Next Article in Journal
Cu(C3H3N3S3)3 Adsorption onto ZnTiO3/TiO2 for Coordination-Complex Sensitized Photochemical Applications
Next Article in Special Issue
Rubus ellipticus Sm. Fruit Extract Mediated Zinc Oxide Nanoparticles: A Green Approach for Dye Degradation and Biomedical Applications
Previous Article in Journal
Physicochemical Characterization of Novel Epoxidized Vegetable Oil from Chia Seed Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Surface Engineering of Nanomaterials with Polymers, Biomolecules, and Small Ligands for Nanomedicine

by
Ana M. Díez-Pascual
Universidad de Alcalá, Facultad de Ciencias, Departamento de Química Analítica, Química Física e Ingeniería Química, Ctra. Madrid-Barcelona, Km. 33.6, 28805 Alcalá de Henares, Madrid, Spain
Materials 2022, 15(9), 3251; https://doi.org/10.3390/ma15093251
Submission received: 18 March 2022 / Revised: 26 April 2022 / Accepted: 28 April 2022 / Published: 30 April 2022

Abstract

:
Nanomedicine is a speedily growing area of medical research that is focused on developing nanomaterials for the prevention, diagnosis, and treatment of diseases. Nanomaterials with unique physicochemical properties have recently attracted a lot of attention since they offer a lot of potential in biomedical research. Novel generations of engineered nanostructures, also known as designed and functionalized nanomaterials, have opened up new possibilities in the applications of biomedical approaches such as biological imaging, biomolecular sensing, medical devices, drug delivery, and therapy. Polymers, natural biomolecules, or synthetic ligands can interact physically or chemically with nanomaterials to functionalize them for targeted uses. This paper reviews current research in nanotechnology, with a focus on nanomaterial functionalization for medical applications. Firstly, a brief overview of the different types of nanomaterials and the strategies for their surface functionalization is offered. Secondly, different types of functionalized nanomaterials are reviewed. Then, their potential cytotoxicity and cost-effectiveness are discussed. Finally, their use in diverse fields is examined in detail, including cancer treatment, tissue engineering, drug/gene delivery, and medical implants.

Graphical Abstract

1. Introduction

Nanomaterials with unique physicochemical properties have recently attracted a lot of attention since they offer a lot of potential in many fields, particularly in biomedical sciences, including drug and gene delivery systems [1,2,3], cancer treatment [4,5], monitoring systems [6], tissue engineering [7], and so forth. New generations of engineered nanostructures, also known as designed and functionalized nanomaterials, have opened up new possibilities in the applications of biomedical approaches such as biomolecular sensing, drug delivery, biological imaging, and therapy. A wide number of nanomaterials have great potential to be used in biomedicine, including nanotubes, nanoparticles, nanoplates, and nanowires, to mention but a few [8,9,10]. Nonetheless, they must meet specific characteristics to be used in biomedical applications [11]. In this regard, their potential cytotoxicity, which can be induced by their structure, chemical content, or features, for example, as well as their biocompatibility, have to be assessed [12]. Their colloidal stability should also be maintained under physiological conditions, ideally across a wide pH range [13]. As a result, it is critical to consider these criteria to ensure the safety, nontoxicity, and biocompatibility of the nanomaterials. Specific interactions with polymers, natural biomolecules, and synthetic ligands of interest are required to modify and functionalize the nanomaterial surface in order to meet these criteria [14,15,16].
The methods for creating and manipulating functionalized nanomaterials (FNMs) open up exciting new opportunities for developing novel multifunctional biological devices [17]. Furthermore, functionalization prevents nanoparticles from agglomeration and makes them compatible in subsequent phases. As a result, FNMs can transport more efficiently after systemic injection and have better pharmacokinetic characteristics in vivo. FNMs can be deeply driven into tissues through narrow capillaries and epithelial coating, leading to improved therapeutic agent delivery to the targeted location [18]. Furthermore, the small size of FNMs enhances exceptional physicochemical features such as solubility, diffusivity, immunogenicity, and the capacity to target the designated region with minimum diffusion to its surrounding [19,20].
The nanomaterial interface can be designed and applied in different ways. These approaches are classified as replacement, noncovalent, and covalent conjugations based on the primary concept of the type of functionalization interaction [21]. The interface between nanoparticles (NPs) and the attached molecules is modified via the replacement approach, which comprises ligand exchange and ligand addition [22]. Noncovalent techniques rely on many interactions, most of them weak, such as electrostatic, Van der Waals, hydrophobic, and hydrogen bonds, and it is particularly useful with metallic nanoparticles [23]. They are straightforward and do not modify the molecular structure nor their interaction with targets. However, these modifications are strongly dependent on parameters such as ionic strength and pH. On the other hand, covalent attachment techniques have been proposed to alter the external functionalization of nanomaterials to bind molecular entities for biomedical purposes, hence giving the nanoparticles additional functionality [24].
This study aims to provide particular examples to cover the different ways of nanomaterial functionalization using polymers, natural biomolecules, and small ligands (Figure 1), via covalent and noncovalent conjugation. Before highlighting specific examples of each type of functionalization, the basis of the functionalization will be summarized. Although some studies on nanoparticle surface modification for medical and nanotechnological application have been reported [8,10,11,12,25], most of them are not updated, deal only with nanoparticles rather than nanomaterials in general, and focus only on either the nanoparticle synthesis or on certain biomedical applications. Thus, the current paper reviews recent studies on nanomaterial surface engineering, divided by nanomaterial type and specialized uses. Besides, the cytotoxicity, cost effectiveness, and use of FNMs as a versatile tool in nanomedicine will be discussed. Due to their beneficial characteristics such as biodegradability and biocompatibility in physiological mechanisms, wide availability, suitability for chemical treatment, and wide range of potential synthesis process from different sources, nanomaterials have been extensively explored in the literature. This article offers novel insights on surface functionalization of nanomaterials, focusing on their therapeutic, diagnostic, tissue-engineering, and medical-implant applications. Following a brief overview of the different surface modification strategies and different types of functionalized nanomaterials, a summary of the most relevant biomedical applications is presented.

2. Strategies for Surface Functionalization of Nanomaterials

The exclusive properties of nanomaterials compared to their microsized counterparts, such as big, specific surface areas and nanometer sizes, have involved huge attention in the scientific community. Depending on the desired final properties, the composition of nanomaterials can vary from metals or metal oxides to carbon or polymers (Figure 2). Metallic nanoparticles (like gold or silver) are beneficial for designing drug delivery and imaging systems, but their safety has to be investigated in detail to prevent undesirable side effects in humans [26]. Iron oxide, with outstanding magnetic properties, is the most common selection as the core of functionalized nanoparticles. Silica NPs are frequently used in drug delivery applications. Mesoporous silica nanoparticles (MSN), with tunable pore size, are widely used to load small molecules, including amino acids, nucleic acids, drugs, and so forth [27]. However, due to reactive surface silanol groups, there are biocompatibility issues regarding the use of silica nanoparticles for nanomedicine. Concerns regarding the toxicity of carbon-based nanomaterials such as carbon nanotubes (CNTs), quantum dots (QDs), and graphene have also been reported [28]. Polymeric nanoparticles are a widespread option for biomedical applications owed to their tailorable physicochemical properties, excellent biocompatibility, and capability to liberate molecules in a continued way [29]. Numerous polymeric micelles such as thiomers, pluronic, polysaccharides, and polyethylene glycol (PEG) have been investigated [14]. In addition, other colloidal nanostructures such as dendrimers, liposomes, polysomes, and cyclodextrins have been designed for targeted applications [30].
Various types of targeting agents have been implemented to be incorporated on the surface of nanoparticles, especially peptides [31], aptamers [32], antibodies [33], polyethylene glycol (PEG) [34], cationic molecules, folic acid [35], drugs, and fluorescent probes, as depicted in Figure 3. It should be noted that biomolecular interactions rely on the chemical modification of the nanoparticle surface when using NPs for in vitro or in vivo applications [36]. Through a ligand–receptor interaction, such targeting moieties can allow nanoparticles to be embodied into cancer cells and tissues. To facilitate active targeting of NPs to receptors, which are located on the surface of the membrane, the nanoparticle surface can be tailored with targeting ligands, resulting in increased cellular internalization and/or selective absorption via receptor-mediated endocytosis [37]. Researchers are particularly interested in discovering new biomarkers and their relevant ligands in targeted medication administration. The binding of NPs to analytes, pathogens, and biomarkers might cause their signal to be amplified, making it easier to detect and image [38]. When the scaffold surface is decorated with bioactive cues to allow FNPs to interact with cells and the extracellular matrix (ECM) to elicit tissue-specific phenotypes, this is referred to as functionalization [39]. Chemists can easily make the suitable functionalities for use in clinics thanks to the easiness of such functionalization. For example, cell surface molecules have been used to identify nanoparticles functionalized with ligands that show a varied affinity for proteins [34].
Furthermore, functionalization has been proven to protect NPs against agglomeration and make them biocompatible materials in other application stages [41]. Functionalization improves the NPs’ physical, chemical, and mechanical characteristics, resulting in synergetic effects [42].

2.1. Functionalization by Covalent Conjugation

The covalent conjugation comprises the reaction of a conjugator (also named linker) with a certain species or chemical group, in a way that the molecules are attached on the nanomaterial surface [43]. Carboxylic acids, amines, thiols, disulfides, phosphates, nitriles, and so forth have been used for covalent conjugation via chemical reactions [44]. Amine groups are the most widely used for functionalization in the biomedical field. The strategy consists in anchoring small molecules or proteins on the nanoparticles. Further, amine functionalization can be used with the aid of n-hydroxysuccinimide (NHS) and different carbodiimides such as EDC. Similarly, carboxylic groups can form ester or amide bonds with alcohol or amine groups on the NPs’ surface [45], Figure 4. On the other hand, conjugation on metallic NPs can be effectively carried out via the thiol moiety. The interaction occurs by reaction of sulfhydryl (RSH) groups on the metallic nanoparticles.

2.2. Functionalization by Noncovalent Conjugation

The noncovalent bonding comprises the attachment of molecules on the surface of nanomaterials without chemical bonding via physical adsorption and/or wrapping of molecules by weak interactions such as hydrophobic (Van der Waals), H-bonding, cation−π, anion−π, π–π, and H−π, that preserve the intrinsic properties of the nanomaterial [43]. This approach has some benefits over the covalent way: (i) it takes place under moderate conditions (water solution at room temperature), thus avoiding structural damage of the nanomaterial; and (ii) enables control of the amount of adsorbed/wrapped molecule. The versatility of this route enables a large number of substances to be coupled to the nanomaterials including polymers, solvents, surfactants, aromatic compounds, etc. In order to offer steric stabilization, nanomaterials have been anchored to biocompatible polymers such as polyethylene glycol (PEG) [46].

2.3. Functionalization by Biomolecules

Biomolecules are outstanding candidates to apply in the surface engineering of nanoparticles. Biomolecule-coated nanoparticles have features that are troublesome or inconceivable to attain with synthetic materials, such as excellent bio-macromolecule distribution with little cytotoxicity. Biomolecules such as proteins, peptides, antibodies, and oligonucleotides can be very valuable for targeting NPs to cancer cells where particular receptors are overexpressed. The synthesis of gold–thiol bonds to create oligonucleotide–AuNP conjugates was one of the first bio-nanotechnology examples reported in the literature [34]. Proteins or peptides boost the penetration of NPs into cells via receptor-mediated endocytosis. On the other hand, transferrin is a glycoprotein that can bind to specific receptors on the cell membrane. A few articles have reported the benefits of using this protein as a target for Au, MSN, and poly(lactic-co-glycolic acid) (PLGA) nanoparticles [30].
Albumins are a class of naturally occurring proteins that, besides being applied to load imaging and therapeutic agents, are valuable for modification of numerous types of NPs, as depicted in Figure 5 [47]. Surface modification of NPs with albumins, such as bovine serum albumin (BSA), provides higher water solubility, increased biocompatibility and blood circulation time, and improved stability and cellular interactions compared with uncoated nanoparticles. Different strategies for conjugation of NPs such as AuNPs with albumin have been reported [48], including: (1) Passive adsorption, so that the charged groups of the protein are anchored to the NP surface via covalent or noncovalent interactions. (2) Active adsorption, which involves the use of modified albumin in order to strengthen the albumin-NP interactions. (3) The use of this protein for NP synthesis, either as a reagent (i.e., reducing agent), foaming, stabilizer, or building block for NP synthesis, resulting in NPs with an albumin coating [49]. The use of albumin encapsulation methods provides some profits, such as the loading of agents with low solubility in order to protect them from degradation. (4) Desolvation cross-linking (coacervation process), used to produce core–shell albumin-NPs. This strategy allows chemical agents to become trapped within albumin capsules, which are very stable and protect from degradation. (5) Emulsification: an albumin solution and a nonaqueous phase are mixed, giving rise to an emulsion, and the NP is dissolved in the oil phase. This methodology is used for the encapsulation of lipophilic drugs and enhances aqueous solubility and biocompatibility. (6) Thermal gelation: an albumin water solution is heated to induce protein unfolding, which results in protein–protein interactions by disulfide and hydrogen bonding, as well as hydrophobic and electrostatic forces. Besides, unfolding induces NP–protein interactions, leading to a protein coating onto the NP surface.

2.4. Functionalization by Polymers

A large number of biocompatible, commercially accessible polymers can be used for functionalization, and are typically chosen based on their specific properties such as hydrophobicity, melting point, and functional groups. Polymers frequently used as NP coatings comprise synthetic polymers (i.e., PEG [50] and PLGA [51]) and natural polymers (such as chitosan [52,53]). Polymers have been used for both covalent and noncovalent conjugation of a wide range of nanomaterials. The covalent approach involves the “grafting” (chemical anchoring) of polymeric segments to the NM surface, and can be implemented via “grafting to”, “grafting from”, “grafing through” and “in situ” tactics (Figure 6). The former is based on the synthesis of a modified polymer prone to react with the functional groups on the surface of the nanomaterial [46]. A shortcoming of this tactic is that the amount of polymer grafted to the nanomaterial is restricted, owed to the low reactivity and large steric barrier of the polymeric segments. In the “grafting from” path the polymer is grown from the NM surface via polymerization of monomers [43]. This approach is effective and manageable, owed to the high reactivity of monomers, allowing a high grafting level. A variation of this strategy is to carry it out via “in situ” polymerization in the presence of the inorganic precursor. Nonetheless, this method requests precise monitor of the amounts of each reagent and the polymerization conditions. In the “grafting through”, a low molecular weight monomer is radically copolymerized with a polymerizable macromonomer in the presence of an initiator.
Polymers are suitable for functionalization because they create a physical barrier around the NPs, preventing the core of the NPs from coming into direct contact with biological receptors. Polymers can produce a physical barrier but with a reduced hydrodynamic radius. As a result, polymer coatings outperform small molecule ligands when imparting macromolecular system characteristics to the particle surface, similar to biological proteins. The use of polymers such as PEG to coat nanoparticles improves passive tumor tissue targeting, increasing permeability, and retention (EPR), as well as biocompatibility [54]. This PEG and other polymer coatings decrease blood serum protein adsorption, lengthen circulation duration, and promote particle absorption into tumor tissues [34]. Using AuNPs synthesized by stacking cationic polyallylamine and anionic poly (acrylic acid) polyelectrolyte layers, Kleinfeldt and coworkers [55] developed an excessively hydrophilic and biocompatible coating that enables colloidal stability. Makvandi et al. [56] investigated the functionalization of various polymers (glyclusters, glydendrimers, glycopolymers) and nanomaterials (Ag2O, CuO, ZnO, Fe3O4, MgO, TiO2, Se, Ni, Pd) for water purification, food containers, fabrics, and medical applications. The benefits and drawbacks of polymer functionalization were investigated and explored in that study. When natural or synthetic polymers are used to functionalize NPs, photo/thermo-responsive properties can be achieved [57]. For instance, chitosan grafted with poly-L-lactide using thiourea-functionalized, and poly-N-isopropyl acrylamide were used to synthesize photo/thermo-triggered micelles [58].

2.5. Functionalization by Small Ligands

Small ligands are a common selection for functionalizing nanomaterial since they are relatively simple to chemically bond to surfaces via functional moieties in their structure. They are an appropriate choice to adjust the nanomaterial properties such as hydrophilicity or charge with a view to improve their biological activity and interaction with other biologically essential ligands, as well as their stability, aqueous solubility, drug loading, and so forth. For instance, silica NPs can be straightforwardly tailored with organosilane molecules such as 3-(aminopropyl) triethoxysilane (APTES) through silane chemistry. It has been reported that APTES-functionalization is an effective method for adjusting drug loading and discharge from mesoporous silica nanoparticles (MSN) [59]. Besides, it is beneficial for many aims, such as the release of low soluble drugs, the targeting of drugs to a chosen position, or to make multifunctional drug delivery and imaging devices. Other ligands such as drugs have been used for tailoring NP surfaces. For example, doxorubicin (DOX, a frequently applied anticancer drug) has been conjugated to Fe3O4 NPs with the aim to develop dual-functional NPs [60]. These modified NPs can destroy tumor cells via the conjugated DOX, and concurrently enable magnetic resonance imaging of the tumor, which is highly valuable. Other small drugs such as methotrexate, that can target the folate receptor on cancer cells, ciprofloxacin [25,61], and so forth, have been conjugated with different nanomaterials.
Various nanomaterials functionalized by small ligands can be added as signal reporters or as carriers for loading more signal reporters in biosensors for analyte detection [62]. Mahmoudpour et al. [63] designed a method for producing aptameric functionalized materials (AFMs). Optical indicators, conducting transducers, carriers, catalysts, and other features, were combined to develop advanced AFMs. Drug delivery, bioimaging, and appropriate sensing have been highlighted as biological uses of improved AFMs. Aptamers have been identified among the most promising prospects for constructing a broad range of sensing platforms due to their unique properties, such as outstanding specificity and sensitivity, easiness of fabrication, and excellent durability in a variety of circumstances. For the manufacturing of aptamer-based nanoprobes, many signals-transduction approaches have been developed. Incorporating numerous aptasensing techniques with NPs has improved biosensor selectivity and sensitivity in recent years [64].

3. Functionalized Nanomaterials

3.1. Metallic Nanoparticles

Metal-based nanomaterials consist of nanoparticles of raw metal, such as gold (Au) and silver (Ag). AuNPs are inert in bulk, while become highly reactive in nanoparticle form [65]. The exciting surface chemistry of AuNPs opens up novel routes for the progress of unexplored multifunctional instruments for biomedical and nanotechnological applications [66]. Nanotechnology applications have drawn a great deal of interest since the late 1980s [67]. The exceptional electrical and optical properties of Au boost their use in biosensing and bioimaging. The use of organic molecules to functionalize Au NPs aids the conjugation of drugs for delivery systems. Thus, AuNPs can be used as photothermal therapeutic agents [56]. For instance, surface-modified AuNPs have been prepared via a layer-by-layer procedure with alternating polyelectrolyte layers of cationic polyallylamine and anionic poly(acrylic acid). Subsequently, papain was covalently immobilized on the modified AuNPs via amide bond between the NH2 groups of papain and the terminal COOH groups of the modified NPs, using EDC and sulfonated NHS as coupling agents, as depicted in Figure 7, to produce a heterogeneous biocatalyst that has been applied in bioanalysis and biopharmaceutical analysis [68].
The conjugation of gold nanorods (AuNRs) onto micelles, via gold-thiolate complex formation, brings photosensitivity to the nanoassembly. The size and surface morphology characterization via TEM (Figure 8) indicated that the mean micellar size was around 15 nm, and the thickness and length of the AuNRs was about 20 and 65 nm, respectively. The percentage of conjugated AuNRs to the micelles was roughly 12%. The attachment of chitosan transfers the photosensitivity of functionalized AuNRs to micelles, and the micelle thermal shrinkage induces the release of paclitaxel, a drug widely used to treat breast cancer [69].
Nejati et al. [19] examined functionalized AuNPs in biomedical applications. To attain this goal, their structure, production, and functionalization were extensively explored and discussed. Gold NPs have been utilized in biological applications, electrochemical technology, and radiation oncology. Multifunctionalization, that is, functionalization that allows for the provision of more than one attribute at a time, provides added value to these NPs due to synergistic effects. Multifunctionalized gold NPs have been discovered to be a viable choice in biomedicine for delivering anticancer drugs and antibiotics for combined photothermal and chemical therapy [70]. AuNPs are suitable for the delivery of the drugs to cellular destinations due to their ease of synthesis, functionalization, and biocompatibility. Figure 9 depicts functionalization of AuNPs for gene and drug delivery. AuNPs functionalized with targeted particular biomolecules can successfully kill tumor cells or bacteria (Figure 9). Large surface-to-volume ratio of AuNPs can carry a huge amount of drug molecules. AuNPs have been applied for the codispensation of protein drugs owed to their skill in penetrating cell membranes, probably because they can interact with the lipids present on the cell surface [70].
Despite the efforts carried out, more studies into intelligent drug delivery based on nanoparticles, particularly gold NPs, is required. Despite numerous publications, only a few clinically authorized drug delivery nano systems are currently accessible in the industry. As a result, an immediate need is found to incorporate animal-model research into clinical practice [70]. Donoso–Gonzalez et al. [71] used cationic cyclodextrin-based polymer (CCD/P) to load phenylethylamine (PhEA) and piperine (PIP) onto gold nanostars (AuNSs). They evaluated the product potential for simultaneous drug loading and SERS-based detection. In addition to PhEA and PIP, the polymer contained AuNSs that had been functionalized with PhEA and PIP, resulting in a unique AuNS-CCD/P-PHEA-PIP nanosystem with an optimum size and Z potential for biomedical applications. Hybrid materials incorporating carbon nanomaterials and AuNPs have also been synthesized. For instance, Shon et al. [72] reported the synthesis of soluble fullerene-linked AuNPs using a modified Brust reaction and subsequent ligand exchange reaction of hexanethiolate-protected Au NPs with 4-aminothiophenol. Amination of C60 with 4-aminothiophenoxide ligands produced the C60-linked AuNPs. This approach enables the control of the optical and photochemical properties of the nanoparticles. Sudeep and coworkers [73] developed a self-assembled photoactive system comprising AuNPs as the central core and fullerene moieties as the photoreceptive hydrophobic shell via functionalization of the NPs with a thiolated fullerene derivative. Yaseen et al. [74] used C60-terminated alkanethiol to synthesize novel fullerenethiol-functionalized gold nanoparticles (C60−AuNPs) of 2 nm diameter with an extremely narrow size distribution. The fullerene-thiol moiety was inserted into the fullerene by the ligand exchange method. Liz–Marzán et al. [75] developed Au core/SiO2 shell nanocomposites with tailorable thickness and good dimensional stability. Citrate-capped AuNPs were first synthesized and then reacted with aminopropyl trimethoxysilane, a widely used coupling agent, which anchored onto the NPs via silanol groups. Active silica was subsequently added, leading to the formation of a fine, dense, and fairly homogeneous silica layer wrapping the NPs.
On the other hand, AgNPs are antibacterial and anti-inflammatory, and possess excellent biocompatibility [76]. The AgNPs can be straightforwardly synthesized via simple, fast, nontoxic, and environmentally friendly means so that they can produce NPs with perfectly defined morphology and size. They were applied as a coating for cardiovascular implants to improve their biocompatibility. Additionally, their antimicrobial, antifungal, antiviral antiangiogenic, and anticancer properties make them suitable in a large number of biomedical and health care areas including device coatings, drug delivery systems, wound dressings, the textile industry, photothermal therapy, and so forth. The biological activity of AgNPs is influenced by many parameters such as the NP shape, size, morphology, state of dispersion, solution rate, reactivity, and ion discharge efficiency, amongst others, which condition their cytotoxicity. The design of AgNPs with uniform functionality, size, and morphology is crucial from a practical viewpoint. Other metallic NPs such as ruthenium and selenium have been applied in nanomedicine [77], in particular for drug delivery.

3.2. Metal Oxide-Based Nanomaterials

A wide number of variations of metal oxide NPs have been used in nanomedicine, such as iron oxide (Fe2O3, Fe3O4), CeO2, titania (TiO2), ZnO, NiO, silica (SiO2), and so forth [78,79,80]. Iron oxide NPs are a fascinating family of nanostructures that have attracted much interest in the medical area because of their negligible toxicity, high biocompatibility, and inherent magnetic properties, which make them perfect candidates for therapeutic and diagnostic goals, particle imaging, and as contrast agents in magnetic resonance imaging (MRI) and ultrasonic techniques [81]. The incorporation of Fe3O4 NPs also enhances the antimicrobial properties [82].
The five most popular strategies to generate hollow iron oxide NPs are the Kirkendall effect, galvanic substitution, chemical etching, nano-template-mediated, and hydrothermal/solvothermal routes [83]. Cheah et al. [84] synthesized iron oxide NPs in diethylene glycol (DEG) by thermal decomposition of iron (III) acetylacetonate (Fe(acac)3), and subsequently changed the surface of the NPs by adding surface ligands (Figure 10). Using this easy production process, surface modification of iron oxide NPs with various covering substances such as dopamine (DOPA), polyethylene glycol with thiol end group (thiol-PEG), and poly(acrylic acid) (PAA) is achievable. The size of these NPs can be precisely controlled at the nanometer scale by continuous growth. TEM images confirmed that the morphology did not change upon functionalization (Figure 10). Besides, NPs with PAA coating can be used as contrast agents. The surface change of oleic-acid-coated iron oxide NPs (Fe3O4-OA) (made by coprecipitation method) with tetraethylorthosilicate was studied by Nayeem et al. [85] using an inverse microemulsion approach (TEOS). To obtain thermally sensitive magnetic nanocomposites (MNCs), Fe3O4/SiO2/P(NIPAm-co-AMPTMA), the surface of iron oxide nanoparticles was tailored using a multistep approach with poly [N-isopropylacrylamide-co-(3-acrylamidopropyl) trimethylammonium chloride], P(NIPAm-co-AMPTMA). Magnetic nanoparticles (MNPs) have been extensively studied as MRI contrast agents to aid in the detection, diagnosis, and treatment of solid cancers. The absorption of superparamagnetic iron oxide NPs (SPIONs) in the endothelial reticulum system (RES) can be used in medical imaging to detect liver neoplasms and metastases. It can also currently differentiate tiny lesions of 2–3 mm. Furthermore, ultrasmall superparamagnetic iron oxide NPs (USPIONs) show promising utility in MRI exams for the identification of lymph node metastases that are 5–10 mm wide [86]. By utilizing the distinct molecular fingerprints of these disorders, the future iteration of active targeting MNPs, which has recently been explored, has the capacity to enhance tumor detection and characterization [86].
Cerium oxide (CeO2) NPs, named as nanoceria, have the unique property of anti-inflammation. They have better redox as well as potential antioxidant properties with therapeutic characteristics. TiO2 has the unique properties of high chemical stability, cytocompatibility, and optical properties [87]. The biocompatible properties of TiO2 NPs have increased their usage in drug delivery, bone substitute materials, bone regeneration, cell and tissue behavior modulation, vascular stents, scaffolds, bioimaging, and biosensors [78]. MSN also have great potential for nanomedicine. In fact, upon functionalization, they can be efficiently targeted to cancer cells [59] and be used for encapsulation and controlled release of drugs [27]. For biomedical applications, ZnO possesses the properties of low toxicity and biodegradability. It can be used for the purpose of drug delivery, gene delivery, biosensing, bioimaging, etc. [88]. CuO NPs have also been used for targeted drug delivery in breast cancer therapy [35].

3.3. Ceramic-Based Nanomaterials

A wide range of ceramics, including Ca3(PO4)2, bioactive glass, Al2O3, ZrO2, CaCO3 and so forth, are getting countless interest in the biomedical field, particularly in the tissue engineering arena. Thus, their outstanding osteoconductivity, resorbability, biocompatibility, biodegradability, and hydrophilicity make them appropriate for numerous hard tissue applications [89,90]. They can be divided into three types: bioinert, bioactive, and resorbable. Resorbable ceramics are progressively adsorbed and substituted by endogenous tissue. They can be synthesized in the forms of nanocrystals, NPs, nanopowders, or nanocoatings. The most popular is Ca3(PO4)2, which is widely applied in the form of NPs and nanocements for orthopaedic and dental uses. The optimal surface charge density, functionality, and solution characteristics of this ceramic account for its fittingness in drug delivery and growth factor uses. Bioactive ceramics such as hydroxyapatite (HDA) NPs are a type of calcium phosphates that have been comprehensively investigated in bone regeneration and antibacterial applications [91,92]. They are osteoconductive and can link to bone tissues via chemical bonding, following the rule of bonding osteogenesis. Furthermore, for bone tissue engineering, bioactive glass is crucial, owed to its outstanding osteoconductivity, osteoinductivity, and biocompatibility [93]. Bioinert bioceramics such as ZrO2 have great chemical stability and in vivo mechanical strength. This oxide is regarded as a nontoxic material and has strong resistance to acids; hence, it is widely used in coatings for metallic load-bearing implants and dentistry. Another widely used oxide is Al2O3, which possesses high hardness and superior heat resistance, and has been applied in arthroplasty, dentistry, and as an antimicrobial coating [94].

3.4. Carbon-Based Nanomaterials

Within carbon-based nanomaterials, carbon nanotubes (CNTs), graphene oxide (GO), and graphene quantum dots (GQDs) have been broadly explored in biomedical applications [13,95,96]. Purification, separation, dispersion, stability, alignment, functionalization, and arrangement of CNTs are critical parameters to be controlled prior to their applications [97]. Since the discovery of CNTs, numerous physical and chemical techniques have been developed to attain these goals [98,99]. Polysaccharides with a broad range of characteristics, large-scale production, and low prices have shown to be highly suitable for CNT functionalization. The use of chitosan for CNT purification and functionalization has been proven to be a strategy to make drug release easier and more effective. Dou et al. [100] described a one-pot tactic for the development of chitosan-coated CNTs via a combination of Diels–Alder reaction and mercaptoacetic acid locking imine (MALI) reaction (Figure 11). Taking into account the broad use of Diels–Alder chemistry and MALI reaction, several carbon nanomaterials with different functional groups might be synthesized and applied to biomedicine.
Graphene and graphene oxide (GO) are 2D carbon-based nanostructures, in the form of nanosheets, that show an optimal combination of biocompatibility, strength, flexibility, and optical transparency, which made them suitable for the design of selective and sensitive sensors of biomolecules, which is crucial for medical sciences and the healthcare industry in order to assess physiological and metabolic parameters [101]. Besides, they show antibacterial and antiviral properties [96,102,103]. Graphene-based systems have proven to be effective via direct interaction with viruses and through photo-induced mechanisms, as well as platforms for other particles or molecules with antiviral properties. GO inactivates the virus by physical disruption: it can adhere to the structure of virus spikes and destroy them with the sharp edges of the GO layers. Its antiviral activity is effective on both DNA and RNA viruses, and depends on the concentration and incubation time. Reduced graphene oxide (rGO) and GO show similar antiviral activity, pointing towards a minor influence of the surface functional groups. The physical interaction of the viruses with their sharp edges seems to be the leading cause for the antiviral activity. Besides, they are negatively charged, which enables electrostatic interaction with the positively charged viruses. The higher interactions result in the destruction and inactivation of the virus. The viruses captured by GO have shown a loss of structural integrity: an RNA is released. The virus can then be identified using the recovered RNA [104]. Another method to inhibit the virus activity is using the GO photocatalytic activity. This approach has been developed by Hu et al. [105] to synthesize GO-aptamer nanosheets that were used to capture MS2 bacteriophage viruses, a small icosahedral nonenveloped RNA virus, which infects E. coli bacteria. This was used as a model for testing the antiviral properties of GO upon illumination with UV light. In this case, the leakage of the virus protein capsid predominates over the physical disruption produced by the sharp edges of the GO sheets.
Carbon quantum dots (CQDs), 0D carbon-based nanomaterials with fluorescence characteristics, also exhibit antimicrobial and antiviral properties [106]. These include amorphous carbon nanoparticles, graphene quantum dots (GQDs), partially graphitized core–shell carbon NPs, and amorphous fluorescent polymeric NPs. Their activity is attributed to the functional groups on their surface. CQDs functionalized with boronic acid demonstrated antiviral efficacy against HCoV-229E Human Coronavirus. HCoV-229E is an enveloped, single-stranded RNA coronavirus. It is one of the viruses that produce the common cold (Coronaviridae family, Human coronavirus 229E species), with a diameter in the range of 120–160 nm. Figure 12 shows two pathways for antiviral activity: (1) the attachment of CQDs (with a mean diameter of about 7 nm to the S-protein of viruses) to prevent infectious contacts between host cells and viruses; and (2) the capacity of CQDs to disrupt RNA genomic replication. Boronic acid functions were crucial in determining antiviral efficacy [107].
Bai et al. [108] developed a molecularly imprinted fluorescent sensor for selective identification of a model drug: paclitaxel. A molecularly imprinted polymer (MIP) shell was grafted on the surface of silane-functionalized Mn:ZnS QDs using a free radical polymerization procedure (Figure 13). Methacryl polyhedral oligomeric silsesquioxane (M-POSS) was utilized to provide a porous structure. Figure 13 depicts the synthesis process and the potential detection mechanism of the drug.
Van Tam et al. [106] used microwave-assisted pyrolysis of fructose to synthesize aniline-functionalized graphene quantum dots (a-GQDs). Then, phenyl boric acid (PBA) was used to modify the a-GQDs, leading to a fluorescence-quenching effect. The a-GQDs/PBA nanomaterial was tested as a fluorescence turn-on sensor for glucose detection, based on the specific interaction between PBA and glucose.
QDs also have great potential for cancer treatment. The selective attachment of FR-positive tumor cells with folic acid/folate (FA) was reported as a fast and easy technique for determining folate receptor (FR) expression in cancer cells. MKN 45, HT 29, and MCF 7 cancer cells were selectively marked using graphene quantum dots with folate coating and nitrogen doping (N-GQDs) [109]. DNA-functionalized QDs have drawn considerable attention in sensing and imaging, as well as cancer therapy [110]. Covalent conjugation, electrostatic interaction, direct dative interactions, and other ways for conjugating DNA to QDs have been documented in the literature [111]. In vitro photothermal imaging was described by Wang et al. [112] as AuNPs-QD complexes combined with DNA as a template. Horo et al. [52] developed DOX-loaded chitosan-AuNPs and beads, both of which were implanted with functionalized silk fibroin. Chitosan was used as a reduction and stabilizing agent to synthesize NPs with dimensions in the range of 3-8 nm. Compared to uncoated materials, coated materials demonstrated a delay in drug release. As a result, drug delivery strategies based on functionalized silk-coated substances may be useful for producing localized and protracted drug release.

3.5. Polymeric Nanomaterials

Polymeric NPs are colloidal particles in the range of 10 nm–1 μm made up of polymers that can be straightforwardly synthesized through chemical reactions in order to tailor the loading and release of drugs and genes. The benefits of these NPs are their easiness to synthesize, high stability, biodegradability, nontoxicity, lengthy blood circulation time, and sustained and targeted delivery. Furthermore, they can be tailored according to their shape, size, surface functional groups, degree of porosity, as well as their mechanical characteristics [113]. They are divided into three main groups: natural, biosynthesized, and chemically synthesized. They can be fabricated into different shapes, including liposomes, dendrimers, nanospheres, nanocapsules, nanogels, and micelles (Figure 2). They are used in wound dressings, pharmaceutical excipients, medical devices, dental materials, and scaffolds [114]. Biodegradable polymers frequently used for the development of polymeric NPs are poly(lactide) (PLA), poly(ɛ-caprolactone) (PCL), PLGA and polycarbohydrates such as alginate, chitosan, and gelatin.
Overall, because of their excellent chemical, physical, and mechanical properties and their versatility of synthesis, functionalized nanomaterials can be employed in a variety of ways. Although functionalized nanoparticles are hardly used in the industrial field up to date, they can aid in developing novel concepts in a variety of industries. Functionalized nanomaterials promise to produce better and cost-effective consumer products and industrial operations. An inappropriate use can have a detrimental effect on surroundings, public health, and safety in various ways [115,116,117].

4. Cytotoxicity: The Role of Functionalization

Chemical composition, crystalline structure, size, and density are parameters that strongly influence nanomaterial toxicity and cytotoxicity [28,118]. Nanomaterial absorption and intracellular localization can be linked to some health hazards due to the nature of nanomaterials and their chemical interactions with cells. Chemical composition, for example, might cause oxidative stress in cells [119]. CNTs are believed to be more poisonous than carbon black or silica nanoparticles and can induce severe lung damage [120]. Asbestos is less hazardous than TiO2, Fe3O4, and ZrO [121]. Another indicator of cytotoxicity based on membrane integrity damage is lactate dehydrogenase (LDH) leakage. Additionally, DNA damage in primary mouse embryofibroblasts (PMEF) treated in vitro with different amounts (5, 10, 20, 50, and 100 μg mL−1) of manufactured nanoparticles (Figure 14) revealed that CNTs and ZnO caused more DNA damage than carbon black (CB) and SiO2 NPs [121].
The crystalline structure also has a strong effect on NPs’ toxicity [122]. For instance, TiO2 NPs, which can naturally appear in three different crystalline forms, i.e., anatase, rutile, and brookite, are reported to have cytotoxic and genotoxic effects. Rutile titania is slightly more lethal than anatase TiO2 NPs. This might be elucidated considering the different reactivity of the two forms: rutile TiO2 NPs are more photocatalytic than anatase and therefore, are capable of producing larger quantities of oxygenated free radicals on their surface. On the other hand, other allotropes have a significant influence on cell viability and, as a result, on human health. Sato et al. [123] discovered that TiO2 allotrope toxicity is affected by the NPs’ environment’s ambient conditions. In the absence of UV radiation, rutile TiO2 NPs (200 nm) caused oxidative DNA injury, whereas TiO2 NPs (10–20 nm) caused oxygen-reactive species (ROS) generation.
Another key component in minimizing nanomaterial toxicity is particle size [124]. Nanoparticles with a smaller size are more prone to pass through biological barriers. Phagocytosis or other pathways can facilitate the entrance of small NPs to cells. NPs can discriminate between adhesive connections because of their ability to infiltrate cells. It can produce forces such as Van der Waals, steric interactions, and electrostatic charges [125]. Furthermore, unlike big nanoparticles, NPs in the size range of 1 to 100 nm are not phagocytized but instead taken up via RME routes. In the lack of specific cell surface receptors, NPs can be absorbed. Most cells can effectively assimilate NPs with size of 50 nm or smaller (causing cytotoxicity). NPs smaller than 20 nm can easily pass through blood arteries and concentrate in organs [126]. NPs with a large surface area, such as NiO (diameter < 25 nm), clump together in liquids, and engage and induce oxidation and DNA damage by interacting with molecules including proteins and DNA [127]. The mechanisms of cell damage by NPs are depicted in Figure 15 [124].

5. Cost-Effective Functionalization

The functionalization of AuNPs using a mixture of DNA and PEG polymers is the most cost-effective and satisfactory method available for nanomaterial cofunctionalization [128,129]. To obtain a comparable level of gold NP binding effectiveness with DNA origami nanostructures, Wang et al. [130] used a significantly smaller amount of thiol-DNA in their technique than pure DNA functionalization. Because of the decreased DNA consumption and lower costs, the use of DNA–NP conjugates in nanotechnology can be scaled up. Figure 16 shows the functionalization process of AuNPs with DNA/PEG polymers [130].

6. Applications of Functionalized Nanomaterials in Biomedicine

Medical diagnosis [131], immunization [132], treatment [133], and even healthcare services have been transformed and influenced by nanotechnology [134]. Chemical functionalization, physical functionalization, and surface synthesis link biological agents with various NPs. It is possible to classify the biomedical applications of nanotechnology into different areas, as summarized in Figure 17 [135]. Additionally, some relevant examples have been provided for each category in Table 1.

6.1. Diagnostic Implications of Functionalized Nanomaterials

Nanomaterials are extensively employed in imaging modes, such as optical coherence tomography and MRI. QDs are semiconductor nanocrystals commonly employed in optical imaging [151]. Imamura et al. [152] used PbS QDs for noninvasive scanning of septic encephalopathy in mice, suggesting that these nanomaterials can be used to image a variety of vascular systems. NIR fluorescence imaging of the mouse brain during therapy with Pbs QDs is shown in Figure 18 [152]. Before administration of QDs, only low-intensity NIR fluorescence signals were distinguished (Figure 18b, middle), due to the extremely low background fluorescence in this spectral zone. When QDs were intravascularly inserted into the mouse, the fluorescence signals arising from the mouse head augmented, and the vascular structure of cerebral blood vessels became visible (Figure 18b, right)
The development of nanoparticles with fluorescence characteristics for in vivo imaging is currently in progress. Because silicon nanocrystals are cell-safe, abundant, and more appealing than QDs [153], they do not necessitate a dense surface coating to protect the nanocrystal center from oxidation and the environment.

6.2. Therapeutic Applications of Functionalized Nanomaterials

Magnetic nanoparticles, AuNPs, and CNTs have been utilized in the field of biomedicine. The application of NPs in postoperative treatment has attracted the attention of many researchers [153]. The use of superparamagnetic Fe3O4, GO, and doxorubicin-incorporated nanofibers has been claimed to reduce the localized regression of breast cancer and develop tissue regeneration [60]. A functionalized peptide that provides specific drug delivery possibilities with improved drug permeability, noteworthy aggregation in the desired target, and high therapeutic efficacy can help with the liposomal formulation in cancer treatment [20]. Docetaxel is a widely used anticancer chemotherapy drug, and transferrin is a blood–plasma glycoprotein that plays a key role in iron metabolism. Fernandes et al. [30] synthesized docetaxel-loaded liposomes functionalized with transferrin (LIP-DTX-TF), and their effects on prostate neoplasms were studied. TEM images demonstrated that the systems were spherical and nanometric in size (Figure 19a) and that the presence of DTX aided in vesicle size reduction, resulting in improved liposome stability (Figure 19b).

7. Functionalized Nanomaterials: Drug/Gene Delivery

Nanomaterials can be functionalized for different purposes, including drug delivery carriers or therapeutic agents for cure and treatment, diagnostic applications in biological imaging, cell labeling, biosensors, and the use of moieties for medical devices such as stents or lenses [154,155,156]. Functionalization can improve biocompatibility and uptake efficiency and simultaneously minimize immune system activation, increasing the material’s bioavailability inside the body. These modifications are beneficial for some drug delivery strategies to ensure that the appropriate doses of the drug are released to the correct area while limiting the detrimental effects of drug molecules on other organs [157]. Drug delivery systems are necessary to improve the efficacy of drug biodistribution. Nanomaterials have been used to carry drugs and genes in passive, active, and direct methods [49,158,159]. Due to the small size of nanoparticles, they can pass across cellular membranes and boundaries. Moreover, the increased surface-to-volume ratio of nanoparticles leads to improved drug loading [160]. Figure 20 displays biological ligands used for active targeting of NP drug carriers [161], and Table 2 summarizes different functionalized nanomaterials applied in drug/gene delivery.

8. Functionalized Nanomaterials: Regenerative Medicine

Reparative and restorative medicine and nanotechnology have gained popularity in recent years, resulting in significant improvements in medical research and clinical practice [180]. Tissue engineering, cell therapy, diagnostics, medication, and gene delivery are examples of regenerative medicine applications that use various functionalized nanoscale materials [24,181,182]. Restorative medicine is a vast field of nanotechnology that strives to regenerate cells and tissues similar enough to their original design and function. Three main types of therapeutic techniques can be found in regenerative medicine: tissue-engineering treatments based on cells; biomaterials; and a combination of the two. Stem cell biology, nanotechnology, and bioengineering have progressed significantly, potentially paving the way for real regenerative medicine for various diseases [183]. Stem cells are known for their capacity to maintain their differentiating potential while intersecting to generate numerous daughter cells. Such daughter cells lack “stem-ness” and use controlled proliferation to produce adult cells of all origins throughout the body (self-renewal) [184]. Using tissue-specific or therapeutic genes, as well as primary cells that overexpress these genes, genetically engineered cell treatment can manufacture proteins with a therapeutic intent, to be used at regeneration platforms or discriminate new cells into the appropriate cellular lineage, assisting in tissue restoration [185].
When bone is formed, it comprises mostly collagen fibers and calcium phosphate, which is converted into hydroxyapatite (HDA). Bone tissue also contains several cellular structures, such as osteoblasts, osteocytes, and osteoclasts, which contribute to its calcification [186]. For bone repair, nanoscaffolds with adequate biophysical characteristics, such as stiffness and cell proliferation, have been employed. A variety of nanofiber matrices have been synthesized in recent years. The vast majority of nanoscaffolds are developed to match genuine bone’s structural, compositional, and biological features [187]. Zhang et al. [188] prepared a chitosan/HDA biomimetic nanocomposite scaffold for assessing the effect of bone marrow MSC mesenchymal stem cells (BMSCs) growth, and explored the molecular mechanism both in vivo and in vitro. It was reported that this hybrid scaffold could encourage the proliferation of BMSCs and trigger the integrin-BMP/Smad signal pathway of BMSCs. In addition, HAD can also be used combined with other polymeric materials such as PEG, PCL, and PLGA, which have displayed improved effects in bone regeneration/repair (Figure 21) [189,190].
Besides nano-HDA, collagen, electrospun silk, anodized titanium, and nanostructured titanium surfaces are some of the primary constituents of materials that mimic the bone extracellular matrix [191]. In primary osteoblasts, nanofibers have improved osteogenesis and biomineralization. Main osteoblasts are limited in their application due to (i) restricted accessibility and intrinsic donor site malady; (ii) limited scaling capability; (iii) age-related behavior; or (iv) possibility of dedifferentiation occurring during in vitro cultivation [94].
Transient gene delivery [192], cell therapy without the need for genetic modification [193], and genetically modified cells [194] are currently three of the most exciting new procedures in the field of tissue engineering. Gene delivery is a therapeutic approach to introducing foreign genetic material directly into host cells in vivo. These genes immediately affect the host tissue, causing it to remodel [195].
Achieving better cell adhesion, motility, and differentiation through nanomedicine is possible thanks to the development of interfaces, components, and substances that mimic the cells’ natural environment. Scientists have developed complex tissue/organ constructions by combining stem cells with scaffolds and stimulating factors as the basis of their tissue-engineering experiments [145,196]. Some of these are currently being utilized therapeutically as part of the standard treatment for various disorders. Scaffolds are transformed into three-dimensional structures that have the appropriate shape, size, architecture, and physical properties for different applications and environments. For this reason, tissue-engineering products are designed to look and behave like natural tissues. In addition to biocompatibility and controllable porosity and permeability, important scaffold characteristics include mechanical and degradation kinetics comparable to those of the desired tissue and support for cell adhesion and proliferation by adding nanotopographies to the biomaterial surface [197].
Biodegradability is a critical property that nanoparticles must possess to be employed safely inside the body. This is a crucial aspect to consider when building scaffolds for tissue engineering and reparative and restorative medicine [198]. Table 3 summarizes functionalized nanomaterials that have been utilized in tissue engineering.

9. Functionalized Nanomaterials: Cancer Therapy

Theranostic nanoprobes for tumors and malignancies have become a prominent focus of research since NP functionalization has been able to be used simultaneously in diagnostic and therapeutic purposes. Surface modification of NPs has been proven to generate targeted accumulation in tumor tissue due to the enhanced permeability and retention (EPR) effect [29,213]. Tumors have more permeable vasculature, a poorly defined lymphatic system, and various substances that aid in increased targeting, as contrasted to normal tissue, such as VEGF and basic fibroblast growth factor [214]. In cancer immunotherapies, NPs can keep track of critical immune cells during metastasis. Different tumor ablation therapies with magnetic NPs such as Fe3O4 have been reported [215] (Figure 22): (a) Magnetic hyperthermia, in which an alternating magnetic field induces NPs to produce heat, boosting tumor necrosis. (b) Photothermal ablation, in which the light absorbed by the NPs is transformed into thermal energy, producing cell death in the neighborhood. (c) Photodynamic therapy, in which photosensitizing agents anchored to NPs are activated via an external light source to make singlet oxygen species that are cytotoxic to cells. As a result, NPs have a high level of target-cell selectivity [216]. Table 4 displays functionalized nanomaterials that have been utilized for cancer treatment.

10. Functionalized Nanomaterials: Medical Implants

Recently, the influence of nanotechnology on the implant field has increased strongly. Nanomaterials with biological-inspired structures are motivating scientists to investigate their potential for enhancing the performance of conventional implants [91,142]. Nanotechnology has the skill to economically substitute many traditional implants and offer numerous novel applications. It can result in more efficient and longer-lasting implants, with reduced infection rates and enhanced bone or tendon healing. In orthopedics, the goal of biomaterials is to substitute injured bone. Improved mechanical properties (e.g., strength, flexibility, hardness, elastic modulus), wear, hydrolysis and corrosion resistance, biocompatibility, osseointegration, bioinertness, and ease of surgical application are required properties to be used in orthopedics [226,227]. Nanomaterials offer an enlarged surface area, a superior stiffness, and a high roughness that can improve the adhesion and proliferation of bone-related proteins and the deposition minerals incorporating Ca [228]. Besides, FMNs can mimic the amounts of the components of natural bones and can aid in sustaining biologically active growth factors and exploit the potential of BMSCs. Numerous studies [229] have been developed to examine the optimal surface properties of FMNs that may support or assist specific protein adsorption, improved osteoblast anchoring, osteoblast differentiation, and new bone formation (Figure 23) [142].
Surface adjustment of nanomaterials is a prospective method to expand the performance and durability as well as to reduce the hazardous side effects that might take place during implant degradation. Surface characteristics have a key role on modulating biological interactions. Specifically, engineered nanomaterials can have a significant impact on molecular and cellular actions; this issue aids in conditioning the comprehensive biological response of an implant (i.e., protein adsorption, cell adhesion, and proliferation). Therefore, several approaches have been settled to modify nanomaterials for orthopedic implants such as anodic oxidation [230], plasma electrolytic oxidation [231], electrochemical plating [232], chemical conversion coating [233], physical vapor deposition, laser surface alloying [234], thermal spraying [235], organic coating [236], and so forth. These methods provide new implant surfaces with tailorable characteristics at the nanoscale. The particular procedure can be chosen based on different factors/goals, including to attain complex geometries and to be suitable for large-scale processing. Metal oxide NPs such as TiO2, ZrO2, and Al2O3 have been used as nanocoatings to enhance the mechanical and biochemical properties of conventional metallic implants [237].

11. Conclusions

Nanotechnology has opened up vast techniques to manipulate and transform the current medical devices or materials utilized for therapy in biomedical sciences and engineering. Numerous nanomaterials can be used in biomedical applications, both organic, such as CNTs, GO, GQDs, and polymeric NPs, and inorganic, such as metallic NPs (Au, Ag), metal oxide NPs (TiO2, Fe3O4, mesoporous SiO2), and ceramic (HAD, CaCO3). Over the last years, numerous approaches have been developed to synthesize surface-engineered nanomaterials, in particular NPs, for drug/gene delivery, diagnostics, cancer therapy, tissue engineering, and medical implants, and the structure–function relationship of these functionalized nanoparticles has been widely explored. The NPs’ surface modification is a potent strategy to improve biocompatibility and uptake, as corroborated by the huge quantity of scientific documents published on this subject. Investigations prove that the conjugation of polymers, biomolecules, and small ligands on the NP surface can successfully increase biocompatibility both in vivo and in vitro, due to the alteration of surface charge and to the inactivation of sensitive functional groups that can influence the stability of the cell membrane. Besides, the incorporation of certain molecules can improve NPs’ passive and active uptake, reducing systemic toxicity in vivo and enabling high precision therapy and/or diagnosis. The binding of functionalization agents on the NP surface can be achieved via covalent and noncovalent tactics. The first is broadly used to link proteins, antibodies, aptamers, and peptides utilized to boost uptake and to achieve active targeting, whereas the second is frequently used for the loading of drugs and other molecules that need to be liberated into the cells. The promise of tissue and organ-specific regeneration therapy has become a reality due to major advances in regenerative medicine and nanomedicine over the previous decade. Preliminary clinical results have shown that functionalization of NPs with specific recognition surface moieties results in improved efficacy and reduced side effects, due to properties such as directed localization in tumors and active cellular uptake. Even though remarkable improvements have been attained, this research arena is still in its early stages, and significant efforts are needed in order to be able to scale up the functionalization approaches developed at the laboratory level and make them reproducible. A prerequisite for progressing in this research area is the development of novel chemical methods to conjugate chemical moieties onto NPs in a safe and consistent manner. In addition, smart and innovative nano-based technologies can offer particular physicochemical properties that could aid in fixing crucial issues associated with the treatments of viral infections such as SARS-CoV-2. Researchers may find this study valuable in analyzing past studies on the topic matter to attain commercial success.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

This article’s data sharing is not applicable as no new data were created or analyzed in this study.

Conflicts of Interest

The author declares that there is no conflict of interest regarding the publication of this article.

Abbreviations

aniline functionalized graphene quantum dotsa-GQDs
aptameric functionalized materialsAFMs
carbon nanotubesCNTs
carbon quantum dotsCQDs
cationic β-cyclodextrin-based polymerCCD/P
docetaxel-loaded liposomes functionalized with transferrinLIP-DTX-TF
dopamineDOPA
dopamine-polyethylene glycol-carboxylic acidDPA-PEG-COOH
enhanced permeability and retentionEPR
extra-cellular matrixECM
folate receptorFR
folic acidFA
folic acid-coated gold nanoparticles conjugated with fluorophoreFA-Au-FITC
functionalized nanoparticlesFNPs
gold nanoparticlesAuNPs
graphene oxideGO
hyaluronic acidHLA
hydroxyapatiteHDA
lactate dehydrogenaseLDH
poly(lactic-co-glycolic acid)PLGA
magnetic nanoparticlesMNPs
magnetic resonance imagingMRI
mesoporous silica nanoparticlesMSN
methacryl polyhedral oligomeric silsesquioxaneM-POSS
nanoparticlesNPs
near-infraredNIR
nitrogen-doped carbon quantum dotsNCQDs
nitrogen-doped graphene quantum dotsN-GQDs
oleic acid-coated iron oxide NPsFe3O4-OA
phenyl boronic acidPBA
phenylethylaminePhEA
photodynamic therapyPDT
photothermalPT
piperinePIP
poly(acrylic acid)PAA
polyethylene glycolPEG
polyethylene glycol with thiol end groupthiol-PEG
polyethylene glycol-gelatin-chitosan-hyaluronidase-5-fluorouracilCS-HYL-5-FU-PEG-G
positron emission tomographyPET
quantum dotsQDs
receptor-mediated endocytosisRME
reticulum endothelial systemRES
sodium alginate (SA)–polyvinyl alcohol (PVA)–bovin serum albuminSA-PVA-BSA
sodium alginate/polyethylene glycol (vinyl alcohol)SA/PVA/Ca
superparamagnetic iron oxide NPsSPIONs
tetraethylorthosilicateTEOS
ultra-small superparamagnetic iron oxide NPUSPIONs

References

  1. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; del Pilar Rodriguez-Torres, M.; Acosta-Torres, L.S.; Diaz-Torres, L.A.; Grillo, R.; Swamy, M.K.; Sharma, S.; et al. Nano based drug delivery systems: Recent developments and future prospects. J. Nanobiotechnol. 2018, 16, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bilal, M.; Qindeel, M.; Raza, A.; Mehmood, S.; Rahdar, A. Stimuli-responsive nanoliposomes as prospective nanocarriers for targeted drug delivery. J. Drug Deliv. Sci. Technol. 2021, 66, 102916. [Google Scholar] [CrossRef]
  3. Rauf, A.; Tabish, T.A.; Ibrahim, I.M.; Hassan, M.R.U.; Tahseen, S.; Sandhu, M.A.; Shahnaz, G.; Rahdar, A.; Cucchiarini, M.; Pandey, S. Design of Mannose-Coated Rifampicin nanoparticles modulating the immune response and Rifampicin induced hepatotoxicity with improved oral drug delivery. Arab. J. Chem. 2021, 14, 103321. [Google Scholar] [CrossRef]
  4. Hong, E.J.; Choi, D.G.; Shim, M.S. Targeted and effective photodynamic therapy for cancer using functionalized nanomaterials. Acta Pharm. Sin. B 2016, 6, 297–307. [Google Scholar] [CrossRef] [Green Version]
  5. Rahdar, A.; Hajinezhad, M.R.; Hamishekar, H.; Ghamkhari, A.; Kyzas, G.Z. Copolymer/graphene oxide nanocomposites as potential anticancer agents. Polym. Bull. 2021, 78, 4877–4898. [Google Scholar] [CrossRef]
  6. Zhang, G.; Khan, A.A.; Wu, H.; Chen, L.; Gu, Y.; Gu, N. The Application of Nanomaterials in Stem Cell Therapy for Some Neurological Diseases. Curr. Drug Targets 2018, 19, 279–298. [Google Scholar] [CrossRef]
  7. Theus, A.S.; Ning, L.; Jin, L.; Roeder, R.K.; Zhang, J.; Serpooshan, V. Nanomaterials for bioprinting: Functionalization of tissue-specific bioinks. Essays Biochem. 2021, 65, 429–439. [Google Scholar] [CrossRef]
  8. Lloyd, J.R.; Byrne, J.M.; Coker, V.S. Biotechnological synthesis of functional nanomaterials. Curr. Opin. Biotechnol. 2011, 22, 509–515. [Google Scholar] [CrossRef]
  9. Díez-Pascual, A.M. Hot Topics in Macromolecular Science. Macromol 2021, 1, 173–176. [Google Scholar] [CrossRef]
  10. Díez-Pascual, A.M. Nanoparticle reinforced polymers. Polymers 2019, 11, 625. [Google Scholar] [CrossRef] [Green Version]
  11. Kobayashi, K.; Wei, J.; Iida, R.; Ijiro, K.; Niikura, K. Surface engineering of nanoparticles for therapeutic applications. Polym. J. 2014, 46, 460–468. [Google Scholar] [CrossRef]
  12. Sanità, G.; Carrese, B.; Lamberti, A. Nanoparticle Surface Functionalization: How to Improve Biocompatibility and Cellular Internalization. Front. Mol. Biosci. 2020, 7, 587012. [Google Scholar] [CrossRef] [PubMed]
  13. Díez-Pascual, A.M. Effect of Graphene Oxide on the Properties of Poly(3-Hydroxybutyrate-co-3-Hydroxyhexanoate. Polymers 2021, 13, 2233. [Google Scholar] [CrossRef] [PubMed]
  14. Razzaq, S.; Rauf, A.; Raza, A.; Akhtar, S.; Tabish, T.A.; Sandhu, M.A.; Zaman, M.; Ibrahim, I.M.; Shahnaz, G.; Rahdar, A.; et al. Multifunctional Polymeric Micelle for Targeted Delivery of Paclitaxel by the Inhibition of the P-Glycoprotein Transporters. Nanomaterials 2021, 11, 2858. [Google Scholar] [CrossRef]
  15. Rahdar, A.; Hasanein, P.; Bilal, M.; Beyzaei, H.; Kyzas, G.Z. Quercetin-loaded F127 nanomicelles: Antioxidant activity and protection against renal injury induced by gentamicin in rats. Life Sci. 2021, 276, 119420. [Google Scholar] [CrossRef]
  16. Er, S.; Laraib, U.; Arshad, R.; Sargazi, S.; Rahdar, A.; Pandey, S.; Thakur, V.K.; Díez-Pascual, A.M. Amino Acids, Peptides, and Proteins: Implications for Nanotechnological Applications in Biosensing and Drug/Gene Delivery. Nanomaterials 2021, 11, 3002. [Google Scholar] [CrossRef]
  17. Mrówczyński, R.; Grześkowiak, B.F. Biomimetic Catechol-Based Nanomaterials for Combined Anticancer Therapies. Nanoeng. Biomater. Biomed. Appl. 2022, 2, 145–180. [Google Scholar] [CrossRef]
  18. Jamir, M.; Islam, R.; Pandey, L.M.; Borah, J. Effect of surface functionalization on the heating efficiency of magnetite nanoclusters for hyperthermia application. J. Alloy. Compd. 2021, 854, 157248. [Google Scholar] [CrossRef]
  19. Nejati, K.; Dadashpour, M.; Gharibi, T.; Mellatyar, H.; Akbarzadeh, A. Biomedical Applications of Functionalized Gold Nanoparticles: A Review. J. Clust. Sci. 2021, 33, 1–16. [Google Scholar] [CrossRef]
  20. Sonju, J.J.; Dahal, A.; Singh, S.S.; Jois, S.D. Peptide-functionalized liposomes as therapeutic and diagnostic tools for cancer treatment. J. Control. Release 2021, 329, 624–644. [Google Scholar] [CrossRef]
  21. Jazayeri, M.H.; Amani, H.; Pourfatollah, A.A.; Pazoki-Toroudi, H.; Sedighimoghaddam, B. Various methods of gold nanoparticles (GNPs) conjugation to antibodies. Sens. Bio-Sens. Res. 2016, 9, 17–22. [Google Scholar] [CrossRef] [Green Version]
  22. Li, H.; Wang, Q.; Liang, G. Phase Transfer of Hydrophobic Nanoparticles Functionalized with Zwitterionic Bisphosphonate Ligands for Renal-Clearable Imaging Nanoprobes. ACS Appl. Nano Mater. 2021, 4, 2621–2633. [Google Scholar] [CrossRef]
  23. Karthik, V.; Selvakumar, P.; Kumar, P.S.; Vo, D.-V.N.; Gokulakrishnan, M.; Keerthana, P.; Elakkiya, V.T.; Rajeswari, R. Graphene-based materials for environmental applications: A review. Environ. Chem. Lett. 2021, 19, 3631–3644. [Google Scholar] [CrossRef]
  24. Díez-Pascual, A.M.; Diez-Vicente, A. L Antibacterial SnO2 nanorods as efficient fillers of poly(propylene fumarate-co-ethylene glycol) biomaterials. Mater. Sci. Eng. C 2017, 78, 806–816. [Google Scholar] [CrossRef]
  25. Alshamrani, M. Broad-Spectrum Theranostics and Biomedical Application of Functionalized Nanomaterials. Polymers 2022, 14, 1221. [Google Scholar] [CrossRef]
  26. Xia, Q.; Huang, J.; Feng, Q.; Chen, X.; Liu, X.; Li, X.; Zhang, T.; Xiao, S.; Li, H.; Zhong, Z.; et al. Size- and cell type-dependent cellular uptake, cytotoxicity and in vivo distribution of gold nanoparticles. Int. J. Nanomed. 2019, 14, 6957–6970. [Google Scholar] [CrossRef] [Green Version]
  27. Hirayama, H.; Amolegbe, S.A.; Islam, M.S.; Rahman, M.A.; Goto, N.; Sekine, Y.; Hayami, S. Encapsulation and controlled release of an antimalarial drug using surface functionalized mesoporous silica nanocarriers. J. Mater. Chem. B 2021, 9, 5043–5046. [Google Scholar] [CrossRef]
  28. Guadagnini, R.; Halamoda Kenzaoui, B.; Walker, L.; Pojana, G.; Magdolenova, Z.; Bilanicova, D.; Saunders, M.; Juillerat-Jeanneret, L.; Marcomini, A.; Huk, A. Toxicity screenings of nanomaterials: Challenges due to interference with assay processes and components of classic in vitro tests. Nanotoxicology 2015, 9, 13–24. [Google Scholar] [CrossRef]
  29. Ellah, N.A.; Abouelmagd, S. Surface functionalization of polymeric nanoparticles for tumor drug delivery: Approaches and challenges. Expert Opin. Drug Deliv. 2017, 14, 201–214. [Google Scholar] [CrossRef]
  30. Fernandes, M.A.; Eloy, J.O.; Luiz, M.T.; Junior, S.L.R.; Borges, J.C.; De la Fuente, L.R.; Luis, C.O.-D.S.; Marchetti, J.M.; Santos-Martinez, M.J.; Chorilli, M. Transferrin-functionalized liposomes for docetaxel delivery to prostate cancer cells. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125806. [Google Scholar] [CrossRef]
  31. Rong, L.; Qin, S.-Y.; Zhang, C.; Cheng, Y.-J.; Feng, J.; Wang, S.-B.; Zhang, X.-Z. Biomedical applications of functional peptides in nano-systems. Mater. Today Chem. 2018, 9, 91–102. [Google Scholar] [CrossRef]
  32. Xie, S.; Ai, L.; Cui, C.; Fu, T.; Cheng, X.; Qu, F.; Tan, W. Functional Aptamer-Embedded Nanomaterials for Diagnostics and Therapeutics. ACS Appl. Mater. Interfaces 2021, 13, 9542–9560. [Google Scholar] [CrossRef] [PubMed]
  33. Farahavar, G.; Abolmaali, S.S.; Gholijani, N.; Nejatollahi, F. Antibody-guided nanomedicines as novel breakthrough therapeutic, diagnostic and theranostic tools. Biomater. Sci. 2019, 7, 4000–4016. [Google Scholar] [CrossRef] [PubMed]
  34. Mout, R.; Moyano, D.F.; Rana, S.; Rotello, V.M. Surface functionalization of nanoparticles for nanomedicine. Chem. Soc. Rev. 2012, 41, 2539–2544. [Google Scholar] [CrossRef]
  35. Mariadoss, A.V.A.; Saravanakumar, K.; Sathiyaseelan, A.; Venkatachalam, K.; Wang, M.-H. Folic acid functionalized starch encapsulated green synthesized copper oxide nanoparticles for targeted drug delivery in breast cancer therapy. Int. J. Biol. Macromol. 2020, 164, 2073–2084. [Google Scholar] [CrossRef]
  36. Wei, W.; Zhang, X.; Zhang, S.; Wei, G.; Su, Z. Biomedical and bioactive engineered nanomaterials for targeted tumor photothermal therapy: A review. Mater. Sci. Eng. C 2019, 104, 109891. [Google Scholar] [CrossRef]
  37. Azevedo, C.; Macedo, M.H.; Sarmento, B. Strategies for the enhanced intracellular delivery of nanomaterials. Drug Discov. Today 2018, 23, 944–959. [Google Scholar] [CrossRef]
  38. Farzin, L.; Shamsipur, M.; Samandari, L.; Sheibani, S. Advances in the design of nanomaterial-based electrochemical affinity and enzymatic biosensors for metabolic biomarkers: A review. Microchim. Acta 2018, 185, 1–25. [Google Scholar] [CrossRef]
  39. Gravely, M.; Safaee, M.M.; Roxbury, D. Biomolecular Functionalization of a Nanomaterial To Control Stability and Retention within Live Cells. Nano Lett. 2019, 19, 6203–6212. [Google Scholar] [CrossRef]
  40. Montaseri, H.; Kruger, C.A.; Abrahamse, H. Review: Organic nanoparticle based active targeting for photodynamic therapy treatment of breast cancer cells. Oncotarget 2020, 11, 2120–2136. [Google Scholar] [CrossRef]
  41. Gole, B.; Sanyal, U.; Banerjee, R.; Mukherjee, P.S. High Loading of Pd Nanoparticles by Interior Functionalization of MOFs for Heterogeneous Catalysis. Inorg. Chem. 2016, 55, 2345–2354. [Google Scholar] [CrossRef] [PubMed]
  42. Bertella, S.; Luterbacher, J.S. Lignin Functionalization for the Production of Novel Materials. Trends Chem. 2020, 2, 440–453. [Google Scholar] [CrossRef]
  43. Díez-Pascual, A.M. Chemical Functionalization of Carbon Nanotubes with Polymers: A Brief Overview. Macromol 2021, 1, 64–83. [Google Scholar] [CrossRef]
  44. Díez-Pascual, A.M. Carbon-Based Nanomaterials. Int. J. Mol. Sci. 2021, 22, 7726. [Google Scholar] [CrossRef] [PubMed]
  45. Sainz-Urruela, C.; Vera-López, S.; Andrés, M.P.S.; Díez-Pascual, A.M. Surface functionalization of graphene oxide with tannic acid: Covalent vs non-covalent approaches. J. Mol. Liq. 2022, 357, 119104. [Google Scholar] [CrossRef]
  46. Díez-Pascual, A.M. Development of Graphene-Based Polymeric Nanocomposites: A Brief Overview. Polymers 2021, 13, 2978. [Google Scholar] [CrossRef]
  47. Chen, Q.; Liu, Z. Albumin carriers for cancer theranostics: A conventionalplatform with new promise. Adv. Mater. 2016, 28, 10557–10566. [Google Scholar] [CrossRef]
  48. Bolaños, K.; Kogan, M.J.; Araya, E. Capping gold nanoparticles with albumin to improve their biomedical properties. Int. J. Nanomed. 2019, 14, 6387–6406. [Google Scholar] [CrossRef] [Green Version]
  49. Chakraborty, A.; Dhar, P. A review on potential of proteins as an excipient for developing a nano-carrier delivery system. Crit. Rev. Ther. Drug Carr. Syst. 2017, 34, 453–488. [Google Scholar] [CrossRef]
  50. Díez-Pascual, A.M.; García-García, D.; Andrés, M.P.S.; Vera, S. Determination of riboflavin based on fluorescence quenching by graphene dispersions in polyethylene glycol. RSC Adv. 2016, 6, 1968. [Google Scholar] [CrossRef]
  51. Díez-Pascual, A.M.; Díez-Vicente, A.L. Multifunctional poly(glycolic acid-co-propylene fumarate) electrospun fibers reinforced with graphene oxide and hydroxyapatite nanorods. J. Mater. Chem. B 2017, 5, 4084–4096. [Google Scholar] [CrossRef] [PubMed]
  52. Horo, H.; Bhattacharyya, S.; Mandal, B.; Kundu, L.M. Synthesis of functionalized silk-coated chitosan-gold nanoparticles and microparticles for target-directed delivery of antitumor agents. Carbohydr. Polym. 2021, 258, 117659. [Google Scholar] [CrossRef] [PubMed]
  53. Díez-Pascual, A.M.; Díez-Vicente, A.L. Electrospun fibers of chitosan-grafted polycaprolactone/poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) blends. J. Mater. Chem. B 2016, 4, 600–612. [Google Scholar] [CrossRef] [PubMed]
  54. Diez-Pascual, A.M.; Diez-Vicente, A.L. Poly(propylene fumarate)/Polyethylene Glycol-Modified Graphene Oxide Nanocomposites for Tissue Engineering. ACS Appl. Mater. Interfaces 2016, 8, 17902–17914. [Google Scholar] [CrossRef] [PubMed]
  55. Kleinfeldt, L.; Gädke, J.; Biedendieck, R.; Krull, R.; Garnweitner, G. Spray-Dried Hierarchical Aggregates of Iron Oxide Nanoparticles and Their Functionalization for Downstream Processing in Biotechnology. ACS Omega 2019, 4, 16300–16308. [Google Scholar] [CrossRef] [Green Version]
  56. Makvandi, P.; Wang, C.Y.; Zare, E.N.; Borzacchiello, A.; Niu, L.N.; Tay, F.R. Metal-Based Nanomaterials in Biomedical Applications: Antimicrobial Activity and Cytotoxicity Aspects. Adv. Funct. Mater. 2020, 30, 1910021. [Google Scholar] [CrossRef]
  57. Mazzotta, E.; Orlando, C.; Muzzalupo, R. New Nanomaterials with Intrinsic Antioxidant Activity by Surface Functionalization of Niosomes with Natural Phenolic Acids. Pharmaceutics 2021, 13, 766. [Google Scholar] [CrossRef]
  58. Delfi, M.; Ghomi, M.; Zarrabi, A.; Mohammadinejad, R.; Taraghdari, Z.; Ashrafizadeh, M.; Zare, E.; Agarwal, T.; Padil, V.; Mokhtari, B.; et al. Functionalization of Polymers and Nanomaterials for Biomedical Applications: Antimicrobial Platforms and Drug Carriers. Prosthesis 2020, 2, 117–139. [Google Scholar] [CrossRef]
  59. Gary-Bobo, M.; Hocine, O.; Brevet, D.; Maynadier, M.; Raehm, L.; Richeter, S.; Charasson, V.; Loock, B.; Morère, A.; Maillard, P.; et al. Cancer therapy improvement with mesoporous silica nanoparticles combining targeting, drug delivery and PDT. Int. J. Pharm. 2012, 423, 509–515. [Google Scholar] [CrossRef]
  60. Jose, J.; Kumar, R.; Harilal, S.; Mathew, G.E.; Parambi, D.G.T.; Prabhu, A.; Uddin, M.S.; Aleya, L.; Kim, H.; Mathew, B. Magnetic nanoparticles for hyperthermia in cancer treatment: An emerging tool. Environ. Sci. Pollut. Res. 2020, 27, 19214–19225. [Google Scholar] [CrossRef]
  61. Sharmeen, S.; Rahman, A.M.; Lubna, M.M.; Salem, K.S.; Islam, R.; Khan, M.A. Polyethylene glycol functionalized carbon nanotubes/gelatin-chitosan nanocomposite: An approach for significant drug release. Bioact. Mater. 2018, 3, 236–244. [Google Scholar] [CrossRef]
  62. Muhammad, M.; Shao, C.S.; Huang, Q. Aptamer-functionalized Au nanoparticles array as the effective SERS biosensor for label-free detection of interleukin-6 in serum. Sens. Actuators B Chem. 2021, 334, 129607. [Google Scholar] [CrossRef]
  63. Mahmoudpour, M.; Ding, S.; Lyu, Z.; Ebrahimi, G.; Du, D.; Dolatabadi, J.E.N.; Torbati, M.; Lin, Y. Aptamer functionalized nanomaterials for biomedical applications: Recent advances and new horizons. Nano Today 2021, 39, 101177. [Google Scholar] [CrossRef]
  64. Mahmoudpour, M.; Karimzadeh, Z.; Ebrahimi, G.; Hasanzadeh, M.; Ezzati Nazhad Dolatabadi, J. Synergizing Functional Nanomaterials with Aptamers Based on Electrochemical Strategies for Pesticide Detection: Current Status and Perspectives. Crit. Rev. Anal. Chem. 2021, 1–28. [Google Scholar] [CrossRef] [PubMed]
  65. Hassanisaadi, M.; Bonjar GH, S.; Rahdar, A.; Pandey, S.; Hosseinipour, A.; Abdolshahi, R. Environmentally Safe Biosynthesis of Gold Nanoparticles Using Plant Water Extracts. Nanomaterials 2021, 11, 2033. [Google Scholar] [CrossRef] [PubMed]
  66. Simon, S.; Ciceo-Lucacel, R.; Radu, T.; Baia, L.; Ponta, O.; Iepure, A.; Simon, V. Gold nanoparticles developed in sol–gel derived apatite—bioactive glass composites. J. Mater. Sci. Mater. Med. 2012, 23, 1193–1201. [Google Scholar] [CrossRef]
  67. Qingling, F.; Wei, J.; Aifantis, K.E.; Fan, Y.; Feng, Q.; Cui, F.-Z.; Watari, F. Current investigations into magnetic nanoparticles for biomedical applications. J. Biomed. Mater. Res. Part A 2016, 104, 1285–1296. [Google Scholar] [CrossRef]
  68. Liu, S.; Höldrich, M.; Sievers-Engler, A.; Horak, J.; Lämmerhofer, M. Papain-functionalized gold nanoparticles as heterogeneous biocatalyst for bioanalysis and biopharmaceuticals analysis. Anal. Chim. Acta 2017, 963, 33–43. [Google Scholar] [CrossRef]
  69. Pourjavadi, A.; Bagherifard, M.; Doroudian, M. Synthesis of micelles based on chitosan functionalized with gold nanorods as a light sensitive drug delivery vehicle. Int. J. Biol. Macromol. 2020, 149, 809–818. [Google Scholar] [CrossRef]
  70. Tiwari, P.M.; Vig, K.; Dennis, V.A.; Singh, S.R. Functionalized gold nanoparticles and their biomedical applications. Nanomaterials 2011, 1, 31–63. [Google Scholar] [CrossRef]
  71. Donoso-González, O.; Lodeiro, L.; Aliaga, Á.E.; Laguna-Bercero, M.A.; Bollo, S.; Kogan, M.J.; Yutronic, N.; Sierpe, R. Functionalization of gold nanostars with cationic β-cyclodextrin-based polymer for drug co-loading and SERS monitoring. Pharmaceutics 2021, 13, 261. [Google Scholar] [CrossRef] [PubMed]
  72. Shon, Y.S.; Choo, H. [60]Fullerene-linked gold nanoparticles: Synthesis and layer-by-layer growth on a solid surface. Chem. Commun. 2002, 21, 2560–2561. [Google Scholar] [CrossRef]
  73. Sudeep, P.K.; Ipe, B.I.; Thomas, K.G.; George, M.V.; Barazzouk, S.; Hotchandani, S.; Kamat, P.V. Fullerene-functionalized gold nanoparticles. A self-assembled photoactive antenna-metal nanocore assembly. Nano Lett. 2002, 2, 29–35. [Google Scholar] [CrossRef]
  74. Yaseen, M.; Humayun, M.; Khan, A.; Usman, M.; Ullah, H.; Tahir, A.; Ullah, H. Preparation, functionalization, modification, and applications of nanostructured gold: A critical review. Energies 2021, 14, 1278. [Google Scholar] [CrossRef]
  75. Liz-Marzán, L.M.; Giersig, M.; Mulvaney, P. Synthesis of nanosized gold−silica core−shell particles. Langmuir 1996, 12, 4329–4335. [Google Scholar] [CrossRef]
  76. Bhargava, A.; Dev, A.; Mohanbhai, S.J.; Pareek, V.; Jain, N.; Choudhury, S.R.; Panwar, J.; Karmakar, S. Pre-coating of protein modulate patterns of corona formation, physiological stability and cytotoxicity of silver nanoparticles. Sci. Total Environ. 2021, 772, 144797. [Google Scholar] [CrossRef]
  77. Matsuo, T. Functionalization of Ruthenium Olefin-Metathesis Catalysts for Interdisciplinary Studies in Chemistry and Biology. Catalysts 2021, 11, 359. [Google Scholar] [CrossRef]
  78. Díez-Pascual, A.M.; Díez-Vicente, A.L. Nano-TiO2 Reinforced PEEK/PEI Blends as Biomaterials for Load-Bearing Implant Applications. ACS Appl. Mater. Interfaces 2015, 7, 5561–5573. [Google Scholar] [CrossRef]
  79. Díez-Pascual, A.M.; Díez-Vicente, A.L. Effect of TiO2 Nanoparticles on the Performance of Polyphenysulfone Biomaterial for Orthopaedic Implants. J. Mater. Chem. B 2014, 2, 7502–7514. [Google Scholar] [CrossRef]
  80. Díez-Pascual, A.M.; Diez-Vicente, A.L. High-Performance Aminated Poly(phenylene sulfide)/ZnO Nanocomposites for Medical Applications. ACS Appl. Mater. Interfaces 2014, 6, 10132–101045. [Google Scholar] [CrossRef] [Green Version]
  81. Díez-Pascual, A.M.; Diez-Vicente, A.L. Antimicrobial and sustainable food packaging based on poly(butylene adipate-co-terephthalate) and electrospun chitosan nanofibers. RSC Adv. 2015, 5, 93095. [Google Scholar] [CrossRef]
  82. Jalalian, S.H.; Taghdisi, S.M.; Hamedani, N.S.; Kalat, S.A.M.; Lavaee, P.; ZandKarimi, M.; Ghows, N.; Jaafari, M.R.; Naghibi, S.; Danesh, N.M.; et al. Epirubicin loaded super paramagnetic iron oxide nanoparticle-aptamer bioconjugate for combined colon cancer therapy and imaging in vivo. Eur. J. Pharm. Sci. 2013, 50, 191–197. [Google Scholar] [CrossRef] [PubMed]
  83. Wei, R.; Xu, Y.; Xue, M. Hollow iron oxide nanomaterials: Synthesis, functionalization, and biomedical applications. J. Mater. Chem. B 2021, 9, 1965–1979. [Google Scholar] [CrossRef] [PubMed]
  84. Cheah, P.; Brown, P.; Qu, J.; Tian, B.; Patton, D.L.; Zhao, Y. Versatile Surface Functionalization of Water-Dispersible Iron Oxide Nanoparticles with Precisely Controlled Sizes. Langmuir 2021, 37, 1279–1287. [Google Scholar] [CrossRef] [PubMed]
  85. Nayeem, J.; Al-Bari, A.A.; Mahiuddin; Rahman, A.; Mefford, O.T.; Ahmad, H.; Rahman, M. Silica coating of iron oxide magnetic nanoparticles by reverse microemulsion method and their functionalization with cationic polymer P(NIPAm-co-AMPTMA) for antibacterial vancomycin immobilization. Colloids Surf. A Physicochem. Eng. Asp. 2021, 611, 125857. [Google Scholar] [CrossRef]
  86. Sun, C.; Lee, J.S.H.; Zhang, M. Magnetic nanoparticles in MR imaging and drug delivery. Adv. Drug Deliv. Rev. 2008, 60, 1252–1265. [Google Scholar] [CrossRef] [Green Version]
  87. Díez-Pascual, A.M.; Díez-Vicente, A.L. Development of Linseed Oil/TiO2 Green Nanocomposites as Antimicrobial Coatings. J. Mater. Chem. B 2015, 3, 4458–4471. [Google Scholar] [CrossRef]
  88. Kundu, M.; Sadhukhan, P.; Ghosh, N.; Chatterjee, S.; Manna, P.; Das, J.; Sil, P.C. pH-responsive and targeted delivery of curcumin via phenylboronic acid-functionalized ZnO nanoparticles for breast cancer therapy. J. Adv. Res. 2019, 18, 161–172. [Google Scholar] [CrossRef]
  89. Al-Harbi, N.; Mohammed, H.; Al-Hadeethi, Y.; Bakry, A.S.; Umar, A.; Hussein, M.A.; Abbassy, M.A.; Vaidya, K.G.; Berakdar, G.A.; Mkawi, E.M.; et al. Silica-Based Bioactive Glasses and Their Applications in Hard Tissue Regeneration: A Review. Pharmaceuticals 2021, 20, 75. [Google Scholar] [CrossRef]
  90. Filho, O.P.; La Torre, G.P.; Hench, L.L. Effect of crystallization on apatite-layer formation of bioactive glass 45S5. J. Biomed. Mater. Res. Off. J. Soc. Biomater. Jpn. Soc. Biomater. 1996, 30, 509–514. [Google Scholar] [CrossRef]
  91. Aina, V.; Cerrato, G.; Martra, G.; Bergandi, L.; Costamagna, C.; Ghigo, D.; Malavasi, G.; Lusvardi, G.; Menabue, L. Gold-containing bioactive glasses: A solid-state synthesis to produce alternative biomaterials for bone implantations. J. R. Soc. Interface 2013, 10, 20121040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Naffakh, M.; Diez-Pascual, A.M. Thermoplastic Polymer Nanocomposites Based on Inorganic Fullerene-like Nanoparticles and Inorganic Nanotubes. Inorganics 2014, 2, 291–312. [Google Scholar] [CrossRef] [Green Version]
  93. Gibson, I.R.; Bonfield, W. Novel synthesis and characterization of an AB-type carbonate-substituted hydroxyapatite. J. Biomed. Mater. Res. Off. J. Soc. Biomater. Jpn. Soc. Biomater. Aust. Soc. Biomater. Korean Soc. Biomater. 2002, 59, 697–708. [Google Scholar] [CrossRef] [PubMed]
  94. Ursino, H.L.; James, B.D.; Ludtka, C.M.; Allen, J.B. Bone tissue engineering. In Tissue Engineering Using Ceramics and Polymers; Elsevier: Amsterdam, The Netherlands, 2022; pp. 587–644. [Google Scholar]
  95. Díez-Pascual, A.M.; Rahdar, A. LbL Nano-Assemblies: A Versatile Tool for Biomedical and Healthcare Applications. Nanomaterials 2022, 12, 949. [Google Scholar] [CrossRef] [PubMed]
  96. Díez-Pascual, A.M. State of the Art in the Antibacterial and Antiviral Applications of Carbon-Based Polymeric Nanocomposites. Int. J. Mol. Sci. 2021, 22, 10511. [Google Scholar] [CrossRef]
  97. Schnorr, J.M.; Swager, T.M. Emerging applications of carbon nanotubes. Chem. Mater. 2011, 23, 646–657. [Google Scholar] [CrossRef] [Green Version]
  98. Naffakh, M.; Díez-Pascual, A.M.; Gómez-Fatou, M.A. New hybrid nanocomposites containing carbon nanotubes, inorganic fullerene-like WS2 nanoparticles and poly(ether ether ketone) (PEEK). J. Mater. Chem. 2011, 21, 7425. [Google Scholar] [CrossRef]
  99. Díez-Pascual, A.M.; Martínez, G.; González-Domínguez, J.M.; Ansón, A.; Martínez, M.T.; Gómez, M.A. Grafting of a hydroxylated poly(ether ether ketone) to the surface of single-walled carbon nanotubes. J. Mater. Chem. 2020, 20, 8285. [Google Scholar] [CrossRef]
  100. Dou, J.; Gan, D.; Huang, Q.; Liu, M.; Chen, J.; Deng, F.; Zhu, X.; Wen, Y.; Zhang, X.; Wei, Y. Functionalization of carbon nanotubes with chitosan based on MALI multicomponent reaction for Cu2+ removal. Int. J. Biol. Macromol. 2019, 136, 476–485. [Google Scholar] [CrossRef]
  101. Sainz-Urruela, C.; Vera-López, S.; San Andrés, M.P.; Díez-Pascual, A.M. Graphene-Based Sensors for the Detection of Bioactive Compounds: A Review. Int. J. Mol. Sci. 2021, 22, 3316. [Google Scholar] [CrossRef]
  102. Díez-Pascual, A.M. Antibacterial Action of Nanoparticle Loaded Nanocomposites Based on Graphene and Its Derivatives: A Mini-Review. Int. J. Mol. Sci. 2020, 21, 3563. [Google Scholar] [CrossRef] [PubMed]
  103. Innocenzi, P.; Stagi, L. Carbon-based antiviral nanomaterials: Graphene, C-dots, and fullerenes. A perspective. Chem. Sci. 2020, 11, 6606–6622. [Google Scholar] [CrossRef] [PubMed]
  104. Xin, Q.; Shah, H.; Nawaz, A.; Xie, W.; Akram, M.Z.; Batool, A.; Tian, L.; Jan, S.U.; Boddula, R.; Guo, B.; et al. Antibacterial carbon-based nanomaterials. Adv. Mater. 2019, 31, e1804838. [Google Scholar] [CrossRef] [PubMed]
  105. Hu, X.; Mu, L.; Wen, J.; Zhou, Q. Covalently synthesized graphene oxide-aptamer nanosheets for efficient visible-light photocatalysis of nucleic acids and proteins of viruses. Carbon 2012, 50, 2772–2781. [Google Scholar] [CrossRef]
  106. Van Tam, T.; Hur, S.H.; Chung, J.S.; Choi, W.M. Novel paper- and fiber optic-based fluorescent sensor for glucose detection using aniline-functionalized graphene quantum dots. Sens. Actuators B Chem. 2021, 329, 129250. [Google Scholar] [CrossRef]
  107. Seifi, T.; Kamali, A.R. Antiviral performance of graphene-based materials with emphasis on COVID-19: A review. Med. Drug Discov. 2021, 11, 100099. [Google Scholar] [CrossRef]
  108. Bai, J.; Chen, L.; Zhu, Y.; Wang, X.; Wu, X.; Fu, Y. A novel luminescence sensor based on porous molecularly imprinted polymer-ZnS quantum dots for selective recognition of paclitaxel. Colloids Surf. A Physicochem. Eng. Asp. 2020, 610, 125696. [Google Scholar] [CrossRef]
  109. Soleymani, J.; Hasanzadeh, M.; Somi, M.H.; Ozkan, S.A.; Jouyban, A. Targeting and sensing of some cancer cells using folate bioreceptor functionalized nitrogen-doped graphene quantum dots. Int. J. Biol. Macromol. 2018, 118, 1021–1034. [Google Scholar] [CrossRef]
  110. Banerjee, A.; Pons, T.; Lequeux, N.; Dubertret, B. Quantum dots–DNA bioconjugates: Synthesis to applications. Interface Focus 2016, 6, 20160064. [Google Scholar] [CrossRef] [Green Version]
  111. Sun, D.; Gang, O. DNA-Functionalized Quantum Dots: Fabrication, Structural, and Physicochemical Properties. Langmuir 2013, 29, 7038–7046. [Google Scholar] [CrossRef]
  112. Wang, G.; Li, Z.; Luo, X.; Yue, R.; Shen, Y.; Ma, N. DNA-templated nanoparticle complexes for photothermal imaging and labeling of cancer cells. Nanoscale 2018, 10, 16508–16520. [Google Scholar] [CrossRef] [PubMed]
  113. Hajikarimi, Z.; Khoei, S.; Khoee, S.; Mahdavi, S.R. Evaluation of the cytotoxic effects of PLGA coated iron oxide nanoparticles as a carrier of 5-fluorouracil and mega-voltage X-ray radiation in DU145 prostate cancer cell line. IEEE Trans. Nanobioscience 2014, 13, 403–408. [Google Scholar] [CrossRef] [PubMed]
  114. Thamake, S.I.; Raut, S.2.; Ranjan, A.P.; Gryczynski, Z.; Vishwanatha, J.K. Surface functionalization of PLGA nanoparticles by non-covalent insertion of a homo-bifunctional spacer for active targeting in cancer therapy. Nanotechnology 2010, 22, 035101. [Google Scholar] [CrossRef] [PubMed]
  115. Du, H.; Parit, M.; Liu, K.; Zhang, M.; Jiang, Z.; Huang, T.-S.; Zhang, X.; Si, C. Multifunctional Cellulose Nanopaper with Superior Water-Resistant, Conductive, and Antibacterial Properties Functionalized with Chitosan and Polypyrrole. ACS Appl. Mater. Interfaces 2021, 13, 32115–32125. [Google Scholar] [CrossRef] [PubMed]
  116. Sofla, R.L.M.; Rezaei, M.; Babaie, A. Investigation of the effect of graphene oxide functionalization on the physical, mechanical and shape memory properties of polyurethane/reduced graphene oxide nanocomposites. Diam. Relat. Mater. 2019, 95, 195–205. [Google Scholar] [CrossRef]
  117. Yan, S.; Wang, W.; Li, X.; Ren, J.; Yun, W.; Zhang, K.; Li, G.; Yin, J. Preparation of mussel-inspired injectable hydrogels based on dual-functionalized alginate with improved adhesive, self-healing, and mechanical properties. J. Mater. Chem. B 2018, 6, 6377–6390. [Google Scholar] [CrossRef] [PubMed]
  118. Saifi, M.A.; Khan, W.; Godugu, C. Cytotoxicity of Nanomaterials: Using Nanotoxicology to Address the Safety Concerns of Nanoparticles. Pharm. Nanotechnol. 2018, 6, 3–16. [Google Scholar] [CrossRef]
  119. Srivastava, V.; Gusain, D.; Sharma, Y.C. Critical Review on the Toxicity of Some Widely Used Engineered Nanoparticles. Ind. Eng. Chem. Res. 2015, 54, 6209–6233. [Google Scholar] [CrossRef]
  120. Madani, S.Y.; Mandel, A.; Seifalian, A.M. A concise review of carbon nanotube’s toxicology. Nano Rev. 2013, 4, 21521. [Google Scholar] [CrossRef] [Green Version]
  121. Yang, H.; Liu, C.; Yang, D.; Zhang, H.; Xi, Z. Comparative study of cytotoxicity, oxidative stress and genotoxicity induced by four typical nanomaterials: The role of particle size, shape and composition. J. Appl. Toxicol. 2009, 29, 69–78. [Google Scholar] [CrossRef]
  122. Katsumiti, A.; Berhanu, D.; Howard, K.T.; Arostegui, I.; Oron, M.; Reip, P.; Valsami-Jones, E.; Cajaraville, M. Cytotoxicity of TiO2 nanoparticles to mussel hemocytes and gill cells in vitro: Influence of synthesis method, crystalline structure, size and additive. Nanotoxicology 2015, 9, 543–553. [Google Scholar] [CrossRef] [PubMed]
  123. Sato, S.; Nakamura, R.; Abe, S. Visible-light sensitization of TiO2 photocatalysts by wet-method N doping. Appl. Catal. A Gen. 2005, 284, 131–137. [Google Scholar] [CrossRef]
  124. Sukhanova, A.; Bozrova, S.; Sokolov, P.; Berestovoy, M.; Karaulov, A.; Nabiev, I. Dependence of Nanoparticle Toxicity on Their Physical and Chemical Properties. Nanoscale Res. Lett. 2018, 13, 1–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Wang, J.; Yao, H.; Shi, X. Cooperative entry of nanoparticles into the cell. J. Mech. Phys. Solids 2014, 73, 151–165. [Google Scholar] [CrossRef]
  126. Elrahman, A.A.; Mansour, F. Targeted magnetic iron oxide nanoparticles: Preparation, functionalization and biomedical application. J. Drug Deliv. Sci. Technol. 2019, 52, 702–712. [Google Scholar] [CrossRef]
  127. Kheirallah, D.A.M.; El-Samad, L.M.; Abdel-Moneim, A.M. DNA damage and ovarian ultrastructural lesions induced by nickel oxide nano-particles in Blaps polycresta (Coleoptera: Tenebrionidae). Sci. Total Environ. 2021, 753, 141743. [Google Scholar] [CrossRef] [PubMed]
  128. Du, Y.; Jin, J.; Liang, H.; Jiang, W. Structural and physicochemical properties and biocompatibility of linear and looped polymer-capped gold nanoparticles. Langmuir 2019, 35, 8316–8324. [Google Scholar] [CrossRef]
  129. Sen, G.T.; Ozkemahli, G.; Shahbazi, R.; Erkekoglu, P.; Ulubayram, K.; Kocer-Gumusel, B. The effects of polymer coating of gold nanoparticles on oxidative stress and DNA damage. Int. J. Toxicol. 2020, 39, 328–340. [Google Scholar] [CrossRef]
  130. Wang, R.; Bowling, I.; Liu, W. Cost effective surface functionalization of gold nanoparticles with a mixed DNA and PEG monolayer for nanotechnology applications. RSC Adv. 2017, 7, 3676–3679. [Google Scholar] [CrossRef] [Green Version]
  131. Singh, A.; Amiji, M.M. Application of nanotechnology in medical diagnosis and imaging. Curr. Opin. Biotechnol. 2022, 74, 241–246. [Google Scholar] [CrossRef]
  132. El-Sayed, A.; Kamel, M. Advances in nanomedical applications: Diagnostic, therapeutic, immunization, and vaccine production. Environ. Sci. Pollut. Res. 2020, 27, 19200–19213. [Google Scholar] [CrossRef] [PubMed]
  133. Mi, Y.; Shao, Z.; Vang, J.; Kaidar-Person, O.; Wang, A.Z. Application of nanotechnology to cancer radiotherapy. Cancer Nanotechnol. 2016, 7, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Aithal, P.S. Nanotechnology Innovations & Business Opportunities: A Review. Int. J. Manag. IT Eng. 2016, 6, 182–204. [Google Scholar]
  135. Elkodous, M.A.; El-Sayyad, G.S.; Abdelrahman, I.Y.; El-Bastawisy, H.S.; Mohamed, A.E.; Mosallam, F.M.; Nasser, H.; Gobara, M.; Baraka, A.; Elsayed, M.; et al. Therapeutic and diagnostic potential of nanomaterials for enhanced biomedical applications. Colloids Surf. B Biointerfaces 2019, 180, 411–428. [Google Scholar] [CrossRef] [PubMed]
  136. Chen, X.; Song, J.; Chen, X.; Yang, H. X-ray-activated nanosystems for theranostic applications. Chem. Soc. Rev. 2019, 48, 3073–3101. [Google Scholar] [CrossRef] [PubMed]
  137. Gharpure, K.; Wu, S.; Li, C.; Lopez-Berestein, G.; Sood, A.K. Nanotechnology: Future of oncotherapy. Clin. Cancer Res. 2015, 21, 3121–3130. [Google Scholar] [CrossRef] [Green Version]
  138. Chen, Z.; Peng, Y.; Li, Y.; Xie, X.; Wei, X.; Yang, G.; Zhang, H.; Li, N.; Li, T.; Qin, X. Aptamer-Dendrimer Functionalized Magnetic Nano-Octahedrons: Theranostic Drug/Gene Delivery Platform for Near-Infrared/Magnetic Resonance Imaging-Guided Magnetochemotherapy. ACS Nano 2021, 15, 16683–16696. [Google Scholar] [CrossRef]
  139. Zeytunluoglu, A.; Arslan, I. Current perspectives on nanoemulsions in targeted drug delivery: An overview. In Handbook of Research on Nanoemulsion Applications in Agriculture, Food, Health, and Biomedical Sciences; IGI Global: Hershey, PA, USA, 2022; pp. 118–140. [Google Scholar]
  140. Dong, Y.; Wu, X.; Chen, X.; Zhou, P.; Xu, F.; Liang, W. Nanotechnology shaping stem cell therapy: Recent advances, application, challenges, and future outlook. Biomed. Pharmacother. 2021, 137, 111236. [Google Scholar] [CrossRef]
  141. Pereira, M.N.; Ushirobira, C.Y.; Cunha-Filho, M.S.; Gelfuso, G.M.; Gratieri, T. Nanotechnology advances for hair loss. Ther. Deliv. 2018, 9, 593–603. [Google Scholar] [CrossRef]
  142. Kumar, S.; Nehra, M.; Kedia, D.; Dilbaghi, N.; Tankeshwar, K.; Kim, K.-H. Nanotechnology-based biomaterials for orthopaedic applications: Recent advances and future prospects. Mater. Sci. Eng. C 2020, 106, 110154. [Google Scholar] [CrossRef]
  143. Zheng, X.; Zhang, P.; Fu, Z.; Meng, S.; Dai, L.; Yang, H. Applications of nanomaterials in tissue engineering. RSC Adv. 2021, 11, 19041–19058. [Google Scholar] [CrossRef] [PubMed]
  144. Kumar, R.; Aadil, K.R.; Ranjan, S.; Kumar, V.B. Advances in nanotechnology and nanomaterials based strategies for neural tissue engineering. J. Drug Deliv. Sci. Technol. 2020, 57, 101617. [Google Scholar] [CrossRef]
  145. Bakopoulou, A.; Papachristou, E.; Bousnaki, M.; Hadjichristou, C.; Kontonasaki, E.; Theocharidou, A.; Papadopoulou, L.; Kantiranis, N.; Zachariadis, G.; Leyhausen, G. Human treated dentin matrices combined with Zn-doped, Mg-based bioceramic scaffolds and human dental pulp stem cells towards targeted dentin regeneration. Dent. Mater. 2016, 32, e159–e175. [Google Scholar] [CrossRef] [PubMed]
  146. Balakrishnan, B.; Joshi, N.; Jayakrishnan, A.; Banerjee, R. Self-crosslinked oxidized alginate/gelatin hydrogel as injectable, adhesive biomimetic scaffolds for cartilage regeneration. Acta Biomater. 2014, 10, 3650–3663. [Google Scholar] [CrossRef]
  147. Wang, C.; Zhang, H.; Chen, B.; Yin, H.; Wang, W. Study of the enhanced anticancer efficacy of gambogic acid on Capan-1 pancreatic cancer cells when mediated via magnetic Fe3O4 nanoparticles. Int. J. Nanomed. 2011, 6, 1929. [Google Scholar]
  148. Haruta, S.; Hanafusa, T.; Fukase, H.; Miyajima, H.; Oki, T. An effective absorption behavior of insulin for diabetic treatment following intranasal delivery using porous spherical calcium carbonate in monkeys and healthy human volunteers. Diabetes Technol. Ther. 2003, 5, 1–9. [Google Scholar] [CrossRef]
  149. Zazo, H.; Colino, C.I.; Lanao, J.M. Current applications of nanoparticles in infectious diseases. J. Control. Release 2016, 224, 86–102. [Google Scholar] [CrossRef]
  150. Rudramurthy, G.R.; Swamy, M.K.; Sinniah, U.R.; Ghasemzadeh, A. Nanoparticles: Alternatives against drug-resistant pathogenic microbes. Molecules 2016, 21, 836. [Google Scholar] [CrossRef]
  151. Smith, B.R.; Gambhir, S.S. Nanomaterials for in vivo imaging. Chem. Rev. 2017, 117, 901–986. [Google Scholar] [CrossRef]
  152. Imamura, Y.; Yamada, S.; Tsuboi, S.; Nakane, Y.; Tsukasaki, Y.; Komatsuzaki, A.; Jin, T. Near-infrared emitting PbS quantum dots for in vivo fluorescence imaging of the thrombotic state in septic mouse brain. Molecules 2016, 21, 1080. [Google Scholar] [CrossRef] [Green Version]
  153. Wang, Y.; Wu, H.; Lin, D.; Zhang, R.; Li, H.; Zhang, W.; Liu, W.; Huang, S.; Yao, L.; Cheng, J.; et al. One-dimensional electrospun ceramic nanomaterials and their sensing applications. J. Am. Ceram. Soc. 2022, 105, 765–785. [Google Scholar] [CrossRef]
  154. Liu, J.; Wang, Z.; Zhao, S.; Ding, B. Multifunctional nucleic acid nanostructures for gene therapies. Nano Res. 2018, 11, 5017–5027. [Google Scholar] [CrossRef]
  155. Patil-Sen, Y. Advances in nano-biomaterials and their applications in biomedicine. Emerg. Top. Life Sci. 2021, 5, 169–176. [Google Scholar] [CrossRef] [PubMed]
  156. Wu, G.; Li, P.; Feng, H.; Zhang, X.; Chu, P.K. Engineering and functionalization of biomaterials via surface modification. J. Mater. Chem. B 2015, 3, 2024–2042. [Google Scholar] [CrossRef] [PubMed]
  157. Singh, T.V.; Shagolsem, L.S. Biopolymer based nano-structured materials and their applications. In Nanostructured Materials and Their Applications; Springer: Singapore, 2021; pp. 337–366. [Google Scholar]
  158. Sargazi, S.; Mukhtar, M.; Rahdar, A.; Barani, M.; Pandey, S.; Díez-Pascual, A.M. Active Targeted Nanoparticles for Delivery of Poly(ADP-ribose) Polymerase (PARP) Inhibitors: A Preliminary Review. Int. J. Mol. Sci. 2021, 22, 10319. [Google Scholar] [CrossRef] [PubMed]
  159. Sivasankarapillai, V.S.; Das, S.S.; Sabir, F.; Sundaramahalingam, M.A.; Colmenares, J.C.; Prasannakumar, S.; Rajan, M.; Rahdar, A.; Kyzas, G.Z. Progress in natural polymer engineered biomaterials for transdermal drug delivery systems. Mater. Today Chem. 2021, 19, 100382. [Google Scholar] [CrossRef]
  160. Bouchoucha, M.; Gaudreault, R.C.; Fortin, M.A.; Kleitz, F. Mesoporous silica nanoparticles: Selective surface functionalization for optimal relaxometric and drug loading performances. Adv. Funct. Mater. 2014, 24, 5911–5923. [Google Scholar] [CrossRef]
  161. Yoo, J.; Park, C.; Yi, G.; Lee, D.; Koo, H. Active Targeting Strategies Using Biological Ligands for Nanoparticle Drug Delivery Systems. Cancers 2019, 11, 640. [Google Scholar] [CrossRef] [Green Version]
  162. Huang, Y.; Cao, L.; Parakhonskiy, B.V.; Skirtach, A.G. Hard, Soft, and Hard-and-Soft Drug Delivery Carriers Based on CaCO3 and Alginate Biomaterials: Synthesis, Properties, Pharmaceutical Applications. Pharmaceutics 2022, 14, 909. [Google Scholar] [CrossRef]
  163. Dizaj, S.M.; Lotfipour, F.; Barzegar-Jalali, M.; Zarrintan, M.-H.; Adibkia, K. Ciprofloxacin HCl-loaded calcium carbonate nanoparticles: Preparation, solid state characterization, and evaluation of antimicrobial effect against Staphylococcus aureus. Artif. Cells Nanomed. Biotechnol. 2016, 45, 535–543. [Google Scholar] [CrossRef] [Green Version]
  164. Ueno, Y.; Futagawa, H.; Takagi, Y.; Ueno, A.; Mizushima, Y. Drug-incorporating calcium carbonate nanoparticles for a new delivery system. J. Control. Release 2005, 103, 93–98. [Google Scholar] [CrossRef] [PubMed]
  165. Hood, J.D.; Bednarski, M.; Frausto, R.; Guccione, S.; Reisfeld, R.A.; Xiang, R.; Cheresh, D.A. Tumor regression by targeted gene delivery to the neovasculature. Science 2002, 296, 2404–2407. [Google Scholar] [CrossRef] [Green Version]
  166. Deng, L.; Li, Q.; Al-Rehili, S.; Omar, H.; Almalik, A.; Alshamsan, A.; Zhang, J.; Khashab, N.M. Hybrid iron oxide–graphene oxide–polysaccharides microcapsule: A micro-matryoshka for on-demand drug release and antitumor therapy in vivo. ACS Appl. Mater. Interfaces 2016, 8, 6859–6868. [Google Scholar] [CrossRef] [Green Version]
  167. Ye, C.; Combs, Z.A.; Calabrese, R.; Dai, H.; Kaplan, D.L.; Tsukruk, V.V. Robust microcapsules with controlled permeability from silk fibroin reinforced with graphene oxide. Small 2014, 10, 5087–5097. [Google Scholar] [CrossRef] [PubMed]
  168. Yang, P.-H.; Sun, X.; Chiu, J.-F.; Sun, H.; He, Q.-Y. Transferrin-mediated gold nanoparticle cellular uptake. Bioconjugate Chem. 2005, 16, 494–496. [Google Scholar] [CrossRef] [PubMed]
  169. Cheng, J.; Gu, Y.-J.; Cheng, S.H.; Wong, W.-T. Surface functionalized gold nanoparticles for drug delivery. J. Biomed. Nanotechnol. 2013, 9, 1362–1369. [Google Scholar] [CrossRef]
  170. Fang, Z.; Li, X.; Xu, Z.; Du, F.; Wang, W.; Shi, R.; Gao, D. Hyaluronic acid-modified mesoporous silica-coated superparamagnetic Fe3O4 nanoparticles for targeted drug delivery. Int. J. Nanomed. 2019, 14, 5785. [Google Scholar] [CrossRef] [Green Version]
  171. Prabha, G.; Raj, V. Sodium alginate–polyvinyl alcohol–bovin serum albumin coated Fe3O4 nanoparticles as anticancer drug delivery vehicle: Doxorubicin loading and in vitro release study and cytotoxicity to HepG2 and L02 cells. Mater. Sci. Eng. C 2017, 79, 410–422. [Google Scholar] [CrossRef]
  172. Rajan, M.; Raj, V.; Al-Arfaj, A.A.; Murugan, A.M. Hyaluronidase enzyme core-5-fluorouracil-loaded chitosan-PEG-gelatin polymer nanocomposites as targeted and controlled drug delivery vehicles. Int. J. Pharm. 2013, 453, 514–522. [Google Scholar] [CrossRef]
  173. Hua, S.; Ma, H.; Li, X.; Yang, H.; Wang, A. pH-sensitive sodium alginate/poly (vinyl alcohol) hydrogel beads prepared by combined Ca2+ crosslinking and freeze-thawing cycles for controlled release of diclofenac sodium. Int. J. Biol. Macromol. 2010, 46, 517–523. [Google Scholar] [CrossRef]
  174. Arshad, R.; Tabish, T.A.; Kiani, M.H.; Ibrahim, I.M.; Shahnaz, G.; Rahdar, A.; Kan, M.; Pandey, S. Hyaluronic Acid Functionalized Self-Nano-Emulsifying Drug Delivery System (SNEDDS) for Enhancement in Ciprofloxacin Targeted Delivery against Intracellular Infection. Nanomaterials 2021, 11, 1086. [Google Scholar] [CrossRef] [PubMed]
  175. Guo, Q.; Jia, Y.; Yuan, M.; Huang, X.; Sui, X.; Tang, F.; Peng, J.; Chen, J.; Lu, S.; Cui, X.; et al. Co-encapsulation of magnetic Fe3O4 nanoparticles and doxorubicin into biodegradable PLGA nanocarriers for intratumoral drug delivery. Int. J. Nanomed. 2012, 7, 1697–1708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Wang, B.; Xu, C.; Xie, J.; Yang, Z.; Sun, S. pH controlled release of chromone from chromone-Fe3O4 nanoparticles. J. Am. Chem. Soc. 2008, 130, 14436–14437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Xie, J.; Xu, C.; Kohler, N.; Hou, Y.; Sun, S. Controlled PEGylation of monodisperse Fe3O4 nanoparticles for reduced non-specific uptake by macrophage cells. Adv. Mater. 2007, 19, 3163–3166. [Google Scholar] [CrossRef]
  178. Saikia, C.; Hussain, A.; Ramteke, A.; Sharma, H.K.; Maji, T.K. Crosslinked thiolated starch coated Fe3O4 magnetic nanoparticles: Effect of montmorillonite and crosslinking density on drug delivery properties. Starch-Stärke 2014, 66, 760–771. [Google Scholar] [CrossRef]
  179. Rezaei, A.; Hashemi, E. A pseudohomogeneous nanocarrier based on carbon quantum dots decorated with arginine as an efficient gene delivery vehicle. Sci. Rep. 2021, 11, 1–10. [Google Scholar] [CrossRef]
  180. Di Marzio, N.; Eglin, D.; Serra, T.; Moroni, L. Bio-Fabrication: Convergence of 3D Bioprinting and Nano-Biomaterials in Tissue Engineering and Regenerative Medicine. Front. Bioeng. Biotechnol. 2020, 8, 326. [Google Scholar] [CrossRef]
  181. Kapat, K.; Shubhra, Q.T.H.; Zhou, M.; Leeuwenburgh, S. Piezoelectric nano-biomaterials for biomedicine and tissue regeneration. Adv. Funct. Mater. 2020, 30, 1909045. [Google Scholar] [CrossRef] [Green Version]
  182. Díez-Pascual, A.M.; Diez-Vicente, A.L. Epoxidized Soybean Oil/ZnO Biocomposites for Soft Tissue Applications: Preparation and Characterization. ACS Appl. Mater. Interfaces 2014, 6, 17277–17288. [Google Scholar] [CrossRef]
  183. Giubilato, E.; Cazzagon, V.; Amorim, M.J.B.; Blosi, M.; Bouillard, J.; Bouwmeester, H.; Costa, A.L.; Fadeel, B.; Fernandes, T.F.; Fito, C.; et al. Risk management framework for nano-biomaterials used in medical devices and advanced therapy medicinal products. Materials 2020, 13, 4532. [Google Scholar] [CrossRef]
  184. Rana, D.; Ramasamy, K.; Leena, M.; Jiménez, C.; Campos, J.; Ibarra, P.; Haidar, Z.S.; Ramalingam, M. Surface functionalization of nanobiomaterials for application in stem cell culture, tissue engineering, and regenerative medicine. Biotechnol. Prog. 2016, 32, 554–567. [Google Scholar] [CrossRef] [PubMed]
  185. Labusca, L.; Herea, D.-D.; Mashayekhi, K. Stem cells as delivery vehicles for regenerative medicine-challenges and perspectives. World J. Stem Cells 2018, 10, 43–56. [Google Scholar] [CrossRef] [PubMed]
  186. Lyons, J.G.; Plantz, M.A.; Hsu, W.K.; Hsu, E.L.; Minardi, S. Nanostructured biomaterials for bone regeneration. Front. Bioeng. Biotechnol. 2020, 8, 922. [Google Scholar] [CrossRef]
  187. Zhu, L.; Luo, D.; Liu, Y. Effect of the nano/microscale structure of biomaterial scaffolds on bone regeneration. Int. J. Oral Sci. 2020, 12, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Zhang, Y.; Venugopal, J.R.; El-Turki, A.; Ramakrishna, S.; Su, B.; Lim, C.T. Electrospun biomimetic nanocomposite nanofiber of hydroxiapatite/chitosan for bone tissue engineering. Biomaterials 2008, 29, 4314–4322. [Google Scholar] [CrossRef] [PubMed]
  189. Liang, X.Y.; Duan, P.G.; Gao, J.M.; Guo, R.S.; Qu, Z.H.; Li, X.F.; He, Y.; Yao, H.Q.; Ding, J.D. Bilayered PLGA/PLGA-HAp Composite Scaffold for Osteochondral Tissue Engineering and Tissue Regeneration. ACS Biomater. Sci. Eng. 2018, 4, 3506–3521. [Google Scholar] [CrossRef]
  190. Naffakh, M.; Diez-Pascual, A.M. WS2 inorganic nanotubes reinforced poly(L-lacticacid)/hydroxyapatite hybrid composite biomaterials. RSC Adv. 2015, 5, 65514. [Google Scholar] [CrossRef]
  191. Cui, Y.; Li, H.; Li, Y.; Mao, L. Novel insights into nanomaterials for immunomodulatory bone regeneration. Nanoscale Adv. 2022, 4, 334–352. [Google Scholar] [CrossRef]
  192. McMillan, A.; Nguyen, M.K.; Gonzalez-Fernandez, T.; Ge, P.; Yu, X.; Murphy, W.L.; Kelly, D.; Alsberg, E. Dual non-viral gene delivery from microparticles within 3D high-density stem cell constructs for enhanced bone tissue engineering. Biomaterials 2018, 161, 240–255. [Google Scholar] [CrossRef]
  193. Perez, J.R.; Kouroupis, D.; Li, D.J.; Best, T.M.; Kaplan, L.; Correa, D. Tissue engineering and cell-based therapies for fractures and bone defects. Front. Bioeng. Biotechnol. 2018, 6, 105. [Google Scholar] [CrossRef] [Green Version]
  194. Veatch, J.R.; Singhi, N.; Srivastava, S.; Szeto, J.L.; Jesernig, B.; Stull, S.M.; Fitzgibbon, M.; Sarvothama, M.; Yechan-Gunja, S.; James, S.E.; et al. A therapeutic cancer vaccine delivers antigens and adjuvants to lymphoid tissues using genetically modified T cells. J. Clin. Investig. 2021, 131, e144195. [Google Scholar] [CrossRef] [PubMed]
  195. Acri, T.M.; Laird, N.Z.; Jaidev, L.R.; Meyerholz, D.K.; Salem, A.K.; Shin, K. Nonviral Gene Delivery Embedded in Biomimetically Mineralized Matrices for Bone Tissue Engineering. Tissue Eng. Part A 2021, 27, 1074–1083. [Google Scholar] [CrossRef] [PubMed]
  196. Ghandforoushan, P.; Hanaee, J.; Aghazadeh, Z.; Samiei, M.; Navali, A.M.; Khatibi, A.; Davaran, S. Novel nanocomposite scaffold based on gelatin/PLGA-PEG-PLGA hydrogels embedded with TGF-β1 for chondrogenic differentiation of human dental pulp stem cells in vitro. Int. J. Biol. Macromol. 2022, 201, 270–287. [Google Scholar] [CrossRef] [PubMed]
  197. Sirkkunan, D.; Pingguan-Murphy, B.; Muhamad, F. Directing Axonal Growth: A Review on the Fabrication of Fibrous Scaffolds That Promotes the Orientation of Axons. Gels 2022, 8, 25. [Google Scholar] [CrossRef]
  198. Abdal-Hay, A.; Sheikh, F.A.; Gómez-Cerezo, N.; Alneairi, A.; Luqman, M.; Pant, H.R.; Ivanovski, S. A review of protein adsorption and bioactivity characteristics of poly ε-caprolactone scaffolds in regenerative medicine. Eur. Polym. J. 2022, 162, 110892. [Google Scholar] [CrossRef]
  199. Joy, J.; Pereira, J.; Aid-Launais, R.; Pavon-Djavid, G.; Ray, A.R.; Letourneur, D.; Meddahi-Pellé, A.; Gupta, B. Gelatin—Oxidized carboxymethyl cellulose blend based tubular electrospun scaffold for vascular tissue engineering. Int. J. Biol. Macromol. 2018, 107, 1922–1935. [Google Scholar] [CrossRef]
  200. Murugesan, B.; Pandiyan, N.; Arumugam, M.; Sonamuthu, J.; Samayanan, S.; Yurong, C.; Juming, Y.; Mahalingam, S. Fabrication of palladium nanoparticles anchored polypyrrole functionalized reduced graphene oxide nanocomposite for antibiofilm associated orthopedic tissue engineering. Appl. Surf. Sci. 2020, 510, 145403. [Google Scholar] [CrossRef]
  201. Jie, W.; Song, F.; Li, X.; Li, W.; Wang, R.; Jiang, Y.; Zhao, L.; Fan, Z.; Wang, J.; Liu, B. Enhancing the proliferation of MC3T3-E1 cells on casein phosphopeptide-biofunctionalized 3D reduced-graphene oxide/polypyrrole scaffolds. RSC Adv. 2017, 7, 34415–34424. [Google Scholar] [CrossRef] [Green Version]
  202. Díez-Pascual, A.M.; Díez-Vicente, A.L. Wound Healing Bionanocomposites Based on Castor Oil Polymeric Films Reinforced with Chitosan-Modified ZnO Nanoparticles. Biomacromolecules 2015, 16, 2631–2644. [Google Scholar] [CrossRef]
  203. Lee, M.-H.; You, C.; Kim, K.-H. Combined effect of a microporous layer and type I collagen coating on a biphasic calcium phosphate scaffold for bone tissue engineering. Materials 2015, 8, 1150–1161. [Google Scholar] [CrossRef] [Green Version]
  204. Siddiqi, N.J.; Abdelhalim, M.A.K.; El-Ansary, A.K.; Alhomida, A.S.; Ong, W.Y. Identification of potential biomarkers of gold nanoparticle toxicity in rat brains. J. Neuroinflammation 2012, 9, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Abdelhalim, M.A.K. Exposure to gold nanoparticles produces cardiac tissue damage that depends on the size and duration of exposure. Lipids Health Dis. 2011, 10, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Love, S.A.; Thompson, J.W.; Haynes, C.L. Development of screening assays for nanoparticle toxicity assessment in human blood: Preliminary studies with charged Au nanoparticles. Nanomedicine 2012, 7, 1355–1364. [Google Scholar] [CrossRef] [PubMed]
  207. Freese, C.; Gibson, M.I.; Klok, H.A.; Unger, R.E.; Kirkpatrick, C.J. Size-and coating-dependent uptake of polymer-coated gold nanoparticles in primary human dermal microvascular endothelial cells. Biomacromolecules 2012, 13, 1533–1543. [Google Scholar] [CrossRef] [PubMed]
  208. Fu, Q.; Rahaman, M.N.; Bal, B.S.; Brown, R.F.; Day, D.E. Mechanical and in vitro performance of bioactive glass scaffolds prepared by a polymer foam replication technique. Acta Biomater. 2008, 4, 1854–1864. [Google Scholar] [CrossRef]
  209. Chen, Q.Z.; Thompson, I.D.; Boccaccini, A.R. 45S5 Bioglass®-derived glass–ceramic scaffolds for bone tissue engineering. Biomaterials 2006, 27, 2414–2425. [Google Scholar] [CrossRef]
  210. Jiang, L.; Chen, D.; Wang, Z.; Zhang, Z.; Xia, Y.; Xue, H.; Liu, Y. Preparation of an electrically conductive graphene oxide/chitosan scaffold for cardiac tissue engineering. Appl. Biochem. Biotechnol. 2019, 188, 952–964. [Google Scholar] [CrossRef]
  211. Shamekhi, M.A.; Mirzadeh, H.; Mahdavi, H.; Rabiee, A.; Mohebbi-Kalhori, D.; Eslaminejad, M.B. Graphene oxide containing chitosan scaffolds for cartilage tissue engineering. Int. J. Biol. Macromol. 2019, 127, 396–405. [Google Scholar] [CrossRef]
  212. Nishida, E.; Miyaji, H.; Takita, H.; Kanayama, I.; Tsuji, M.; Akasaka, T.; Sugaya, T.; Sakagami, R.; Kawanami, M. Graphene oxide coating facilitates the bioactivity of scaffold material for tissue engineering. Jpn. J. Appl. Phys. 2014, 53, 06JD04. [Google Scholar] [CrossRef] [Green Version]
  213. Arshad, R.; Fatima, I.; Sargazi, S.; Rahdar, A.; Karamzadeh-Jahromi, M.; Pandey, S.; Díez-Pascual, A.M.; Bilal, M. Novel Perspectives towards RNA-Based Nano-Theranostic Approaches for Cancer Management. Nanomaterials 2021, 11, 3330. [Google Scholar] [CrossRef]
  214. Feng, X.; Jiang, D.; Kang, T.; Yao, J.; Jing, Y.; Jiang, T.; Feng, J.; Zhu, Q.; Song, Q.; Dong, N.; et al. Tumor-homing and penetrating peptide-functionalized photosensitizer-conjugated PEG-PLA nanoparticles for chemo-photodynamic combination therapy of drug-resistant cancer. ACS Appl. Mater. Interfaces 2016, 8, 17817–17832. [Google Scholar] [CrossRef] [PubMed]
  215. Revia, R.A.; Zhang, M. Magnetite nanoparticles for cancer diagnosis, treatment, and treatment monitoring: Recent advances. Mater. Today 2016, 19, 157–168. [Google Scholar] [CrossRef] [PubMed]
  216. Naidoo, C.; Kruger, C.A.; Abrahamse, H. Photodynamic therapy for metastatic melanoma treatment: A review. Technol. Cancer Res. Treat. 2018, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Heo, D.N.; Yang, D.H.; Moon, H.-J.; Lee, J.B.; Bae, M.S.; Lee, S.C.; Lee, W.J.; Sun, I.-C.; Kwon, I.K. Gold nanoparticles surface-functionalized with paclitaxel drug and biotin receptor as theranostic agents for cancer therapy. Biomaterials 2012, 33, 856–866. [Google Scholar] [CrossRef]
  218. Dilnawaz, F.; Singh, A.; Mohanty, C.; Sahoo, S.K. Dual drug loaded superparamagnetic iron oxide nanoparticles for targeted cancer therapy. Biomaterials 2010, 31, 3694–3706. [Google Scholar] [CrossRef]
  219. Dash, B.; Jose, G.; Lu, Y.-J.; Chen, J.-P. Functionalized reduced graphene oxide as a versatile tool for cancer therapy. Int. J. Mol. Sci. 2021, 22, 2989. [Google Scholar] [CrossRef]
  220. Shen, J.M.; Gao, F.Y.; Yin, T.; Zhang, H.X.; Ma, M.; Yang, Y.J.; Yue, F. cRGD-functionalized polymeric magnetic nanoparticles as a dual-drug delivery system for safe targeted cancer therapy. Pharmacol. Res. 2013, 70, 102–115. [Google Scholar] [CrossRef]
  221. Zhang, Q.; Neoh, K.G.; Xu, L.; Lu, S.; Kang, E.T.; Mahendran, R.; Chiong, E. Functionalized mesoporous silica nanoparticles with mucoadhesive and sustained drug release properties for potential bladder cancer therapy. Langmuir 2014, 30, 6151–6161. [Google Scholar] [CrossRef]
  222. Kim, S.H.; Lee, J.E.; Sharker, S.M.; Jeong, J.H.; In, I.; Park, S.Y. In vitro and in vivo tumor targeted photothermal cancer therapy using functionalized graphene nanoparticles. Biomacromolecules 2015, 16, 3519–3529. [Google Scholar] [CrossRef]
  223. Wu, Y.-F.; Wu, H.-C.; Kuan, C.-H.; Lin, C.-J.; Wang, L.-W.; Chang, C.-W.; Wang, T.-W. Multi-functionalized carbon dots as theranostic nanoagent for gene delivery in lung cancer therapy. Sci. Rep. 2016, 6, 1–12. [Google Scholar] [CrossRef]
  224. Xia, Y.; Chen, Y.; Hua, L.; Zhao, M.; Xu, T.; Wang, C.; Li, Y.; Zhu, B. Functionalized selenium nanoparticles for targeted delivery of doxorubicin to improve non-small-cell lung cancer therapy. Int. J. Nanomed. 2018, 13, 6929–6939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Zhang, X.; Wu, J.; Williams, G.R.; Niu, S.; Qian, Q.; Zhu, L.-M. Functionalized MoS2-nanosheets for targeted drug delivery and chemo-photothermal therapy. Colloids Surf. B Biointerfaces 2019, 173, 101–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Ammarullah, M.I.; Afif, I.Y.; Maula, M.I.; Winarni, T.I.; Tauviqirrahman, M.; Akbar, I.; Basri, H.; Van der Heide, E.; Jamari, J. Tresca Stress Simulation of Metal-on-Metal Total Hip Arthroplasty during Normal Walking Activity. Materials 2021, 14, 7554. [Google Scholar] [CrossRef] [PubMed]
  227. Basri, H.; Syahrom, A.; Prakoso, A.T.; Wicaksono, D.; Amarullah, M.I.; Ramadhoni, T.S.; Nugraha, R.D. The Analysis of Dimple Geometry on Artificial Hip Joint to the Performance of Lubrication. J. Phys. Conf. Ser. 2019, 1198, 042012. [Google Scholar] [CrossRef]
  228. Webster, T.J.; Siegel, R.W.; Bizios, R. Osteoblast adhesion on nanophase ceramics. Biomaterials 1999, 20, 1221–1227. [Google Scholar] [CrossRef]
  229. Tran, P.A.; Sarin, L.; Hurt, R.H.; Webster, T.J. Opportunities for nanotechnologyenabled bioactive bone implants. J. Mater. Chem. 2009, 19, 2653–2659. [Google Scholar] [CrossRef]
  230. Zinger, O.; Anselme, K.; Denzer, A.; Habersetzer, P.; Wieland, M.; Jeanfils, J.; Hardouin, P.; Landolt, D. Time-dependent morphology and adhesion of osteoblastic cells on titanium model surfaces featuring scale-resolved topography. Biomaterials 2004, 25, 2695–2711. [Google Scholar] [CrossRef]
  231. Chu, P.K.; Chen, J.Y.; Wang, L.P.; Huang, N. Plasma-surface modification of biomaterials. Mater. Sci. Eng. R Rep. 2002, 36, 143–206. [Google Scholar] [CrossRef] [Green Version]
  232. Yu, G.; Hu, L.; Vosgueritchian, M.; Wang, H.; Xie, X.; McDonough, J.R.; Cui, X.; Cui, Y.; Bao, Z. Solution-processed graphene/MnO2 nanostructured textiles for high-performance electrochemical capacitors. Nano Lett. 2011, 11, 2905–2911. [Google Scholar] [CrossRef]
  233. Rojaee, R.; Fathi, M.; Raeissi, K. Electrophoretic deposition of nanostructured hydroxyapatite coating on AZ91 magnesium alloy implants with different surface treatments. Appl. Surf. Sci. 2013, 285, 664–673. [Google Scholar] [CrossRef]
  234. Vorobyev, A.Y.; Guo, C. Direct femtosecond laser surface nano/microstructuring and its applications. Laser Photonics Rev. 2013, 7, 385–407. [Google Scholar] [CrossRef]
  235. Bolelli, G.; Bellucci, D.; Cannillo, V.; Lusvarghi, L.; Sola, A.; Stiegler, N.; Müller, P.; Killinger, A.; Gadow, R.; Altomare, L.; et al. Suspension thermal spraying of hydroxyapatite: Microstructure and in vitro behaviour. Mater. Sci. Eng. C 2014, 34, 287–303. [Google Scholar] [CrossRef] [PubMed]
  236. Sima, F.; Davidson, P.M.; Dentzer, J.; Gadiou, R.; Pauthe, E.; Gallet, O.; Mihailescu, I.N.; Anselme, K. Inorganic-organic thin implant coatings deposited by lasers. ACS Appl. Mater. Interfaces 2014, 7, 911–920. [Google Scholar] [CrossRef] [PubMed]
  237. McEntire, B.; Bal, B.S.; Rahaman, M.; Chevalier, J.; Pezzotti, G. Ceramics and ceramic coatings in orthopaedics. J. Eur. Ceram. Soc. 2015, 35, 4327–4369. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the functionalization of different types of nanomaterials by polymers, natural biomolecules, and synthetic ligands, and their applications in nanomedicine.
Figure 1. Schematic representation of the functionalization of different types of nanomaterials by polymers, natural biomolecules, and synthetic ligands, and their applications in nanomedicine.
Materials 15 03251 g001
Figure 2. Representation of different types of organic and inorganic nanomaterials used in nanomedicine.
Figure 2. Representation of different types of organic and inorganic nanomaterials used in nanomedicine.
Materials 15 03251 g002
Figure 3. Representation of a nanoparticle functionalized with different types of ligands, polymers, therapeutic compounds, and biomolecules. Adapted from Ref. [40], copyright 2020, with permission from Impact Journals LLC.
Figure 3. Representation of a nanoparticle functionalized with different types of ligands, polymers, therapeutic compounds, and biomolecules. Adapted from Ref. [40], copyright 2020, with permission from Impact Journals LLC.
Materials 15 03251 g003
Figure 4. Representation of the covalent functionalization of graphene oxide (GO) with a biological macromolecule, tannic acid (TA) via formation of ether and ester linkages. Reproduced from Ref. [45], copyright 2022, with permission from Elsevier.
Figure 4. Representation of the covalent functionalization of graphene oxide (GO) with a biological macromolecule, tannic acid (TA) via formation of ether and ester linkages. Reproduced from Ref. [45], copyright 2022, with permission from Elsevier.
Materials 15 03251 g004
Figure 5. Representation of the surface modification of different types of NPs with albumin. Reproduced from Ref. [47], copyright 2016, with permission from John Wiley and Sons.
Figure 5. Representation of the surface modification of different types of NPs with albumin. Reproduced from Ref. [47], copyright 2016, with permission from John Wiley and Sons.
Materials 15 03251 g005
Figure 6. Schematic illustration of polymer grafting approaches: “grafting from”, “grafting to”, “grafting through” and “in situ” preparation in the presence of an inorganic precursor.
Figure 6. Schematic illustration of polymer grafting approaches: “grafting from”, “grafting to”, “grafting through” and “in situ” preparation in the presence of an inorganic precursor.
Materials 15 03251 g006
Figure 7. Schematic representation of the functionalization process of AuNPs with papain. The surface modification was achieved by a layer-by-layer (LbL) approach via activation of COOH groups of the modified AuNPs with EDC and sulfonated NHS as coupling agents, followed by amide bonding with the NH2 groups of papain. (PAH+, polyallylamine hydrochloride; PAA−, polyacrylic acid sodium). Adapted from Ref. [68], copyright 2017, with permission from Elsevier.
Figure 7. Schematic representation of the functionalization process of AuNPs with papain. The surface modification was achieved by a layer-by-layer (LbL) approach via activation of COOH groups of the modified AuNPs with EDC and sulfonated NHS as coupling agents, followed by amide bonding with the NH2 groups of papain. (PAH+, polyallylamine hydrochloride; PAA−, polyacrylic acid sodium). Adapted from Ref. [68], copyright 2017, with permission from Elsevier.
Materials 15 03251 g007
Figure 8. (a) TEM image of polymeric micelles, (b) SEM image of AuNRs, and (c) TEM image of AuNRs coated by polymeric micelles. Adapted from Ref. [69], copyright 2020, with permission from Elsevier.
Figure 8. (a) TEM image of polymeric micelles, (b) SEM image of AuNRs, and (c) TEM image of AuNRs coated by polymeric micelles. Adapted from Ref. [69], copyright 2020, with permission from Elsevier.
Materials 15 03251 g008
Figure 9. (a) PEGylated gold nanoparticles for gene delivery. (b) Functionalized gold nanoparticles for drug delivery. Adapted from Ref. [70], copyright 2011, with permission from MDPI.
Figure 9. (a) PEGylated gold nanoparticles for gene delivery. (b) Functionalized gold nanoparticles for drug delivery. Adapted from Ref. [70], copyright 2011, with permission from MDPI.
Materials 15 03251 g009
Figure 10. (Top) Synthesis of Fe3O4 NPs in diethylene glycol (DEG) by thermal decomposition of acetylacetonate (Fe(acac)3), and surface modification by adding surface ligands. (Bottom) TEM images of Fe3O4 NPs (a), Fe3O4 NPs functionalized with dopamine (b), Fe3O4 NPs surface modified with polyethylene glycol with thiol end group (thiol-PEG) (c), and Fe3O4 NPs modified with poly(acrylic acid) (PAA) (d). Adapted from Ref. [84], copyright 2021, with permission from the American Chemical Society.
Figure 10. (Top) Synthesis of Fe3O4 NPs in diethylene glycol (DEG) by thermal decomposition of acetylacetonate (Fe(acac)3), and surface modification by adding surface ligands. (Bottom) TEM images of Fe3O4 NPs (a), Fe3O4 NPs functionalized with dopamine (b), Fe3O4 NPs surface modified with polyethylene glycol with thiol end group (thiol-PEG) (c), and Fe3O4 NPs modified with poly(acrylic acid) (PAA) (d). Adapted from Ref. [84], copyright 2021, with permission from the American Chemical Society.
Materials 15 03251 g010
Figure 11. Functionalization of carbon nanotubes with chitosan based on MALI reaction. Adapted from Ref. [100], copyright 2019, with permission from Elsevier.
Figure 11. Functionalization of carbon nanotubes with chitosan based on MALI reaction. Adapted from Ref. [100], copyright 2019, with permission from Elsevier.
Materials 15 03251 g011
Figure 12. Scheme of the antiviral action of functionalized graphene quantum dots (GQDs). (a) Viral illnesses are caused by binding between the coronavirus (HCoV-229E) S-protein and the host cell receptor. (b) The presence of GQDs can prevent such binding. (c) This mechanism can inhibit the viral genome replication. Adapted from Ref. [107], copyright 2021, with permission from Elsevier.
Figure 12. Scheme of the antiviral action of functionalized graphene quantum dots (GQDs). (a) Viral illnesses are caused by binding between the coronavirus (HCoV-229E) S-protein and the host cell receptor. (b) The presence of GQDs can prevent such binding. (c) This mechanism can inhibit the viral genome replication. Adapted from Ref. [107], copyright 2021, with permission from Elsevier.
Materials 15 03251 g012
Figure 13. Schematic illustration for the preparation process and possible detection principle of the POSS-MIP/QDs. Adapted from Ref. [108], copyright 2021, with permission from Elsevier.
Figure 13. Schematic illustration for the preparation process and possible detection principle of the POSS-MIP/QDs. Adapted from Ref. [108], copyright 2021, with permission from Elsevier.
Materials 15 03251 g013
Figure 14. DNA damage determined by comet assay in PMEF cells exposed to NPs. Cells were respectively treated with 5 μg mL−1 of CB, CNT, SiO2, and ZnO for one day. Damage was evaluated by (A) tail length, (B) tail DNA, (C) tail moment, (D) Olive tail moment. Values shown are the mean from 50 images. * p < 0.05; ** p < 0.01 in comparison to blank. Adapted from Ref. [121], copyright 2009, with permission from Wiley & Sons, Inc.
Figure 14. DNA damage determined by comet assay in PMEF cells exposed to NPs. Cells were respectively treated with 5 μg mL−1 of CB, CNT, SiO2, and ZnO for one day. Damage was evaluated by (A) tail length, (B) tail DNA, (C) tail moment, (D) Olive tail moment. Values shown are the mean from 50 images. * p < 0.05; ** p < 0.01 in comparison to blank. Adapted from Ref. [121], copyright 2009, with permission from Wiley & Sons, Inc.
Materials 15 03251 g014
Figure 15. Mechanisms of cell damage by nanoparticles. (1) Physical damage of membranes. (2) Structural changes in cytoskeleton components. (3) Disturbance of transcription and oxidative damage of DNA. (4) Damage of mitochondria. (5) Disturbance of lysosome functioning. (6) Generation of reactive oxygen species. (7) Disturbance of membrane protein functions. (8) Synthesis of inflammatory factors and mediators. Adapted from Ref. [124], copyright 2018, with permission from Springer Nature.
Figure 15. Mechanisms of cell damage by nanoparticles. (1) Physical damage of membranes. (2) Structural changes in cytoskeleton components. (3) Disturbance of transcription and oxidative damage of DNA. (4) Damage of mitochondria. (5) Disturbance of lysosome functioning. (6) Generation of reactive oxygen species. (7) Disturbance of membrane protein functions. (8) Synthesis of inflammatory factors and mediators. Adapted from Ref. [124], copyright 2018, with permission from Springer Nature.
Materials 15 03251 g015
Figure 16. AuNPs are functionalized in two stages: first with DNA/PEG polymers comprising variable amounts of DNA, and then with rectangular DNA origami. Adapted from Ref. [130], copyright 2017, with permission from The Royal Society of Chemistry.
Figure 16. AuNPs are functionalized in two stages: first with DNA/PEG polymers comprising variable amounts of DNA, and then with rectangular DNA origami. Adapted from Ref. [130], copyright 2017, with permission from The Royal Society of Chemistry.
Materials 15 03251 g016
Figure 17. Summary of the applications of nanomaterials in biomedicine. Adapted from Ref. [135], copyright 2019, with permission from Elsevier.
Figure 17. Summary of the applications of nanomaterials in biomedicine. Adapted from Ref. [135], copyright 2019, with permission from Elsevier.
Materials 15 03251 g017
Figure 18. (a) Setup for NIR fluorescence imaging of cerebral arteries. (b) Imaging of a mouse head. Bright field micrograph (left), NIR fluorescence image without (middle), and with QDs (right). (c) NIR fluorescence pictures of cerebral blood vessels. The upper image shows the fluorescence after the scalp has been removed, whereas the lower micrograph shows the fluorescence after separation—with one-millimeter scale bars. Taken from Ref. [152], copyright 2016, with permission from MDPI.
Figure 18. (a) Setup for NIR fluorescence imaging of cerebral arteries. (b) Imaging of a mouse head. Bright field micrograph (left), NIR fluorescence image without (middle), and with QDs (right). (c) NIR fluorescence pictures of cerebral blood vessels. The upper image shows the fluorescence after the scalp has been removed, whereas the lower micrograph shows the fluorescence after separation—with one-millimeter scale bars. Taken from Ref. [152], copyright 2016, with permission from MDPI.
Materials 15 03251 g018
Figure 19. (a) Transmission Electron Microscopy of liposomes: (A) empty liposomes (LIP); (B) docetaxel-loaded liposomes (LIP-DTX); (C) empty transferrin functionalized liposomes (LIP-TF); and (D) docetaxel-loaded liposomes functionalized with transferrin (LIP-DTX-TF). (b) In vitro release profile of free DTX and encapsulated in liposomes in PBS buffer pH 7.4. Adapted from Ref. [30], copyright 2021, with permission from Elsevier.
Figure 19. (a) Transmission Electron Microscopy of liposomes: (A) empty liposomes (LIP); (B) docetaxel-loaded liposomes (LIP-DTX); (C) empty transferrin functionalized liposomes (LIP-TF); and (D) docetaxel-loaded liposomes functionalized with transferrin (LIP-DTX-TF). (b) In vitro release profile of free DTX and encapsulated in liposomes in PBS buffer pH 7.4. Adapted from Ref. [30], copyright 2021, with permission from Elsevier.
Materials 15 03251 g019
Figure 20. Illustration of biological ligands for active targeting of nanoparticle drug carriers. Taken from Ref. [161], copyright 2019, with permission from MDPI.
Figure 20. Illustration of biological ligands for active targeting of nanoparticle drug carriers. Taken from Ref. [161], copyright 2019, with permission from MDPI.
Materials 15 03251 g020
Figure 21. Illustration of hydroxyapatite-based scaffold-induced regeneration of bone. Reprinted from Ref. [189], copyright 2018, with permission from the American Chemical Society.
Figure 21. Illustration of hydroxyapatite-based scaffold-induced regeneration of bone. Reprinted from Ref. [189], copyright 2018, with permission from the American Chemical Society.
Materials 15 03251 g021
Figure 22. Schematic representation of tumor ablation therapies with iron oxide nanoparticles (NPs). Reproduced from Ref. [215], copyright 2016, with permission from Elsevier.
Figure 22. Schematic representation of tumor ablation therapies with iron oxide nanoparticles (NPs). Reproduced from Ref. [215], copyright 2016, with permission from Elsevier.
Materials 15 03251 g022
Figure 23. Comparison of bone regeneration using nanomaterials and traditional materials. Nanomaterials show improved protein adsorption, osteoblast anchoring, and differentiation compared to traditional materials. Reproduced from Ref. [142], copyright 2020, with permission from Elsevier.
Figure 23. Comparison of bone regeneration using nanomaterials and traditional materials. Nanomaterials show improved protein adsorption, osteoblast anchoring, and differentiation compared to traditional materials. Reproduced from Ref. [142], copyright 2020, with permission from Elsevier.
Materials 15 03251 g023
Table 1. Applications of functionalized nanomaterials in nanomedicine.
Table 1. Applications of functionalized nanomaterials in nanomedicine.
ApplicationExampleRef.
Diagnostic ImagingX Ray
Tomography
Magnetic resonance imaging
Photothermal imaging
[136]
[137]
[138]
[112]
TherapyDrug delivery
Gene and stem cell therapy
Hair growth
[139]
[140]
[141]
Medical implantsOrthopaedic
Cardiovascular
Neurological
Dental
[142]
[143]
[144]
[145]
Tissue EngineeringBone
Cartilage
[54]
[146]
AnticancerPaclitaxel
DOX
Docetaxel
Gambogic acid
[108]
[60]
[30]
[147]
SensingGlucose
Insulin
Metabolic biomarkers
[106]
[148]
[38]
Antimicrobial and AntiviralStreptomycin, penicillin
Coronavirus
E. coli
Airborne viruses
[149]
[107]
[104]
[150]
Table 2. Functionalized nanomaterials used for drug/gene delivery.
Table 2. Functionalized nanomaterials used for drug/gene delivery.
NanomaterialFunctionSize (nm)Drug/GeneTarget Organ & IndicationRef.
Porous CaCO3Intranasal drug
carrier
2000–3200InsulinPostprandial hyperglycemia in diabetes[162]
CaCO3 NPsDrug/gene delivery116Ciprofloxacin HClS. Aureus[163]
CaCO3Drug delivery40–200Hydrophilic drugs and
bioactive proteins
(validamycin)
Inflamed region[164]
Cationic NPsGene delivery50–100Raf gene, ATPμ-RafAngiogenic blood vessels (tumor-bearing mice)[165]
Fe3O4@GODrug release and antitumor therapy200–1000Hybrid microcapsuleTumor cells targeting[166]
GO flakesDrug release1000–2000DOX
microcapsules
-[167]
AuNPsDrug delivery100--Nasopharyngeal carcinoma
cells
[168]
FA-Au-FITC 1Drug delivery for cancer therapy4–7DOXCytoplasm[169]
HLA 2-Si/Fe3O4 NPsDrug delivery for cancer therapy40–110DOXTumor tissues[170]
Fe3O4-SA-PVA-BSA 3Drug delivery240–460DOXCancer cells[171]
CS-HYL-5-FU-PEG-G 4Drug delivery300–580COLO-205 and HT-29 colonCancer cells[172]
SA/PVA/Ca 5Drug delivery
system
500–1000Diclofenac sodium-[173]
PLGA 6-Fe3O4Drug delivery
system
675-FluorouracilProstate carcinoma cell[113]
HLA-NanoemulsionDrug delivery
system
Ciprofloxacin-[174]
Fe3O4Drug delivery
system
20Gambogic acidCapan-1 pancreatic cancer cells[147]
PLGA-Fe3O4 NPsIntratumoral drug delivery200–300DOXMurine Lewis lung carcinoma cells[175]
Fe3O4 conjugate oleate/oleylamineDrug release12ChromoneHeLa cells[176]
Fe3O4/DPA-PEG-COOH 7Drug delivery9Dextran, PEGMacrophage Cells[177]
Thiolated starch-coated Fe3O4Drug delivery40–50IsoniazidHuman body cells[178]
Zn-doped Fe3O4
nano-octahedral core
Drug delivery10–20DOX and HSP70/HSP90 siRNAsTumor cells[138]
Arginine-NCQDs 8Gene delivery6–11EGFP geneMammalian cells[179]
1 Folic acid-coated gold nanoparticles conjugated with a fluorophore; 2 Hyaluronic acid-modified mesoporous silica-coated Fe3O4 NPs; 3 Fe3O4 nanoparticles coated with a mixture of sodium alginate (SA), polyvinyl alcohol (PVA), and bovine serum albumin (BSA); 4 Polyethylene glycol-gelatin-chitosan-hyaluronidase-5-fluorouracil; 5 Sodium alginate/polyethylene glycol (vinyl alcohol); 6 Poly(lactic-co-glycolic acid); 7 Dopamine-polyethylene glycol-carboxylic acid; 8 Nitrogen-doped carbon quantum dots.
Table 3. Functionalized nanomaterials utilized in tissue engineering.
Table 3. Functionalized nanomaterials utilized in tissue engineering.
NanomaterialFunctionSize (nm)TissuePurpose & OutcomesRef.
PEG-GOTissue engineering50BoneImproved thermal stability, hydrophilicity, water absorption, biodegradation, mechanical, viscoelastic, and
antibacterial properties
[54]
Oxidized alginate/
gelatin hydrogel
Tissue regeneration100–200Cartilage regeneration for the treatment of osteoarthritisUsefulness of the hydrogel in encouraging cellular migration and proliferation[146]
OCMC 1Tissue engineering2000–4000BALB/c3T3 cells
in rates
Biocompatibility, spinnability of hydrogel through electrospinning[199]
Pd/PPy/rGO NC 2Tissue engineering2–4BoneBiocompatibility, osteoproliferation, and bacterial infection prevention[200]
3D macro-rGO/PPYBone tissue
engineering
100–400BackboneCasein phosphopeptide as bioactive for bone engineering, osteoblastic performance,
biological properties
[201]
Chitosan-ZnOSoft tissue
engineering
180 Improved hydrophilicity,
porosity, water absorption, oxygen permeability,
biodegradability, antibacterial and wound healing
[202]
Biphasic Calcium
Phosphate
Bone tissue
engineering
1–2MG63 cellsMicropores and collagen coating influence cellular function, in vitro cellular
behavior, scaffold–osteoblast interactions
[203]
AuNPs/glass-ceramic
matrix
Bone tissue
engineering
5–10BoneIn vitro hydroxyapatite synthesis, controlled release of gold species, biocompatibility, and antibacterial activity of AuNPs[91]
AuNPsTissue
engineering
20Rat brainAuNP biochemical effects on the rat brain, biomarkers of AuNP toxicity[204]
AuNPsTissue Engineering10–50Cardiac tissueEffects of AuNPs on the histological deformities of rat heart tissue, toxicity, therapeutic and diagnostic potential of NPs, and their interaction with proteins and other cells[205]
AuNPsTissue engineering30 nmSubsets of cells
in human organs
NP toxicity in human blood, hemolysis, development of ROS 3, platelet condensation in cell subsets[206]
AuNPs/polymeric
coatings
Tissue engineering18, 35, 65Endothelial cells from human dermisNP toxicity, uptake behavior, and uptake quantification[207]
Bioactive glass scaffoldsTissue engineering50–100Bone repairOsteoblastic cells for bone reconstruction[208]
Na2Ca2Si3O9Bone tissue
engineering
500BoneBioactive and biodegradable scaffold effects, mechanical support[209]
Bioactive glass-
ceramics/apatite
Bone tissue
engineering
8–20BoneCrystallization rate of bioactive glasses on the kinetics of HAD formation[90]
Ca10(PO4)6(OH)2Bone tissue
engineering
1000–2000Trabecular boneExtent and nature of carbonate substitution on HDA[93]
GO/Chitosan
Scaffold
Cardiac tissue––Cardiac tissueInvestigate cell survival, cell adhesion, development of intercellular networks, genes, and proteins expression[210]
GO/Chitosan
Scaffold
Cartilage repair35–60Cartilage tissueNanocomposite effect on human tissue, effects of GO[211]
GO-coated collagen
scaffolds
Tissue engineering––Mouse osteoblastic MC3T3-E1 cellsInfluence of the GO coating on cell growth and differentiation, biocompatibility and biodegradability of collagen scaffolds, bioactivity studies[212]
Nanocrystalline apatite/AuNPsTissue engineering2–25Bone tissue reconstructionToxicity of NPS in simulated physiological fluid[66]
1 Gelatin − oxidized carboxymethyl cellulose. 2 Nanocrystalline cellulose. 3 Reactive oxygen species.
Table 4. Functionalized nanomaterials used for cancer therapy.
Table 4. Functionalized nanomaterials used for cancer therapy.
NanomaterialFunctionalization AgentSize (nm)DrugPurpose & OutcomesRef.
ZnO NPsPBA40CurcuminHigh drug-loading and release rates, in vitro and in vivo antitumor efficacy[88]
AuNPsBeta-cyclodextrin with PEG, biotin, PTX,
rhodamine B
30–50PTXCytocompatibility,
stability, and biomolecule binding ease
[217]
SPION5TR1 Aptamer57EpirubicinMagnetic resonance (MR) traceability, nontoxicity, increased permeability,
retention effect
[82]
Fe3O4 NPSGlycerol monooleate144PTX, rapamycin, alone or combinedIntravenous administration of hydrophobic drugs[218]
rGO 1Fe3O4 NPs54.8CamptothecinpH-responsive drug release profile, good biocompatibility, excellent photodynamic[219]
Fe3O4 MNPs+ PLGAcitric acid130–140DOX, verapamilLoading hydrophilic and hydrophobic drugs[220]
MSN 2β-cyclodextrin with hydroxyl, amino, and thiol groups75.5DOXHigher mucoadhesive on the urothelium[221]
rGOHA-PEG-g-poly(dimethylaminoethyl methacrylate)120−190-Biocompatibility, in vitro cellular uptake sensitive to cancer cells[222]
MSNGalactose277CamptothecinMSN targeting to cancer cells[59]
rPEI- Cdots 3FA143-Biocompatible, good siRNA gene delivery carrier[223]
PLGA NPsbis(sulfosuccinimidyl) suberate (BS3)184CurcuminPromote the loading of low-soluble drugs and aid in sustained released[114]
ZnO NPsPBA414CurcuminCurcumin distribution to the sialic acid is much easier by PBA conjugation[88]
Se NPs(Arg–Gly–Asp–d-Phe–Cys [RGDfC]) cyclic peptide18DOXAntitumor efficacy in vivo, effective cellular uptake A549[224]
CuO NPsFA, starch108.83Cytochrome CAntioxidants, anticancer, antimicrobial, drug-carrier[35]
MoS2FA, BSA133DOXExcellent photothermal conversion ability[225]
1 Reduced graphene oxide. 2 Mesoporous silica nanoparticles. 3 Reducible polyethyleneimine passivated carbon dots.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Díez-Pascual, A.M. Surface Engineering of Nanomaterials with Polymers, Biomolecules, and Small Ligands for Nanomedicine. Materials 2022, 15, 3251. https://doi.org/10.3390/ma15093251

AMA Style

Díez-Pascual AM. Surface Engineering of Nanomaterials with Polymers, Biomolecules, and Small Ligands for Nanomedicine. Materials. 2022; 15(9):3251. https://doi.org/10.3390/ma15093251

Chicago/Turabian Style

Díez-Pascual, Ana M. 2022. "Surface Engineering of Nanomaterials with Polymers, Biomolecules, and Small Ligands for Nanomedicine" Materials 15, no. 9: 3251. https://doi.org/10.3390/ma15093251

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop