubMed.
  227. A. Hardiansyah, A. Randy, R. T. Dewi, M. Angelina, N. Yudasari, S. Rahayu, I. M. Ulfah, F. Maryani, Y.-W. Cheng and T.-Y. Liu, Polymers, 2022, 14, 3163 CrossRef CAS PubMed.
  228. S. Agila and J. Poornima, in 2015 IEEE 15th International Conference on Nanotechnology (IEEE-NANO), IEEE, 2015, pp. 1058–1061 Search PubMed.
  229. N. S. Tehrani, M. Masoumi, F. Chekin and M. S. Baei, Russ. J. Appl. Chem., 2020, 93, 1221–1228 CrossRef CAS.
  230. M. Krishani, W. Y. Shin, H. Suhaimi and N. S. Sambudi, Gels, 2023, 9, 100 CrossRef CAS PubMed.
  231. M. I. Echeverria Molina, K. G. Malollari and K. Komvopoulos, Front. Bioeng. Biotechnol., 2021, 9, 617141 CrossRef PubMed.
  232. P. Zarrintaj, F. Seidi, M. Y. Azarfam, M. K. Yazdi, A. Erfani, M. Barani, N. P. S. Chauhan, N. Rabiee, T. Kuang and J. Kucinska-Lipka, Composites, Part B, 2023, 258, 110701 CrossRef CAS.
  233. M. Z. A. Mahmud, M. D. Islam and M. H. Mobarak, J. Nanomater., 2023, 2023, 9270064 Search PubMed.
  234. S. D. Purohit, R. Bhaskar, H. Singh, I. Yadav, M. K. Gupta and N. C. Mishra, Int. J. Biol. Macromol., 2019, 133, 592–602 CrossRef CAS PubMed.
  235. S. D. Purohit, H. Singh, R. Bhaskar, I. Yadav, S. Bhushan, M. K. Gupta, A. Kumar and N. C. Mishra, Front. Mater., 2020, 7, 250 CrossRef.
  236. W. T. Godbey and A. Atala, Ann. N. Y. Acad. Sci., 2002, 961, 10–26 CrossRef CAS PubMed.
  237. N. A. Kurniawan, Curr. Opin. Organ Transplant., 2019, 24, 590–597 CrossRef CAS PubMed.
  238. S. Caddeo, M. Boffito and S. Sartori, Front. Bioeng. Biotechnol., 2017, 5, 40 CrossRef PubMed.
  239. C. I. Chamorro, S. Zeiai, N. Juul, O. Willacy, J. Huo, J. Hilborn and M. Fossum, Int. J. Mol. Sci., 2022, 23, 12703 CrossRef CAS PubMed.
  240. W. R. Dong, Y. Q. Xiao, Y. J. Piao and Y. H. Chen, Acad. J. First Med. Coll. PLA, 2004, 24, 969–974 CAS.
  241. F. J. O'Brien, Mater. Today, 2011, 14, 88–95 CrossRef.
  242. Y. Hou, W. Wang and P. Bártolo, Int. J. Bioprint., 2020, 6, 266 CrossRef CAS PubMed.
  243. C. Shuai, B. Peng, P. Feng, L. Yu, R. Lai and A. Min, J. Adv. Res., 2022, 35, 13–24 CrossRef CAS PubMed.
  244. N. Mahmoudi and A. Simchi, Mater. Sci. Eng., C, 2017, 70, 121–131 CrossRef CAS PubMed.
  245. M. S. Ramasamy, R. Bhaskar, K. B. Narayanan, S. D. Purohit, S. S. Park, A. Manikkavel, B. Kim and S. S. Han, Mater. Today Commun., 2022, 33, 104659 CrossRef.
  246. F. Puza, S. Rostami, B. Özçolak-Aslan, S. Odabaş, K. D. Jandt and B. Garipcan, Adv. Eng. Mater., 2023, 25, 2200777 CrossRef CAS.
  247. S. Kumar, S. Raj, S. Jain and K. Chatterjee, Mater. Des., 2016, 108, 319–332 CrossRef CAS.
  248. C. D. Grande Tovar, J. I. Castro, C. H. Valencia, P. A. Zapata, M. A. Solano, E. Florez López, M. N. Chaur, M. E. Valencia Zapata and J. H. Mina Hernandez, Molecules, 2020, 25, 2308 CrossRef PubMed.
  249. C. H. Valencia-Llano, M. A. Solano and C. D. Grande-Tovar, Polymers, 2021, 13, 3877 CrossRef CAS PubMed.
  250. J. T. Rashkow, Y. Talukdar, G. Lalwani and B. Sitharaman, ACS Biomater. Sci. Eng., 2017, 3, 2533–2541 CrossRef CAS PubMed.
  251. G. L. Breitenbach, B. S. Caldas, M. C. G. Pellá, E. C. Muniz and D. C. Dragunski, Colloids Surf., A, 2024, 701, 134872 CrossRef CAS.
  252. C. Fu, H. Bai, Q. Hu, T. Gao and Y. Bai, RSC Adv., 2017, 7, 8886–8897 RSC.
  253. S. Nagarajan, C. Pochat-Bohatier, C. Teyssier, S. Balme, P. Miele, N. Kalkura, V. Cavaillès and M. Bechelany, RSC Adv., 2016, 6, 109150–109156 RSC.
  254. G. Khamidov, Ö. Hazman, M. N. Hasanovich and I. Erol, J. Drug Delivery Sci. Technol., 2024, 91, 105258 CrossRef CAS.
  255. M. U. A. Khan, S. I. A. Razak, M. N. M. Ansari, R. M. Zulkifli, N. Ahmad Zawawi and M. Arshad, Polymers, 2021, 13, 3611 CrossRef CAS PubMed.
  256. G. Lalwani, A. M. Henslee, B. Farshid, L. Lin, F. K. Kasper, Y.-X. Qin, A. G. Mikos and B. Sitharaman, Biomacromolecules, 2013, 14, 900–909 CrossRef CAS PubMed.
  257. B. Farshid, G. Lalwani and B. Sitharaman, J. Biomed. Mater. Res., Part A, 2015, 103, 2309–2321 CrossRef CAS PubMed.
  258. J. Yang, J. Zhang, M. Yusoff, N. A. Roslan and M. H. Razali, Dig. J. Nanomater. Biostruct., 2025, 20, 301–314 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.