Next Article in Journal
Electrospun Fibres with Hyaluronic Acid-Chitosan Nanoparticles Produced by a Portable Device
Next Article in Special Issue
Ultrastructural Features of Gold Nanoparticles Interaction with HepG2 and HEK293 Cells in Monolayer and Spheroids
Previous Article in Journal
On the Sensitivity of the Ni-rich Layered Cathode Materials for Li-ion Batteries to the Different Calcination Conditions
Previous Article in Special Issue
Sensitive SQUID Bio-Magnetometry for Determination and Differentiation of Biogenic Iron and Iron Oxide Nanoparticles in the Biological Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Comprehensive Survey on Nanobiomaterials for Bone Tissue Engineering Applications

1
Department of Materials Science and Nanotechnology, Deenbandhu Chhotu Ram University of Science and Technology, Murthal 131039, India
2
Department of Biotechnology, Deenbandhu Chhotu Ram University of Science and Technology, Murthal 131039, India
3
Department of Bio and Nanotechnology, Guru Jambheshwar University of Science and Technology, Hisar 125001, India
4
School of Mechanical Engineering, Lovely Professional University, Phagwara 144411, India
5
Dipartimento di Meccanica, Matematica e Management, Politecnico di Bari, 70125 Bari, Italy
6
Department of Design, Manufacturing & Engineering Management, University of Strathclyde, Glasgow G1 1XJ, UK
7
Department of Mechanical Engineering, Imperial College London, London SW7 2AZ, UK
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(10), 2019; https://doi.org/10.3390/nano10102019
Submission received: 1 September 2020 / Revised: 8 October 2020 / Accepted: 9 October 2020 / Published: 13 October 2020
(This article belongs to the Special Issue Advances in Nanomaterials in Biomedicine)

Abstract

:
One of the most important ideas ever produced by the application of materials science to the medical field is the notion of biomaterials. The nanostructured biomaterials play a crucial role in the development of new treatment strategies including not only the replacement of tissues and organs, but also repair and regeneration. They are designed to interact with damaged or injured tissues to induce regeneration, or as a forest for the production of laboratory tissues, so they must be micro-environmentally sensitive. The existing materials have many limitations, including impaired cell attachment, proliferation, and toxicity. Nanotechnology may open new avenues to bone tissue engineering by forming new assemblies similar in size and shape to the existing hierarchical bone structure. Organic and inorganic nanobiomaterials are increasingly used for bone tissue engineering applications because they may allow to overcome some of the current restrictions entailed by bone regeneration methods. This review covers the applications of different organic and inorganic nanobiomaterials in the field of hard tissue engineering.

1. Introduction

Nanobiomaterials denote nanometer-sized materials whose structures and constituents have significant and novel characteristics with a strong impact on healing and medicine [1,2]. They include metals, ceramics, polymers, hydrogels, and novel self-assembled materials [3]. Rapid developments in nanotechnology not only led to create new materials and tools for biomedical applications, but also changed the way of using these materials in science and technology [4,5].
Human bone is a dynamic tissue that can rebuild and remodel in the body throughout life [6]. The human bone is a hierarchical assembly of nano- to macro-scale organic and inorganic components involved in transmitting physio-chemical and mechano-chemical cues [7,8]. The schematic of Figure 1 shows that normal human bone contains 30% organic collagen fibrils and 70% inorganic minerals [9,10,11,12], while 2% of the total volume is occupied by bone cells, osteoblasts, osteoclasts, lining cells, progenitor cells, and adipocytes [13,14]. Crystalline phases form 65% of the dry weight of the mineral matrix and most part of calcined fraction in calcium phosphate [15,16].
The continuously growing population and the higher complexity of human interactions have generated new bone-related diseases (e.g., bone tumors, bone infections, and bone loss). This requires effective handling and treatment for bone regeneration [17,18]. Tissue engineering has revolutionized orthopedic and surgical studies, providing a new direction in the field based on nanoscale surface modification to simulate properties of extracellular matrix (ECM) and new foundations of structural variables of autologous tissue [19,20]. Tissue engineering is used to generate, restore, and/or replace tissues and organs by using biomaterials and helps to produce similar native tissue or organ [21].
Nanotechnology solved many questions in tissue engineering by modifying regenerative strategies [22]. Biomedical applications of nanotechnology became a hot subject because different nanomaterials are utilized for the synthesis of scaffolds or implants [23,24]. These nanomaterials may be metallic [25], ceramic, or polymeric [26] with different structural forms such as tubes, rods, fibers, and spheres [27]. Various properties of materials such as physiochemical, electrical, mechanical, optical, catalytic, and magnetic properties can be improved at the nanoscale [28] and tailored to specific applications. Nanomaterials synthesized through top-down or bottom-up approaches [29] (Figure 2) have outstanding properties, which are used for biomedical applications particularly in tissue engineering [22].
Existing biomaterials often do not integrate with host tissue completely. This may cause infection and foreign body reactions that lead to implant failure [30]. Indeed, nanostructured biomaterials imitate the natural bone’s extracellular matrix (ECM), producing an artificial microenvironment that promotes cell adhesion, proliferation and differentiation [31]. The specific biological, morphological, and biochemical properties of nanobiomaterials attract researchers to use them for the hard tissue engineering [32]. Nanostructured biomaterials can be used to fabricate high-performance scaffolds or implants with tailored physical, chemical, and biological properties. Several natural and synthetic nanostructured biomaterials are now available for the fabrication of scaffolds with decent bioactivity [33].
This survey article presents the most relevant applications of nanobiomaterials to bone tissue engineering, trying to highlight how organic and inorganic nanobiomaterials can deal with the above mentioned requirements on bone regeneration and the multiple challenges entailed by such a complicated subject. A broad overview of the various types of nanobiomaterials and their applications in the field of hard tissue engineering is provided.
Besides the introductory articles mentioned above and other general articles on topics related to nanomaterials and the different contexts where they operate, the present survey covers some 550 technical papers focusing on types, fabrication, and applications of nanobiomaterials. These articles have been selected using three widely used academic search engines: Scopus, Web of Science, and Google Scholar. For that purpose, keywords such as “tissue engineering”, “bone tissue engineering”, “tissue regeneration”, “scaffolds”, “nanomaterials”, and “nanobiomaterials” as well as their combinations have been used as input for the search process. High priority is given to peer-reviewed journal articles with respect to book chapters and conference proceedings, which count for some 10 papers, less than 1.75% of the total number of surveyed articles. More detailed statistics on the articles surveyed for each topic will be reported at the end of the corresponding subsections.
The paper is structured as follows. Section 2 and Section 3 describe the various types of organic and inorganic nanobiomaterials and their applications in bone tissue engineering/regenerative medicines, drug/gene delivery, anti-infection properties, coatings, scaffold fabrication, and cancer therapy; generalizing conclusions are given at the end of each section. The conclusion section summarizes the main findings of this survey.

2. Nanobiomaterials

Nanobiomaterials cover a wide variety of biomaterials including natural and artificial materials, used for various applications in tissue engineering [34,35,36]. These materials can be classified into two categories, i.e., organic nanobiomaterials and inorganic nanobiomaterials, where the former is characterized by the presence of carbon-containing constituents. Organic–inorganic hybrids are much more effective biomaterials than pure polymers, bioglasses, metals, alloys, and ceramics [37] as they try to combine best properties of constituents following the general concept of composite material. Section 2.1 will review the different types of organic nanobiomaterials, while Section 2.2 will review the different types of inorganic nanobiomaterials.

2.1. Organic Nanobiomaterials

Nanostructured materials have characteristics like biocompatibility, nontoxicity, and non-carcinogenicity. When used for replacement or restoration of body tissue, they are regarded as organic nanobiomaterials [38]. Several research groups have shifted their attention from metallic to organic nanomaterials, such as lipids, liposomes, dendrimers, and polymers including chitosan, gelatin, collagen, or other biodegradable polymers [39]. Organic materials are combinations of a few of the lightest elements, particularly hydrogen, nitrogen and oxygen, and carbon-containing chemical compounds located within living organisms [40]. Proteins, nucleic acids, lipids, and carbohydrates (the polysaccharides) are the basic types of organic materials [41].
Table 1 presents a general classification of organic nanobiomaterials and summarizes representative applications of each material in tissue engineering. The following subsections present a general description of each nanomaterial type listed in Table 1 and a detailed literature survey on the corresponding developments for tissue engineering.

2.1.1. Lipids

Lipids are small hydrophobic or amphiphilic molecules [67]. They can be classified as fatty lipids of acylglycerol, phospholipids such as glycerides, seduction lipids, sterols, demonstrations of lipids played, lipids, and polylactide Kane [68]. Lipids are essential agents for the physiological and pathophysiological functioning of cells [69]. Generally, 10–1000 nm sized spherical lipid nanoparticles are synthesized [70]. All organisms consist of lipids as basic components, among other ingredients. The use of these lipids in pharmaceutical and biomedical fields can solve the problem of biocompatibility and biodegradation [71]. Besides liposomes (lipids arranged in the formation), other unique structures (e.g., hexagonal, spongy, solid structure, etc.) resulting from lipid polymorphisms also are available [72,73]. The latter have better stability and production efficiency than liposomes [48]. Lipid nanocarriers are better than polymeric nanoparticles (NPs) in terms of biocompatibility and lower toxicity, production cost and scalability, and encapsulation efficiency of highly lipophilic actives [74,75]. Lipid nanocarriers such as solid lipid nanoparticles (SLN) [76], nanostructured lipid carriers (NLC) [77], lipid nanocapsules (LNC) [78], and drug–lipid conjugates [79] are used for various administration routes (i.e., parenteral, oral, and topical ones) [80]. Lipid polymer hybrid nanoparticles (LPHNs) can also be used in the area of bioimaging agents for medicinal diagnostics as delivery vehicles like iron oxide, quantum dots (QDs) fluorescent dyes, and inorganic nanocrystals [81].

2.1.2. Liposomes

Liposomes were discovered in the mid-1960s by A. D. Bangham [82]. The vesicle of the liposome is easily fabricated in a laboratory and made of one or more phospholipid bilayers [83] (Figure 3). These are self-assembled versatile particles with diameters ranging from nanometer to micrometer scale [84]. Resembling lipid cell membranes, the nature of phospholipid depends on the length of fatty acid chains [48]. They have the ability to encapsulate and carry hydrophobic aqueous agents [82]. They exhibit many advantages over other carrier systems [85,86].
Bone morphogenetic protein-2 (BMP-2) is one of the most potent proteins in bone regeneration [87]. For this reason, encapsulation of BMP-2 in nanomaterials has attracted great interest. BMP-2-loaded liposomal-based scaffolds may possess better osteoinductivity and bone formation ability [88].
Liposomes can carry drugs directly to the site of action and sustain their levels without causing toxicity for long periods [89]. By changing the composition of lipids, liposome properties can change. Some liposome preparations for anticancer drugs have successfully released on the market by acquiring FDA’s approval [83]. Gentamycin- and vancomycin-integrated liposome-loaded particles are employed for manufacturing of scaffolds [90]. The integration of bioactive aspirin into a liposome delivery system would have a beneficial impact on stem cell osteoblast differentiation [91]. The initial drug amount and the chemical and physical drug properties are considerable factors for the encapsulation efficiency [92]. DOXIL®, the first FDA-approved nanodrug, which consists of liposomes encapsulating doxorubicin, was prepared by this remote loading method [93]. This method can also be used for preparing liposomes encapsulating other drugs such as daunorubicin and vincristine [94]. Liposomal systems are highly used despite being the oldest of the non-viralgene-delivery vehicles [95]. Scaffolds used as delivery vehicles for bioactive agents offer many advantages such as enhanced and extended gene expression, and the ability to control a localized delivery of cargo [96] (see Figure 3 and Figure 4).

2.1.3. Dendrimers

Dendrimers are the newest class of highly-defined macromolecules, which differs from simple polymers by branching at each repeating unit [97]. Their step-by-step controlled synthesis is used worldwide for molecular chemistry, while their repeating structure made of monomers relate them to the world of polymers [98,99]. The repetitively branched nanometer-scale dimension of dendrimers is an ideal candidate for a variety of tissue engineering [100], molecular imaging [101], and drug delivery [102] applications. Dendrimers can be a main component of scaffolds mimicking cross-linkers, chemical surface modifiers, and charge modifiers, as well as natural extracellular matrices [103].
The combination of dendrimers with other conventional structural polymers, such as proteins, carbohydrates and linear synthetic polymers, leads to obtain new physical, mechanical and biochemical properties of hybrid structures [100,104]. The center of dendrimer may be composed of polypropylimine (PPI), di-aminobutyl (DAB), polyamidoamine (PAMAM), and ethylenediamine (EDA), along with various surface residues such as amine, carboxyl, and alcoholic groups [105]. A dendrimer can be synthesized for particular use in different parts with controlled properties like solubility and thermal stability [106].
Dendrimer–drug conjugation is a better approach to the encapsulation of cytotoxic pharmaceuticals. In this way, numerous cytotoxic and anticancer drugs, and targeted individuals such as monoclonal antibodies, peptides, and folic acid, can be conjugated to a single dendrimer molecule [107]. The drug is covalently conjugated to the dendrimer rather than complexed (Figure 5) [108] and these conjugates are relatively more stable.
Dendrimers are a good choice for hydrophobic moieties and poorly water-soluble drugs [109]. PAMAM dendrimer/DNA complexes were employed to encapsulate functional fast biodegradable polymer films used for substrate-mediated gene delivery [110].
The physicochemical characteristics, such as solubility and pharmacokinetics, of dendrimers are better than those of linear polymers. Therefore, dendrimers are ideal candidates for incorporation into scaffolds used for tissue engineering applications [111,112]. A few scaffolds were fabricated with dendrimers such as poly(caprolactone) chains conjugated to a poly(L-lysine) dendritic core to fabricate an HA-composite [113], linear PCL/n-HA hybrids [114], N-hydroxy succinimide/1-ethyl-3-(3-dimethyl aminopropyl) carbodiimide (NHS/EDC) cross-linked scaffold [115], and dexamethasone carboxymethyl chitosan/PAMAM [116] for in vitro bone regeneration.

2.1.4. Polymeric Nanomaterials

Polymeric nanoparticles of size range 10 nm to 1 μm are the most advanced noninvasive approaches to tissue engineering and drug delivery applications [117]. They are comprised of repeating units of chain-like macromolecules with multiple structures and compositions [118]. In general, polymeric nanoparticles can be used for different applications by changing the physicochemical properties of nanoparticles. Polymers are differently processed to produce nanofibers [119], spherical nanoparticles [120] and polymeric micelles [121] for specific applications.
There are several techniques to synthesize polymer-based nanoparticles, applied in tissue engineering [122]. Gelation [123], emulsion–solvent evaporation [124], nanoprecipitation [125], salting-out [122], and desolvation process [126] are generally preferred for natural polymers, like proteins and polysaccharides. Similar to other nanoparticle systems, polymer-based nanoparticles or nanocomposites can be functionalized to perform active targeting [127].
Polymeric nanoparticles alter and may enhance the pharmacokinetic and pharmacodynamic properties used for various drug types because they show controlled and sustained release properties [128]. They offer a variety of benefits ranging from the administration of non-soluble drugs to protection of unstable compounds [129]. These nanoparticles can be loaded with therapeutic or bioactive molecules (Figure 6) either by dispersion or adsorption within the polymer matrix, or encapsulation [130,131].
Drug release may occur directly from nanoparticles through diffusion and polymeric nanoparticles may dissociate into monomers [132]. Polymers used for nanoparticle fabrication should be degradable via enzymatic or non-enzymatic routes under common metabolic pathways [133,134]. Drug-containing polymeric nanoparticles must be stable during migration to the plasma, that is, at almost neutral pH [135].
Chitosan, collagen, gelatin, hyaluronic acid, alginate, and albumin are representative examples of natural biopolymers [136,137]. Polymeric nanoparticles are one of the fastest-growing platforms for the applications in tissue engineering because of their biocompatibility, biodegradability, low cytotoxicity, high permeation, ability to deliver poorly soluble drugs, and retaining bioactivity after degradation [117]. Some newly designed polymeric nanoparticles are sensitive to pH, temperature, oxidizing/reducing agents, and magnetic field which support a high efficiency and specificity for tissue engineering applications [138,139]. Due to good biocompatibility and adjustable chemical composition, and their ability to reorganize, polymeric nanoparticles are very promising as nanobiomaterials for the fabrication of scaffolds or bone substitutes [140]. Plasma protein-based nanoparticles have shown high biodegradability, bioavailability, long in vivo half-lives, and long shelf lives without any toxicity. Blood plasma is a complex mixture of 100,000 proteins, but only two of these proteins have been used in drug administration and tissue regeneration [141,142].

Chitosan

Chitosan is a natural and nontoxic linear biopolymer synthesized from alkaline N-deacetylation of chitin [143]. It can be extracted from exoskeleton of crustacean shells (i.e., crabs and shrimps) some microbes, yeast, and fungi [144]. It has different molecular weights and is soluble in various organic solutions at pH 6.5 and below. The shape of the chitosan nanoparticles is affected by the degree of deacetylation [145,146]. The presence of amine and hydroxyl group leads the use of these compounds in many research areas [147,148]. Chitosan has outstanding biochemical properties, making it very attractive for applications in many areas including tissue engineering and/or regenerative medicine (Figure 7) [149]. Chitosan nanoparticles carry well therapeutic agents and biomolecules because of their high biocompatibility and biodegradability. Because of their small size, they can pass through biological barriers in vivo and deliver the drugs at the targeted site [150].
Applications of Chitosan: Scaffolds prepared from chitosan and ceramics, especially hydroxyapatite, may have superior osteoconductive properties [151]. Bone morphogenetic protein-2 (BMP-2)-loaded chitosan nanoparticles used for the coating of Ti implants were selected in order to examine bone regeneration in mice [55]. Chitosan and growth factor (BMP-7) were used to functionalize a thick electrospun poly(ε-caprolactone) nanofibrous implant (from 700 μm to 1 cm thick), which produced a fish scale-like chitosan/BMP-7 nano-reservoir. This nanofibrous implant mimicked the extracellular matrix and enabled in vitro colonization and bone regeneration [152]. There, the polycationic nature of chitosan entails an antimicrobial behavior at nanoscale [153]. Besides the orthodontic field, there are relevant applications of chitosan in skin healing, nerve regeneration, and oral mucosa [39]. Nanobioglass incorporated chitosan-gelatin scaffolds showed excellent cytocompatibility and ability to accelerate the crystallization of bone-like apatite in vitro [154,155]. The nanocomposite of chitosan/hydroxyapatite-zinc oxide (CTS/HAp-ZnO) supporting organically modified montmorillonite clay (OMMT) was synthesized and used for hard tissue engineering applications [156]. BMP-2 and BMP-7 loaded poly(3-hydroxybutyrate-co3-hydroxyvalerate) nanocapsules were used for the fabrication of chitosan-poly(ethylene oxide) scaffolds [157]. Mili et al. [158] used nerve growth factor (NGF) loaded chitosan nanoparticles for neural differentiation of canine mesenchymal stem cells. Freeze-dried nano-TiO2/chitosan scaffolds showed high biocompatibility and antibacterial effects [159]. Chitosan-poly(vinyl alcohol)-gum tragacanth (CS/PVA/GT) hybrid nanofibrous scaffolds showed 20 MPa ultimate tensile strength and supported L929 fibroblast cells growth [160]. Collagen–chitosan–calcium phosphate microsphere scaffolds fused with glycolic acid did not show relevant differences in their degradation, cytocompatibility, porosity, and Young’s modulus [160,161].

Collagen

The main constituents of living human bone are collagen type-1 (protein) and calcium phosphate or hydroxyapatite (mineral) [162]. Collagen is the major structural protein of the soft and hard tissues in living organisms [163]. It can have a significant role in preserving biological and structural integrity of extracellular matrix (ECM) [164]. It is a versatile material that is widely used in the biomedical field (Figure 8) due to advantages including high biocompatibility and biodegradability [165]. Collagen is mainly used as a carrier for drug delivery as well as osteogenic and bone filling material [166]. Collagen matrix was also used to deliver gene promoting bone synthesis [167]. Collagen with recombinant human bone morphogenetic protein-2 was used to monitor bone formation [168].
Bone morphogenic protein (BMP)-loaded collagen activates osteoinduction in the host tissue [169]. Collagen-based nanospheres/nanoparticles can be used as a systematic delivery carrier for various therapeutic agents or biomolecules [166]. As collagen type-I and hydroxyapatite are a basic part of the bone, hydroxyapatite and collagen were used to fabricate scaffolds that enhance osteoblast differentiation and accelerate osteogenesis [170].
Collagen-based biomaterials in various formats such as 3-D scaffolds have been employed for tissue engineering [171]. The combination of collagen with elastin was successfully fabricated and in vitro tests proved the adhesion and proliferation of cells without any cytotoxicity [172]. Collagen-based inks were used for 3D bioprinting employed for tissue repairing and scaffold fabrication. The collagen-based ink was extruded with a temperature stage of −40 °C, followed by freeze-drying and cross-linking by using 1-ethyl-(3-3-dimethylaminopropyl) hydrochloride solution [173].

Gelatin

Gelatin represents a derivative of collagen, extracted by collagen hydrolysis from the skin, bones, and/or connective tissues of animals. It is a cost effective, biocompatible, and biodegradable polymer, which supports cross-linking of functional groups. Gelatin is a versatile polymer that is known for his wealth merits [174]. Pharmaceutical or medical grade gelatin has fragility and transparency for tablet coatings, suspensions, capsule formulations, and nano-formulations (Figure 9).
Because of their great biocompatibility, the injected gelatin-loaded nanoparticles have been reported in the skeletal system [24,175,176]. It is a polyampholyte at pH 9 (gelatin A) and pH 5 (gelatin B). Gelatin nanoparticles are used as a biomaterial for the delivery of biomolecules and therapeutic agents [177]. However, digestive process of gelatin showed low antigenicity, with the formation of harmless metabolic products. In order to prevent infectious disease transmission, genetic engineering approaches were used for the production of human recombinant gelatin [178,179].
At the nanoscale, gelatin shows high biocompatibility, biodegradability, and low immunogenicity [180]. The presence of a higher number of functional groups on polymer backbone helps with crosslinking and chemical modification [181]. The cross-linking is necessary to stabilize the macromolecular structure of gelatin is not stable at normal body temperature due to the low melting temperature [182,183].
Gelatin methacryloyl (GelMA) hybrid hydrogel demonstrated a wide range of tissue engineering applications. When exposed to light irradiation, GelMA scaffolds convert into hydrogels with tunable mechanical properties [184]. Gelatin enables therapeutic cell adhesion without comprising cell phenotypes [185]. Porous HA-gelatin microparticles (1 to 100 μm) support human osteoblast-like Saos-2 cells growth and cell delivery [186]. A mechanically strong gelatin–silk hydrogel composite was prepared by direct blending of gelatin with amorphous Bombyxmori silk fibroin (SF) [187]. Gelatin coated polyamide (PA) scaffold showed good biomechanical, cell attachment, and wound healing characteristics while being transplanted to nude rats [188]. Poly(lactide-co-glycolide) (PLGA)–gelatin fibrous scaffolds possess the highest Young’s modulus (770 ± 131 kPa) and tensile strength (130 ± 7 kPa) [189]. Methacrylamide-modified gelatin (GelMOD) 3D CAD scaffolds showed excellent stability in culture medium and support porcine mesenchymal stem cell adhesion and subsequent proliferation [190].
Gelatin-based microcarriers used embryonic stem cell delivery for the applications in tissue engineering [191]. The magnetic nanoparticles were assembled with magnetic gelatin membranes to produce 3D multilayered scaffolds (Figure 10), which are used for controlled distribution of magnetically labeled stem cells [192].

Poly (Lactic-co-glycolic) acid (PLGA)

PLGA is considered as one of the most efficient tissue engineering materials due to its (i) high biocompatibility, (ii) biodegradability, (iii) potential to interact with biological materials, and (iv) clinical use approved by FDA [193]. Biodegradable biomolecule-loaded PLGA nanoparticles can be used for the preparation of a drug delivery system, which can be further utilized in scaffold fabrications [194,195,196]. These nanoparticles may increase the mechanical properties of the scaffolds but decrease swelling behavior without changing the morphology of the scaffold [197]. Afterward, this system is effective to prepare a controlled release platform for model drugs that favors the bio-distribution and development of clinically relevant therapies [198]. Different methods such as gas foaming [199], porogen leaching [200], solid freedom fabrication [201], and phase separation [202] can be used for PLGA scaffolds fabrication.

2.1.5. Carbon Nanostructures

Carbon nanomaterials are great candidate materials for bone tissue engineering due to their conductivity, lightweight, stability and strength [203]. Nanostructures such as fullerenes, carbon nanotubes, carbon nanofibers, and graphene are the most common structures (Figure 11) [204,205].
Han et al. [205] pointed out that carbon is biocompatible and can be used in many clinical applications, such as prosthetic heart valves. However, a pure form of carbon nanomaterials cannot be used as a substrate for bone tissue [206]. Therefore, carbon-based materials are used in combined form to fabricate scaffolds [207]. Carbon nanostructures doped or reinforced compositions became more popular due to their high performance and compatibility with bone tissues [203].
Carbon nanofibers (CNF) are cylindrical or conical structures of various diameters and lengths. The interior structure of the CNF contains an improved layout of graphene sheets. Graphene is a single-layer two-dimensional material composed of long-edged reactive carbon atoms. Graphene leaves are characterized by stable dispersion and orientation of nanofillers [208,209,210].
Carbon nanotubes (CNT) enhance mechanical and electrical properties, which helps to generate innovative products. CNTs are one of the ideal and favorable materials used for designing novel polymer composites [211]. Many authors focused on the progress of composite materials fabrication-integrating CNTs to enhance its applications in biomedical field [212,213,214,215,216].
Nanodiamonds (4–10 nm) are typically different from other nanostructures as they are sp3 hybridized [203]. They show admirable protein binding ability and can be used as a carrier for some biomolecules such as BMP-2 [217]. The carbon nanotube/gold hybrids are employed commonly for the delivery of the anticancer drug doxorubicin hydrochloride into A549 lung cancer cell line [218].
Nanoscaffolds can be produced by electrospinning poly(ε-caprolactone) (PCL) and different types of carbon nanomaterials such as carbon nanotubes, graphene, and fullerene [219]. Mesoporous silica (mSiO2) decorated carbon nanotubes (CNTs) hybrid composite were used for the simultaneous applications of gentamicin and protein cytochrome C delivery and imaging [220]. Single-walled carbon nano-horns encapsulated with positively charged lipids complex were used for targeted drug and protein delivery [221].

2.1.6. Summary and Statistical Analysis of the Survey on Organic Nanobiomaterials

The survey on organic nanobiomaterials presented in Section 2.1 regarded some 200 articles. Polymeric nanomaterials, carbon nanostructures, and nanocomposite materials are the most widely investigated subject (60.8% of the studies), followed by dendrimers (21.1%) and lipids/liposomes (18.1%). While dendrimers and lipid/liposomes are mainly utilized as nanocarriers, the other nanomaterials cover a much broader spectrum of applications. The development of new nanomaterials (especially carbon nanomaterials or materials including natural bone constituents such as, for example, collagen) that can improve tissue regeneration, cell growth, and drug/protein delivery currently represents the main research area in the field of organic nanobiomaterials with a strong tendency to design hybrid materials and improve fabrication techniques of the resulting nanocomposite materials/scaffolds/structures. Such a trend has become very clear in the last 5–6 years. However, much work remains to be done in order to fully understand interactions between different phases of nanocomposite materials and cell/tissues to be repaired/treated. Another important issue strictly related to the above mentioned one is how to “optimize” the composition of the nanocomposite for the specific purposes on which the material itself is designed.

2.2. Inorganic Nanobiomaterials

Inorganic biomaterials are those lacking carbon element and they are widely employed for in vivo and in vitro biomedical research [222]. These crystalline or glass structured nanomaterials are used to replace or restore a body tissue [36]. The main applications of inorganic biomaterials, including bioceramics and bioglasses, are for orthopedics and dentistry. Modifications in composition and fabrication techniques may produce a range of biocompatible materials such as bioceramics [223]. Natural bone also includes inorganic materials like calcium (Ca) and phosphorus (P) in the form of hydroxyapatite (HA) crystals, as well as carbonate (CO32−), potassium (K), fluoride (F), chlorine (Cl), sodium (Na), magnesium (Mg), and some trace elements including copper (Cu), zinc (Zn), strontium (Sr), iron (Fe), and silicon (Si) [224]. Therefore, it is very logical to investigate on nanomaterials based on these inorganic constituents.
Table 2 presents a general classification of inorganic nanobiomaterials and summarizes representative applications of each material in tissue engineering. The following subsections present a general description of each nanomaterial type listed in Table 2 and a detailed literature survey on the corresponding developments for tissue engineering.

2.2.1. Nano Silica

A huge amount of investigations on biomedical applications of silica nanostructures have been carried out in the past decade [249]. The ability to synthesize uniform, porous and dispersible nanoparticles, together with the fact that particles’ size and shape can be easily controlled [226], certainly favored the variety of applications of silica in tissue engineering [250]. Furthermore, as silica is biocompatible and chemically stable [251,252], it has been used also for biomedical imaging and medication administration [225], either itself or as a coating of other compounds [251].
Mesoporous silica nanoparticles (MSNPs) have been used as a drug delivery vehicle [253] and to improve mechanical properties of biological materials. It was noted their use as well as for sustained and prolonged release or administration of intracellular genes in bone tissue engineering [226]. MSNPs work as efficient biocompatible nanocarriers due to (i) high visibility, (ii) dispersibility, (iii) binding capability to a target tissue, (iv) ability to load and deliver large concentrations of cargos, and (v) triggered or controlled release of cargos [250]. The functioning of MSNPs can be tailored by modifying the silanol group present within the pore interiors and on the outer surface. These positive chemical moieties are adsorbed by negatively charged SiO groups at neutral pH, through electrostatic interactions (Figure 12).
Anitha et al. [254] reported a composite matrix containing crystalline rod-shaped core with uniform amorphous silica sheath (Si–n HA), which showed good biocompatibility, osteogenic differentiation, vascularization, and bone regeneration potential. Silicate containing hydroxyapatite stimulates cell viability of human mesenchymal stem cells for extended proliferation [255]. Zhou et al. [256] synthesized PLGA–SBA15 composite membranes with different silica contents by electrospinning method; these membranes showed better osteogenic initiation then the pure PLGA membranes. Ding et al. [257] successfully fabricated levofloxacin (LFX)-loaded polyhydroxybutyrate/poly(ε-caprolactone) (PHB/PCL) and PHB/PCL/sol–gel-derived silica (SGS) scaffolds, which support the growth of MG-63 osteoblasts. A microfluidic device was used to generate photo-cross-linkable gelatin microgels (GelMA), coupled with providing a protective silica hydrogel layer for applications in injectable tissue constructs [258]. Dexamethasone (DEX)-loaded aminated mesoporous silica nanoparticles (MSNs-NH2) were prepared via electrophoretic deposition (EPD) and successfully incorporated within poly(l-lactic acid)/poly(ε-caprolactone) (PLLA/PCL) matrix to fabricate composite nanofibrous scaffolds for bone tissue engineering applications [259].

2.2.2. Nano Bioglass

Bioglasses (BG) have been intensively investigated as biomaterials since their discovery in 1969 and first developments in the 1970s made by L. Hench [260]. Compared to common glass, bioglass contains less silica and higher amounts of calcium and phosphorous. As a biomaterial for tissue engineering, bioglass is applied independently or in combination with a number of polymers [261] (Figure 13). BG can arouse fibroblasts with higher bioactivity by accelerating bioactive growth factors and proteins as compared to untreated fibroblasts [262].
BG degrade slowly when implanted into the targeted patient’s site and release ions, which favors the biosynthesis of hydroxyapatite [263]. The silica-rich surface of bioglass promotes the exchange of Ca2+ and PO43− with physiological fluid, which leads to the generation of a Ca–P layer [264,265]. This biodegradation may be enhanced by the presence of a SiO2 network, which forms non-bridging silicon-oxygen bonds [266]; the low connectivity of the SiO2 network enhances dissolution of bioglass while the presence of Na and Ca forms Si–O–Si bonds and reduces dissolution rate. Mesoporous BG can be fabricated using the sol-gel method, which can be a good carrier for targeted drug delivery [267]. The sol–gel method was also used by Kumar et al. [268] to develop bioglass nanoparticles with a higher content of silica, which are suited for bone tissue applications.
Bioglass nanoparticles show high biocompatibility and surface area, which can enhance in vitro osteoconductivity as compared to layer and microsized particles of bioglass [269]. The size of the particles can be modified by changing the synthesis parameters and techniques. However, because of its brittleness, the glass alone cannot be used to heal large bone defects [270]. In order to solve this issue, Bioglass 45S5 was used with poly(D,L-lactide) (PDLLA), a biodegradable polymer, to form a composite scaffold with enhanced biomechanical characteristics [271]. The early failure of a bioglass composite at the interface occurs because of nonuniform mechanical strength, phase separation, nonhomogeneous mixture, and different degradation properties of two compounds. A hybrid composite of poly(methyl methacrylate) (PMMA) and bioactive glass was manufactured via the sol-gel method (Figure 14) to enhance physicochemical and mechanical properties [272].
An elastin-like polypeptidic and bioglass (ELP/BG) hydrogel was also fabricated that is mechanically robust, injectable, and self-healable. This ELP/BG biocomposite can be useful for drug delivery and tissue engineering purposes [273]. A 3D construct of type-I collagen and 45S5 Bioglass meets the basic requirements of a scaffold including biocompatibility, osteoconductivity, osteoinductivity, and biodegradability [274]. Bioglass nanoparticles were also used with bacterially derived poly(3-hydroxybutyrate) to fabricate bioactive composite film using a fermentation technique [275]. Different glass modifiers (Mg2+, Ca2+, and Sr2+) were used to prepare borosilicate bioactive glasses through a melt-quenching technique which showed good antibacterial properties [276]. Poly(propylene fumarate) (PPF) was used to functionalize bioglass particles that enhance the bioactivity and cell adhesion, proliferation, and bone regeneration [277].

2.2.3. Nano Hydroxyapatite

Hydroxyapatite (Ca10(PO4)6(OH)2) is a significant natural mineral constituent of bones (70% wt.) and teeth (96% wt.) [278,279]. Synthetic HA is a biocompatible ceramic material, used for biomedical applications (Figure 13) because it may replicate the behavior of mineral part of the bone [280,281]. It shows outstanding biocompatibility with bones, teeth, skin, and muscles, both in vitro and in vivo [282,283]. The stoichiometric molar ratio Ca/P in synthetic HA of 1.67 is not the actual ratio in the hydroxyapatite of normal bones, because of the presence of other elements such as C, N, Fe, Mg, and Na [284]. Hydroxyapatite (HA) can be easily synthesized by using different methods such as hydrothermal, sol–gel, and co-precipitation methods [285]. The comparison of mineral compositions of hydroxyapatite, bone and teeth is shown in Table 3 [286,287].
HA shows such excellent biocompatibility, bio-inertia and bioactivity without toxicity, immunogenicity [288,289]. It has a good ability to make bonds with bone directly and it is primarily used in therapeutic applications such as implants and fillers for bones and teeth in different forms [290]. To overcome the low mechanical strength of hydroxyapatite scaffolds, a large number of natural and synthetic polymers were combined with HA such as collagen, polyethylene, polylactic acid, alginates, poly(methyl methacrylate), and polycaprolactone [136].
Woodard et al. [291] compared the activity of nano- and microsized ceramic materials in the body. Their studies demonstrated a substantial increase in osteoblast adhesion and protein adsorption in nanomaterials. The major components of the inorganic nanostructure can have a higher biological activity than micro-components [245]. Polydopamine (pDA)-templated hydroxyapatite (tHA) was introduced into polycaprolactone (PCL) matrix to make bioactive tHA/PCL composite based fibrous scaffold; in vitro and in vivo investigations (Figure 15) showed a favorable cytocompatibility at a given concentration of tHA (0–10% wt.) [292].
A new type of scaffold with bamboo fiber (5%) incorporated nano-hydroxyapatite/poly(lactic-co-glycolic) (30%) was fabricated via freeze-drying; bamboo fibers improved biomechanical properties of n-HA/PLGA composite scaffolds thus developing a superior potential for bone tissue engineering [293]. Sol–gel synthesized hydroxyapatite–TiO2-based nanocomposites synthesized in supercritical CO2 have better Young’s and flexural moduli than PCL/HAp composites [294]. A set of techniques including molding/particle leaching and plasma-treated surface deposition were used to fabricate bilayered PLGA/PLGA-HAp composite scaffold [295]; the in vivo rat model experiment proved that the new composite is suitable for osteochondral tissue engineering applications. Electrospinning mediated poly(ε-caprolactone)−poly(ethylene glycol)−poly(ε-caprolactone) (PCL–PEG–PCL, PCEC) and nano-hydroxyapatite (n-HA) composite scaffolds showed good biocompatibility and nontoxicity [296]. Hydroxyapatite/Na(Y/Gd)F4:Yb3+, Er3+ composite fibers [297], and gadolinium-doped mesoporous strontium hydroxyapatite nanorods [298] were successfully used in drug storage/release applications.

2.2.4. Silver Nanoparticles

Silver proved its bactericidal activities against many bacteria since 1000 B.C. Now silver is used as an antiseptic, antibacterial, and antitumor agent [299]. Because of their strong antibacterial activity against both Gram-positive and Gram-negative bacterial strains, silver nanoparticles were widely used for fabricating antibacterial nanocomposite-based scaffolds and coated implants [235,237]. Furthermore, silver may be combined with different materials such as CNT [300], chitosan, HA [301], and manganite [302] to get a specific function. Ag-doped or coated implants allow reducing the number of bacterial infections without interfering with bone cell growth in the body (Figure 16 [303]). The antimicrobial activity of Ag has been reported against Escherichia coli [304], Candida albicans [11], Vibrio cholera [305], and Staphylococcus aureus [306].
In addition to antibacterial activity, Ag nanoparticles promote wound healing, reduce scar formation, and reduce inflammation [307]. Silver nanoparticles have many applications as antimicrobial agents when combined with different biological substances [308]. A micrometer-sized surface-enhanced Raman spectroscopy (SERS) substrate, core–shell microparticles composed of solid carbonate core coated with silver nanoparticles, and polyhedral multishell fullerene-like structure were developed for biomedical applications [309]. Soft poly(vinyl alcohol) (PVA) hydrogel films containing silver particles prepared on solid biodegradable poly(l-lactic acid) (PLLA) exhibit both antibacterial and reduced cell adhesion properties [310]. Biocompatible maleimide-coated silver nanoparticles (Ag NPs) can be used as co-cross-linkers for the preparation of a nanocomposite gelatin-based hydrogel. Covalently bound Ag nanoparticles support swelling and drug release properties of composite hydrogel without producing toxicity [311]. In situ fabricated Ag NPs (4-19 nm) and immobilized on titanium by using a plasma immersion ion implantation process motivated osteoblast differentiation in rat bone marrow stem cells (BMSCs) [312]. Patrascu et al. [313] fabricated collagen/hydroxyapatite-silver nanoparticles (COLL/HA-Ag)-based antiseptic composite for biomedical applications.

2.2.5. Gold Nanoparticles

Gold nanoparticles (GNPs) are defined as a colloid of nanometer-sized particles with better properties than bulk gold. They are produced in different shapes such as spheres [314], rods [315], star-like [316], and cages [317]. GNPs possess unique characteristics, such as easy-to-control, nanoscale size, easy preparation, high surface area, easy functionalization, and excellent biocompatibility, that make them highly suited for many tissue engineering and more in general for biotechnology applications [318,319]. GNPs are definitely superior over other types of nanoparticles in terms of low toxicity and colloidal stability. Furthermore, they present an outstanding physicochemical behavior, which is related to local plasmon resonance phenomena.
Gold nanoparticles were utilized for biosensing [320], bioimaging [321,322,323], cancer therapy [324], gene delivery to enhance osteogenic differentiation [325], and photo-thermal therapy [229,314,316,323]. Gold nanoparticles were also combined with other materials such as silica (to produce core and shell nanoparticles) [318,323] as well as natural polymers (to improve mechanical properties) and synthetic polymers (to enhance biocompatibility) [318].
Due to their excellent biocompatibility and chemical inertness, gold nanoparticles became the ideal choice for the preparation of scaffolds in many cases [318]. The mission of GNPs in tissue engineering and regenerative medicine is to act as a multimodal tool in order to improve scaffold properties, cell differentiation and intracellular growth factor delivery (Figure 17), while monitoring cellular events in real-time [326].

2.2.6. Titanium Dioxide

Titanium is widely used in many surgical applications (e.g., prostheses and implants) because of its excellent biocompatibility, good mechanical properties, and lower mass density than steel [327]. The low density and high specific strength of titanium results in lightweight implants with good mechanical properties [238,239]. Furthermore, the smooth surface of Ti mesh prevents bacterial contamination instead of adsorbate materials. Therefore, titanium mesh provides an excellent solution to guide bone regeneration [243].
Nanostructured TiO2 materials of various morphologies such as nanoparticles, nanorods, nanowires, nanotubes, and other hierarchical nanostructures can be produced using different techniques such as, for example, microwaves [328,329], hydrothermal/solvothermal processes [330,331], sol–gel [332,333], anode oxidation [334,335], chemical vapor deposition [336,337], sonochemical processes [338,339], and green synthesis [340,341,342].
As can be seen from Figure 18, nanostructured TiO2 is a multifunctional material for a wide range of applications in engineering and biomedical areas. Interestingly, TiO2 nanoparticles represent a miniature of electrochemical cells capable of light-induced redox chemistry. This quality can be used for manipulating biomolecules and cell metabolic processes. TiO2 nanoparticles prove to have a higher affinity for binding proteins and other cellular components when used within cellular environment [343,344]. TiO2 nanoparticles can also be used to enhance photodynamic therapy (PDT) and sonodynamic therapy (SDT) [345].
Titanium nanotubes (TNTs) possess excellent biocompatibility and drug-releasing performance. Furthermore, they can be generated on the surface of the existing medical implants [346,347]. The physical adsorption of the drugs promotes the anti-inflammatory properties of the TNTs, and with improved osteoblast adhesion, the drug-eluting technique is extended [348].
TiO2 based scaffolds are biocompatible, have good osteoconductive performance and antibacterial properties [349], and show high porosity, excellent interconnectivity, and sufficient mechanical strength [350,351]. Nanostructured TiO2 can be combined with several polymers including polylactic acid (PLA) [352]; poly(ether-ether ketone) (PEEK) [353]; poly(lactic-co-glycolic acid) (PLGA) [354]; and inorganic materials such as SiO2 [355], Al2O3 [356], bioglass [357], hydroxyapatite [358], graphene [359], and calcium phosphate [360].
Nano-TiO2 surface coated implants can limit autoimmune reactions between the underlying bone tissue surfaces and the implant [361]. However, material deterioration or generation of chronic inflammation in the implanted tissues may reduce success rate [361,362]. Various TiO2 nanostructures were used for loading and eluting cefuroxime as an antibiotic on orthopedic implants [363].

2.2.7. Zirconia

Zirconia was first recognized by M.H. Klaproth in 1789 and used as a pigment for ceramics for a long time [364]. Since the 1970s, zirconia received massive consideration as a biomedical material in association to its chemical and biological inertness [365]. Consequently, zirconia was also used to overcome the brittleness of alumina and the consequent failure of implants [366], and as a material for the repair and replacement of bones due to its unique biomechanical properties [367].
Investigations on zirconia biomaterials began in the 1960s. Classical orthopedics studied for many years have used zirconia in the area of hip replacement [368,369]. Zirconium oxide (zirconia) possesses improved mechanical properties and has become one of the most popular ceramic materials in the field of healthcare due to its high biocompatibility and low toxicity [364,370].
Zirconia is one of the most useful structured ceramics because it provides high resistance to bending and fracture. However, zirconium oxide with a low fracture toughness due to the presence of alumina abrasive grains [371] also was introduced as an alternative to having excellent wear resistance due to the unwanted release of orthopedic alumina. Porous zirconia stents can be manufactured by cutting CAD/CAM blocks in the desired shape, and zirconia stents assembled with HA significantly increase the volume of new bone formation in vivo [372].
While it might be concluded that zirconia has one of the best combinations of mechanical strength, fracture resistance, biocompatibility, and biological activity, its performance can be further enhanced via a proper modification of material’s surface or by combining the material with some other bioactive ceramics and glass [367]. In addition, as a result of the introduction of Zr into the Ca-Si system, no toxicity was observed. Previous studies confirmed that the optimum content of zirconium and strontium increases the surface energy of the magnesium alloy and enhances the ability to stimulate bone formation around the implant [373,374]. Hydroxyapatite and fluorapatite slurry coated zirconia scaffolds induce osteoconductivity and enhance bonding strength up to 33 MPa [375]. The dispersion of zirconia with alumina lead to produce ZrO2-toughened alumina (Al2O3), known as zirconia-toughened alumina (ZTA) [376].
Zirconia (ZrO2)/β-tricalcium phosphate (β-TCP) composite has shown excellent mechanical properties and supports osteoblast regeneration [377]. Silk fibroin-chitosan-zirconia (SF/CS/nano ZrO2) and chitin–chitosan/nano ZrO2 composites provide a suitable environment for cell infiltration and colonization [378,379]. Different temperature based hydroxyapatite-zirconium composites such as 873 K (HZ600), 923 K (HZ650), and 973 K (HZ700) demonstrated that osteoblast growth and mineralization were not influenced by any composite [380]. A new biphasic calcium phosphate (BCP) scaffold reinforced with zirconia (ZrO2) was fabricated through the fused deposition modeling (FDM) technique. The 90% BCP and 10% ZrO2 scaffold thus created had significantly better mechanical properties than 100% BCP and 0% ZrO2 scaffold [381].
ZrO2 nanoparticle (NP)-doped CTS–PVA–HAP composites (ZrCPH I–III) showed improvement in the tensile strength of ZrCPH I–III with respect to the CTS–PVA–HAP scaffold [382]. Sol–gel cum solvothermal derived mesoporous titanium zirconium (TiZr) oxide nanospheres were used for ibuprofen, dexamethasone, and erythromycin drugs loading and in vitro release studies [383]. The excellent biocompatibility of Zr makes it a good material for metal–organic frameworks (MOFs). Surface functionalization of Zr-fumarate MOF (Figure 19) was used for dichloroacetate (DCA) drug loading, which is more efficient at transporting the drug mimic calcein into HeLa cells [384].

2.2.8. Alumina

Since 1975, the bio-inertness of alumina has been confirmed. Alumina has very high hardness and resistance to abrasion on the Moh scale next to diamond [385]. In addition, the crystalline nature of alumina makes it insoluble at room temperature in regular chemical reagents [386]. Alumina has been used in many fabrications of artificial implants since it was inserted into an artificial femur head in the 1970s [387]. Pure and densified alumina, α-Al2O3 (corundum), was the first ceramic material used in the biomedical field for dental restorations, cochlear implants, and load-bearing hip prostheses [388]. As porous alumina does not degrade under in vitro and in vivo environments, it may be used for biosensing [389], good electrical insulation [390], and immune isolation [391].
Properties such as abrasion resistance, power and chemical inertness favor the use of alumina in hard tissue engineering [392]. If the alumina is implanted in bone marrow, no toxic effects are generated in the surrounding tissue [393]. However, the high stiffness of alumina may lead to have a high elastic incompatibility between the biological tissue and the implant [394]. The tensile strength of alumina can be increased by reducing grain size and increasing its density [395]. In view of their good mechanical behavior, alumina implants are characterized by long-time survival predictions [396].
A significant feature in applications involving open and aligned porous structures, such as bone tissue scaffolds, catalysts, and membranes, is the anisotropic nature of porous alumina ceramics [397]. The α-alumina is the most stable oxide amongst transient and metastable types [398]. It should be noted that essential physico-chemical properties of alumina surface are significantly affected by the protein adsorption process. For example, the presence of liquid solutions nearby the implanted site can cause accelerated protein adsorption on the alumina’s surface [399]. Piconi et al. [394] reported the in vitro biocompatibility of alumina with various cell lines such as fibroblasts and osteoblasts, and immunological cells with various cell environments.
The particle size of alumina may affect biocompatibility, particularly when using nanoparticles because of their high surface/volume ratio [400]. Alumina suspensions (70% wt.) and wheat flour (20–30% vol.) were used to synthesize different particle sized porous alumina ceramics [401]. Hydroxyapatite/alumina composite based foam was synthesized via a precipitation method under a variety of pH values that showed a good concentration of Ca2+ and PO43− contents [402]. The chemical modification of porous alumina surface with vitronectin and peptide (i.e., arginine-glycine-aspartic acid cysteine (RGDC)) enhanced bone cell adhesion and production of extracellular matrix [403].
Porous anodic alumina (PAA) can be fabricated on the surface of other materials through anodization process [404,405]. It can be considered a good nanocontainer to load active agents such as drugs or biomolecules [406]. Evaporation induced self-assembly derived mesoporous aluminum oxide was used for the delivery of poor-water soluble compound Telmisartan (anti-blood pressure drug) with 45% loading efficiency [407]. The drug is not loaded within the pores of the PAA completely, but the surface itself can hold some of this load, which can be quite high; this promotes another phase release [408,409].
Calcium phosphate with 20% alumina (Ca3(PO4)2–Al2O3) bio-ceramic composite revealed enhanced biocompatibility and mechanical properties [410]. Using alumina nanowires reinforcement in polyhydroxy butyrate-chitosan (PHB-CTS/3% Al2O3) scaffolds enhanced the mechanical properties of the scaffold. The addition of alumina increased by ten times the tensile strength of PHB-CTS/3% Al2O3, which became higher than its counterpart for the original PHB-CTS scaffold [411].
Al2O3 coating was used for improving the performance of stainless steel 316L and Ti-6Al-4V implants [412]. In general, coating materials are used to protect the surface of the implant material and the interface with the biological system at hand [413]. Nanorod-like HA-coated porous Al2O3 was fabricated by anodic oxidation that revealed excellent biological activity in vitro [414].

2.2.9. Copper

Copper ions stimulate the proliferation of human vein endothelial cells and mesenchymal stem cells (MSCs) but not human dermal fibroblasts [415,416]. Copper nanoparticles can also act as antifungal and antibacterial agents [417]. Copper is commonly used in bone implants for its antimicrobial activity against a wide range of pathogens [418]. As copper is an essential component of the body, it may be more suitable for in vivo applications [25].
The importance of copper has been studied extensively because its deficiency can lead to osteoporosis [419]. Cu also stimulates angiogenesis and collagen deposition, which are key elements in wound healing [420]. The use of copper-based biomaterials is cost-effective compared to other vital materials based on gold and silver [421].
Copper ions were incorporated into biologically active scaffolds for controlled release to improve vascular strengthening and antimicrobial action for prolonged periods [422]. Copper-doped wollastonite (Cu–Ws) particles (1184 nm) have shown biocompatibility towards mouse mesenchymal stem cells (mMSC) up to 0.05 mg/ml concentration [423]. A freeze-dried chitosan/hydroxyapatite/copper-zinc alloy (CS/nHAp/nCu–Zn) composite-based scaffold showed lower degradation and higher protein adsorption without producing toxicity towards rat osteoprogenitor cells [422]. Collagen-copper-doped bioactive glass (CuBG-CS) scaffolds exhibited enhanced mechanical properties (up to 1.9-fold) and osteogenesis (up to 3.6-fold) than chitosan [424].
Copper nanoparticles were investigated also for wound healing applications. 1 μM concentration of 80 nm CuNPs was found not to be toxic to the cultured fibroblast, endothelial, and keratinocyte cells, and it supported endothelial cell migration and proliferation [425]. CuNPs may alter the structure of proteins and enzymes, affecting their normal functions and causing inactivation of bacterial functions at the injury site [426]. Chen et al. reported the cytotoxic effect of copper NPs towards human histolytic lymphoma (U937) and human cervical cancer cells by inducing apoptosis [427]. CuNPs were used to design a special drug delivery system for chemotherapy. For example, Figure 20 illustrates a mesoporous, upconversion, nanoparticles (mUCNPs)-based controlled-release drug carrier system exhibiting higher upconversion luminescence emission intensity [428].

2.2.10. Magnetic Nanoparticles

Magnetic elements (i.e., iron, nickel, cobalt, and their oxides) were utilized for the fabrication of nanomaterials for different medical applications [429,430] such as MRI, drug delivery, medical diagnostics, cancer therapy, biosensoring, and magneto-optic devices. Magnetic nanoparticles can be synthesized through different techniques including co-precipitation [431], microemulsion [432], hydrothermal synthesis [433], sol–gel process [434], polyol synthesis [435], flow injection [436], sonolysis/sonochemical methods [437], microwave irradiation [438], electrochemical synthesis [439], solvothermal method [440], chemical vapor deposition [441], laser pyrolysis [442], and green synthesis [443] using biomass or biotemplate.
Due to high magnetic flux density, magnetic nanoparticles were used for drug targeting [444] and bio-separation [445], including cell sorting [446]. Sun et al. [447] analyzed metallic, bi-metallic, magnetic cationic liposomes and superparamagnetic iron oxide nanoparticles for imaging and drug delivery. The surface of magnetic nanoparticles also needs to be functionalized to recognize specific targets (Figure 21) [448]. Polyethylene glycol (PEG) is one of the best polymers used for the functionalization of magnetic nanoparticles by surface modification [449]. Interestingly, surface modified magnetic nanoparticles reduce nonspecific interaction with biological molecules.
Magnetic manipulation is another important advantage of magnetic nanoparticles [450]. It is done by labeling cells with magnetic nanoparticles that can easily be controlled by remote control or external magnetic field [451]. The magnetic nanoparticles, which are usually smaller than 10 nm can be easily transported through skin lipid matrix and hair follicles to the stratum granulosum, where it is condensing between corneocytes [452].
In orthopedic surgery, implant-associated infection is a serious issue, as stated in the previous sections. Infection around a bone graft can lead to serious illness or failure of surgery. Drug-loaded Fe3O4 composites promote cell adhesion, proliferation, and osteogenic differentiation of hBMSCs [453,454,455]. In stem cell therapy for bone regeneration, an application of these NPs is the magnetic targeting of stem cells to the deserved locations, known as magnetic homing of stem cells. For example, penetration of ferumoxide-labeled cells into porous hydroxyapatite ceramic implanted in a rabbit ulnar defect was significantly facilitated by this approach, which improved bone formation even in the chronic process [456].

2.2.11. Summary and Statistical Analysis of the Survey on Inorganic Nanobiomaterials

The survey on inorganic nanobiomaterials presented in Section 2.2 covered some 230 articles, practically the same as its counterpart for organic biomaterials not counting about 30 articles on fabrication techniques of silica and magnetic nanoparticles.
While the number of technical papers appears to be rather uniformly distributed among the ten types of inorganic nanomaterials considered in this survey, it should be noted that most studies focused on nanoparticles and their functionalization for drug/gene/therapy delivery, cell labeling, biosensing, and bioimaging (75%), followed by studies on development and fabrication of new composite materials and scaffolds (25%).
Gold and titania present the largest variety of nanostructures and the latter material may also be available in the form of nanotubes. Gold nanoparticles may represent the best solution for most applications in view of the possibility of controlling size and dimensions of nanostructures as well as for their special physical properties (for example, local plasmon resonance). However, massive utilization of GNPs is obviously limited by the high cost of gold. Silica and titania nanoparticles also are widely utilized as standalone materials or in combination with gold and silver nanoparticles.
Similar to what has been observed for organic nanobiomaterials, a rapidly growing research area in the field of inorganic nanobiomaterials for bone tissue engineering is to hybridize them with other materials (e.g., chitosan, PLA, PLGA, collagen, and hydroxyapatite) to enhance mechanical properties, biocompatibility and osteogenetic properties of the modified materials. Development of high-performance scaffolds comprised of multiple materials is the final stage of this complicated process.

3. Applications of Nanobiomaterials

Nanobiomaterials have outstanding mechanical, chemical, electrical and optical properties, which make them highly suited for a variety of biological applications [70]. Nanotechnologies made it possible to develop new nanoscale materials (nanobiomaterials) with upgraded surface area to volume ratio, enabling more surface interactions [457,458,459]. As nanobiomaterials possess very specific properties that may be tailored to specific targets (i.e., solubility (for otherwise insoluble drugs), carriers for hydrophobic entities, multifunctional capability, active and passive targeting, ligands (size exclusion), and reduced toxicity), they have tremendous potential for disease identification (as imaging tools), care delivery, and prevention in new ways [107]. Nanobiomaterials are special kinds of materials that are introduced into the body for the treatment of damaged hard tissues [460]. The huge variety of biomedical applications of nanobiomaterials are illustrated in Figure 22 [428,461,462,463].
Nanobiomaterials have well-defined nanostructures such as size, shape, channels, pore structure, and surface domain [464]. Nanoscale dimension enables nanobiomaterials to develop critical physical and chemical characteristics that enhance their performance [465,466]. The properties and behaviors of nanobiomaterials, therefore, allow the diagnosis, monitoring, treatment, and prevention of diseases [467]. Nano-size materials show more catalytic reactions at their surface than macro-sized or conventional materials [468]. The nanoscale biomaterials create biomimetic feature towards most of the proteins which support further biological reactions such as cell attachment, growth, proliferation and generation of new tissue [36].

3.1. Bone Regeneration

A perfect bone and cartilage repair scaffold materials should neither suppress the activity of normal cells nor induce toxicity during and after implantation [469]. Figure 23 illustrates the basic cycle of tissue regeneration using nanobiomaterials or derived scaffolds [255].
The various synthetic nanostructured matrices are able to stimulate cell differentiation with a focus on preserving the structural features, composition, and biology of natural bone tissue [470]. The main constituents used so far in this regard are nano-hydroxyapatite [471], anodized titanium [472], collagen [473], and silver-incorporated calcium silicate. Nanobiomaterials (1–100 nm) generated from polymers, metals, ceramics, and composites act as effective constituents for hard tissue and play a significant role in osteointegration on nanostructured surfaces [474].
Alumina has been widely used for the fabrication of knee and hip joint prosthesis with low wear rates [475]. Bioactive glasses were used as a prosthesis for the restoration of the ossicular chain of the middle ear and oral implant to preserve the alveolar ridge from bone resorption [476,477]. Different metal oxides such as ZnO, Fe2O3, TiO2, and Al2O3, and polymers such as PLA, PGA, and their copolymers were used with bioactive glass systems for hard tissue engineering applications [478,479].

3.2. Drug Delivery

As mentioned above, various nanobiomaterials can be used for bone regeneration, prevention of infections, and osteointegration [480,481]. A nanoparticle that functions as carrier can stabilize the bioactive molecules through encapsulation [482], facilitating targeting cellular delivery and targeted drug release [483,484]. Nanospheres, tubes and capsules are widely accepted tools for targeted and sustained release drug delivery because of their small size and high specific surface area, which encapsulates the drug molecules and shows high reactivity to the surrounding tissues [485]. The materials selected for nanosphere fabrication depend on application principles and requirements. Some factors in this regard include size, drug characteristics, surface properties, biodegradability and biocompatibility of materials and drug release profile [486]. The 2D and 3D structures of scaffolds can be useful for the drug (poorly soluble drugs) loading purpose in tissue engineering (Figure 24).

3.3. Gene Delivery

The rapid development of nanotechnology made available novel DNA and RNA delivery systems for gene therapy (Figure 25) that can be used instead of viral vectors [487]. Gene therapy collectively refers to therapies aimed at manipulating gene expression in living organisms by supplying exogenous DNA or RNA that is incorporated or not incorporated to cure or prevent diseases [488]. There is great incentive to work towards safer and targeted viral vectors and to engineer more effective non-viral systems that can achieve secure, effective gene therapy in humans because of the enormous potential for gene therapies to influence medicine.
There are a number of nanocarriers used for gene delivery (Figure 25) [489] applications which are based on lipids [490,491], liposomes [492,493,494], dendrimers [495], polymers [496,497], graphene [498,499], carbon nanotubes (CNTs) [500,501], mesoporous silica [502], gold nanoparticles [503,504], magnetic nanoparticles [505,506], and other types of inorganic nanoparticles.
The number of clinical trials for cancer [507], liver disease [508], hemophilia [509], and bone regeneration [510,511,512] is continuously increasing due to the promising opportunity to correct gene disorders. Nanomaterials are used for gene delivery because of their small size and superior stability [513]. Before the use of these nanomaterials, surfaces need to be functionalized with small biomolecules or polymers to adapt their physiochemical properties such as hydrophobicity, charge density, and binding affinity [514,515]. Factors including molecular weight, biodegradability, rigidity, charge density, pKa value, solubility, crystallinity, and hydrophobicity ensure effective and safe gene delivery [516,517].
Surface-modified graphene oxide through cationic polymers such as polyethylenimine (PEI) provides a large surface area for the encapsulation of DNA molecules [518]. DNA/drug molecules attached graphene oxide conjugated provide an outstanding platform for the immobilization of nucleotides on its surface [519]. Mesoporous silica nanospheres (MSNs) and functionalized single-walled carbon nanotubes (SWCNT) represent an excellent gene delivery system [520]. In Figure 26, a potential route is recorded for the use of dendrimers as vectors of gene delivery. As plasmid DNA penetrates the cell membrane, it makes (in vitro) a complex between dendrimer and DNA (called dendriplex). This complex is transported through the blood system to the specific cell. Finally, the DNA moves through the cytoplasm to reach the nucleus for gene expression in series [521,522].

3.4. Anti-Infective Nanobiomaterials

Disease, injury, and trauma can lead to serious bacterial infections, which cause disease and adverse complications in host tissues and even death of patients [523]. Nanobiomaterials made from polymers, metals, and ceramics might be a potential source of infection when they are introduced into the body [524,525]. Virus and bacterial infections cause unregulated damage that leads to organ failure [526]. In polymeric biomaterials, the most common bacterial infections are powered by Staphylococcus epidermidis (S. epidermidis) from skin and Staphylococcus aureus (S. aureus), which may identified on metallic biomaterials [527]. Ceramics and metals can represent an alternative because of their resistance to infection. However, in presence of minor imperfections on the surface or microfractures, pathogens, such as bacteria, can form a colony [528]. Biomaterials from natural sources were used as alternative as scaffolds for promoting regeneration but they carry a risk for pathogenic transmission [529].

3.5. Nanobiomaterials for Coating

Micro/nanoscale tissue engineering scaffolds play a vital role on the organization of natural extracellular matrix [530]. Nanostructured 3D scaffolds enhance cell functioning, migration, differentiation, proliferation, and extracellular matrix formation [531].
Nanobiomaterials used for coatings include silica (SiO2), titania (TiO2), zirconia (ZrO2), alumina (Al2O3), zinc oxide (ZnO), CNT, graphene, and various combined oxides [532]. Simple calcium phosphate coating method on metals, glasses, inorganic ceramics and organic polymers (such as PLGA, PS, PP, and silicone), collagens, and silk fibers can improve biocompatibility or enhance the bioreactivity for orthopedic applications [494,533]. TiO2 and Al2O3 can be used as biologically active coating agents, supporting cell adhesion, growth, osteogenic differentiation, bone matrix production, and mineralization [534]. Nanostructured TiO2 has a positive effect on the performance of bone cells. TiO2 is available in the form of nanocrystals [535], nanofibers [536], nanoparticles [537], also immobilized on nanotubes [538]. TiO2 nanotube coating on any substrate enhances hydroxyapatite formation in SBF [539]. Nano silica coating on Ti-6Al-4V alloys generates apatite and supports adhesion and attachment of human osteoblast-like Saos-2 cells [540]. Nitinol coated stainless steel has shown enhanced biocompatibility but Ni ions produce an allergic response and toxicity [541]. Zirconia coated pure and yttrium-stabilized nanostructure promote deposition of apatite from SBF, which supports cell adhesion and growth [542]. Zinc oxide doped with alumina or functionalized with the silane coupling agent KH550 supports the proliferation of fibroblasts [543].
Carbon nanotubes have been used with various synthetic and natural polymers or minerals for the improvement of mechanical properties [544]. CNT and other nano-carbon forms stimulate cell adhesion and growth of osteogenic cells. Graphene-based films and composites used for biomaterial coatings can be obtained from pure or oxidized graphene. These graphene-based films improve the osteogenic differentiation manifested by collagen I and osteocalcin, high calcium phosphate deposition, and high alkaline phosphatase activity [545,546]. Due to the antimicrobial impression of graphene, graphene oxide (GO), and their derivatives, these materials can be used for implant coating [547]. Graphene oxide (GO) coating on the collagen scaffold induces morphological changes depending on GO concentration [548]. The application of GO improved physical properties like compressive strength as well as adsorption of Ca and proteins without changing porosity [549]. Graphene oxide-silk fibroin (GO-SF) composite used as an alternative to coating with collagen, showed improved biomechanical properties and proved could work in cellular environments [550].
HA can accelerate new bone formation by coating on titanium and tantalum scaffolds. It was demonstrated that after 6 weeks of implantation with titanium and tantalum scaffolds coated is possible to reach fully dense bone formation [551]. Calcium-phosphate-coated Fe foam showed better differentiation and proliferation rate of human mesenchymal stem cells than uncoated Fe foam [552]. Polymer-coated mesoporous silica nanoparticles are effective, cell-specific targeted chemotherapeutic agent delivery method [553]. In rat calvarial defects, HA-coated PLGA scaffolds alone promote bone regeneration and increased exposure to HA nanoparticles on the scaffold surface has been documented to result in accelerated bone deposition by local progenitors [554].

3.6. Nanostructured Scaffolds

Scaffolds are artificial constructs that provide support, tensile strength, and aid in tissue ingrowth [555]. They can also serve as carriers for growth factors, drugs and other required ingredients [556]. Scaffolds mimic the presence of extracellular matrix and allow the replacement of tissue without producing any harmful disturbance with respect to surrounding tissues. An ideal scaffold should be biocompatible, biodegradable, bioactive, non-toxic, mechanically stable, biodegradable, and bioresorbable (Figure 27) [557]. The amalgamation of organic and inorganic materials with scaffolds may enhance morphology and mechanical properties, thus supporting better cell attachment and proliferation [558].
Scaffold properties can be improved by using nanoparticles because organic and inorganic minerals in natural bone have nanoscale structures [559]. Many studies found that the addition of titanium and iron improve biological and mechanical properties such as collagen synthesis and apatite generation [560,561]. In addition, engineered nanofibrous scaffolds are also suitable for loadbearing applications and can replace natural extracellular matrix (ECM) with artificial ECM. The nanofibrous scaffold can therefore get a much more suitable environment for cellular growth and eventual regeneration of the bone [562]. Nanofiber-based scaffolds have been fabricated by using different synthetic polymers including PCL [563,564,565,566], PLLA [567,568], copolymer [569], PLGA [61], and chitosan [569].
Different kinds of metallic nanoparticles can be used for the synthesis of composite-based scaffolds with enhanced mechanical characteristics, cell adhesion, and bone tissue generating capacity [12]. The incorporation of titanium, iron, and alumoxane in a scaffold can improve mechanical properties, collagen synthesis, calcium deposition, and alkaline phosphatase activity [561].
Graphene and its derivatives were used as reinforcement material for fibrous scaffolds, films, and hydrogels [570]. The graphene and graphene oxide incorporation into hydrogels yield enhancements in mechanical properties without producing adverse effects on encapsulated fibroblast cells [571]. Carbon-based nanomaterials can be used to improve mechanical strength of scaffolds [572]. Alumina, titania, bioglass, and hydroxyapatite support osteoblast adhesion and growth [573].
Nanobiomaterial-based composite structures are an efficient platform for the synthesis of engineered scaffolds and application in bone tissue engineering (Figure 28) [574]. Nanocomposite-based scaffolds exhibit inherent characteristics such as porous and rough surface and increased wettability, which promote fast bone regeneration. These nanocomposite-based scaffolds provide a porous structure for nutrients exchange and increased protein adsorption. Scaffolds exhibited micro/nano-scaled porous structural pathway for cell–scaffold interaction and integrin-triggered signaling pathway. The nanoscale features support bone cells (osteoblast) and bone-derived stem cells proliferation, migration, cell signaling, stem cell fate, and genetic cell fate. The nanobiomaterials based scaffold have notable mechanical and biological advantages and can induce bone tissue regeneration [531]. The nanostructured materials improve morphological characteristics of scaffolds that may enhance osteoinduction, bone cell attachment, differentiation, proliferation, and natural bone cell growth within the extracellular matrix [12].
Hydroxyapatite (HA) has attracted attention because of its inherent biological compatibility and bone conduction as well as its similarity with bone minerals [575]. For this reason, HA was combined with a number of synthetic and natural polymers such as polycaprolactone [576], poly (lactic acid) (PLA) [577], polyethylene, poly(lactic-co-glycolic acid) (PLGA) [203], collagen [578], gelatin [148], and chitosan [579] to fabricate scaffolds. These composite based scaffolds showed improved mechanical properties, porosity and biocompatibility without or with significantly less adverse effects.

3.7. Bone Cancer Therapy

Cancer is the uncontrolled growth of tissues that could lead to invasion into other organs without proper regulation or differentiation [580]. Conventional cancer therapy is associated with multiple adverse side effects [581]. Bone metastases or “bone mets” occur when cancer cells from the primary tumor relocate to the bone and also spread in the prostate, breast, and lung, which leads to painful (75% of patients) and devastating skeletal-related events (SREs) [582,583]. Depending on the stage of the disease, history, and the overall health of the patient, disease management includes a combination of therapies as shown in Figure 29 [584].
The different types of nanoparticles (NPs) used as carriers for small-molecule drugs, proteins, and nucleic acids [585] can be localized to specific disease locations for the treatment of bone metastasis [586]. Nanoparticles also improve the efficiency of other methods used for treating bone metastasis [587]. The effectiveness of NPs depends on their accumulation in vascularized solid tumors via the enhanced permeability and retention (EPR) effect [588]. A wide variety of nanomaterials have been developed in the 1 to 100 nm range and include various anti-tumor drugs (Figure 30) by fine-tuning the chemical structure, scale, and shape (morphology) that can regulate the nanomaterials’ functionality [589].

4. Counter-Indications

There is inevitably some sort of interaction between the organic or inorganic materials and the biological environment when individual or composite biomaterials are put in contact with the tissues and fluids of the human body. The basic clinical research may decide that the materials should not cause any local or systemic adverse reactions. Recent studies exposed that nanosized materials can easily penetrate biological membranes of normal cells and enter vascular system to facilitate redistribution in different tissues. Nanomaterials, which by themselves are not very harmful, could become toxic if are ingested in higher concentration. The toxicity of metal based biomaterials to the liver is an important basis for the safety assessment of nanosized materials. Metal based nanoparticles release ions which may enter the cells and affect the functions of organelles, leading to liver injury. Various factors including amount, composition, pH, and fabrication techniques may decide the compatibility and cytotoxicity of biomaterials. The research is ongoing to improve the existing technologies which may produce highly compatible substitutes without producing adverse effects.

5. Conclusions

This review article provided a broad overview of the various types of organic and inorganic nanobiomaterials and their applications in the field of hard tissue engineering. Besides classifying nanobiomaterials, the survey covered several key aspects like bone/cartilage regeneration, drug/gene delivery, anti-infection properties, coatings, scaffold fabrication, and cancer therapy. A total of 550 articles selected by means of web search engines widely used in science and engineering were reviewed in this study. Interestingly, the number of reviewed articles was approximately the same for organic and inorganic biomaterials.
Biomaterials science is a highly multidisciplinary area. Developments in life science and nanotechnology enabled scientists and engineers to conceive new designs and improve the existing bone structure. For example, advances in nanotechnology allowed for the development of novel methods for fabricating new nanostructured scaffolds possessing a higher efficiency in tissue regeneration.
Nanomaterials represent an excellent tool for research and therapeutic approaches in bone tissue engineering. Organic nanomaterials are more biocompatible, nontoxic, and help more with cell regeneration than inorganic nanomaterials. However, inorganic nanomaterials provide better mechanical strength and inertness to chemical agent. From the references cited in this survey it appears that nanoparticles, graphene and nanocomposites are the most diffused types of nanostructures used for hard tissue applications. An important research trend which results in a rapidly growing number of published articles is the development of new composite nanobiomaterials especially for scaffold applications.
Interactions between bone cells and nanomaterials depend on the composition of nanoparticles. Proper selection of nanoparticles may result in faster bone regeneration and recovery. Besides composition, the overall performance of a nanobiomaterial depends on porosity, microstructure, mechanical properties and functionality. Nanomaterials-based scaffolds also play a major role in three-dimensional tissue growth. Nanostructural modifications provide a favorable environment for bone regeneration.
The survey presented in the article proved that tissue engineering supports (i) application of engineering design methods to functionally engineered tissues, (ii) development of novel biomaterials for constructing scaffolds that mimic extracellular matrix, and (iii) creating artificial microenvironments. Nanobiomaterials represent an excellent tool for research and therapeutic approaches in bone tissue engineering. However, further investigations should be aimed at producing advanced nanobiomaterials suitable for hard tissue engineering that can fill the gap between biomaterial fabrication and clinical implementation.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Patra, J.K.; Das, G.; Fraceto, L.F.; Campos, E.V.R.; Rodriguez-Torres, M.d.P.; 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] [Green Version]
  2. Parratt, K.; Yao, N. Nanostructured biomaterials and their applications. Nanomaterials 2013, 3, 242–271. [Google Scholar] [CrossRef] [PubMed]
  3. Lei, Y.; Lijuan, Z.; Webster, T.J. Nanobiomaterials: State of the art and future trends. Adv. Eng. Mater. 2011, 13, B197–B217. [Google Scholar]
  4. Ramos, A.P.; Cruz, M.A.E.; Tovani, C.B.; Ciancaglini, P. Biomedical applications of nanotechnology. Biophys. Rev. 2017, 9, 79–89. [Google Scholar] [CrossRef] [PubMed]
  5. Capek, I.B.T.-S. Nanocomposite Structures and Dispersions; Elsevier: Amsterdam, The Netherlands, 2006; Chapter 1; Volume 23, pp. 1–69. [Google Scholar]
  6. Florencio-Silva, R.; Rodrigues, G.; Sasso-Cerri, E.; Simões, M.J.; Cerri, P.S.; Cells, B. Biology of bone tissue: Structure, function, and factors that influence bone cells. BioMed Res. Int. 2015, 2015, 421746. [Google Scholar] [CrossRef] [Green Version]
  7. Eliaz, N.; Metoki, N. Calcium phosphate bioceramics: A review of their history, structure, properties, coating technologies and biomedical applications. Materials 2017, 10, 334. [Google Scholar] [CrossRef] [Green Version]
  8. McMahon, R.; Wang, L.; Skoracki, R.; Mathur, A. Development of nanomaterials for bone repair and regeneration. J. Biomed. Mater. Res. Part B Appl. Biomater. 2013, 101, 387–397. [Google Scholar] [CrossRef] [PubMed]
  9. Kalenderer, Ö.; Turgut, A. Bone. In Musculoskeletal Research and Basic Science; Korkusuz, F., Ed.; Springer International Publishing: Cham, Switzerland, 2016; Chapter 18; pp. 303–321. [Google Scholar]
  10. Farbod, K.; Nejadnik, M.R.; Jansen, J.A.; Leeuwenburgh, S.C.G. Interactions between inorganic and organic phases in bone tissue as a source of inspiration for design of novel nanocomposites. Tissue Eng. Part B Rev. 2013, 20, 173–188. [Google Scholar] [CrossRef] [Green Version]
  11. Feng, Q.L.; Wu, J.; Chen, G.Q.; Cui, F.Z.; Kim, T.N.; Kim, J.O. A mechanistic study of the antibacterial effect of silver ions on Escherichia coli and Staphylococcus aureus. J. Biomed. Mater. Res. 2000, 52, 662–668. [Google Scholar] [CrossRef]
  12. Walmsley, G.G.; Mcardle, A.; Tevlin, R.; Momeni, A.; Atashroo, D.; Hu, M.S.; Feroze, A.H.; Wong, V.W.; Lorenz, P.H.; Longaker, M.T.; et al. Nanotechnology in bone tissue engineering. Nanomedicine 2015, 11, 1253–1263. [Google Scholar] [CrossRef] [Green Version]
  13. Mohamed, A.M. An overview of bone cells and their regulating factors of differentiation. Malays. J. Med. Sci. 2008, 15, 4–12. [Google Scholar] [PubMed]
  14. Kargozar, S.; Mozafari, M.; Hamzehlou, S.; Brouki Milan, P.; Kim, H.-W.; Baino, F. Bone tissue engineering using human cells: A Comprehensive review on recent trends, current prospects, and recommendations. Appl. Sci. 2019, 9, 174. [Google Scholar] [CrossRef] [Green Version]
  15. Combes, C.; Cazalbou, S.; Rey, C. Apatite biominerals. Minerals 2016, 6, 34. [Google Scholar] [CrossRef] [Green Version]
  16. Glimcher, M. Bone: Nature of the calcium phosphate crystals and cellular, structural, and physical chemical mechanisms in their formation. Rev. Miner. Geochem. 2006, 64, 223–282. [Google Scholar] [CrossRef]
  17. Stevens, M.M. Biomaterials for bone tissue engineering. Mater. Today 2008, 11, 18–25. [Google Scholar] [CrossRef]
  18. Liu, S.; Gong, W.; Dong, Y.; Hu, Q.; Chen, X.; Gao, X. The effect of submicron bioactive glass particles on in vitro osteogenesis. RSC Adv. 2015, 5, 38830–38836. [Google Scholar] [CrossRef]
  19. Cheng, C.W.; Solorio, L.D.; Alsberg, E. Decellularized tissue and cell-derived extracellular matrices as scaffolds for orthopaedic tissue engineering. Biotechnol. Adv. 2014, 32, 462–484. [Google Scholar] [CrossRef] [Green Version]
  20. Nie, X.; Wang, D.-A. Decellularized orthopaedic tissue-engineered grafts: Biomaterial scaffolds synthesised by therapeutic cells. Biomater. Sci. 2018, 6, 2798–2811. [Google Scholar] [CrossRef]
  21. Amini, A.R.; Laurencin, C.T.; Nukavarapu, S.P. Bone tissue engineering: Recent advances and challenges. Crit. Rev. Biomed. Eng. 2012, 40, 363–408. [Google Scholar] [CrossRef] [Green Version]
  22. Zhang, L.; Webster, T.J. Nanotechnology and nanomaterials: Promises for improved tissue regeneration. Nano Today 2009, 4, 66–80. [Google Scholar] [CrossRef]
  23. Kim, S.; Kim, D.; Cho, S.W.; Kim, J.; Kim, J.-S. Highly efficient RNA-guided genome editing in human cells via delivery of purified Cas9 ribonucleoproteins. Genome Res. 2014, 24, 1012–1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Wang, C.; Shen, H.; Tian, Y.; Xie, Y.; Li, A.; Ji, L.; Niu, Z.; Wu, D.; Qiu, D. Bioactive nanoparticle—Gelatin composite scaff old with mechanical performance comparable to cancellous bones. ACS Appl. Mater. Interfaces 2014, 6, 13061–13068. [Google Scholar] [CrossRef] [PubMed]
  25. Dhivya, S.; Ajita, J.; Selvamurugan, N. Metallic nanomaterials for bone tissue engineering. J. Biomed. Nanotechnol. 2015, 11, 1675–1700. [Google Scholar] [CrossRef] [PubMed]
  26. Moreno-Vega, A.I.; Gómez-Quintero, T.; Nuñez-Anita, R.E.; Acosta-Torres, L.S.; Castaño, V. Polymeric and ceramic nanoparticles in biomedical applications. J. Nanotechnol. 2012, 2012, 936041. [Google Scholar] [CrossRef] [Green Version]
  27. Kim, N.J.; Lee, S.J.; Atala, A. Biomedical nanomaterials in tissue engineering. In Woodhead Publishing Series in Biomaterials; Gaharwar, A.K., Sant, S., Hancock, M.J., Hacking, S.A.B.T.-N., Eds.; Woodhead Publishing: Cambridge, UK, 2013; pp. 1e–25e. [Google Scholar]
  28. Gajanan, K.; Tijare, S.N. Applications of nanomaterials. Mater. Today Proc. 2018, 5, 1093–1096. [Google Scholar] [CrossRef]
  29. Wang, Y.; Xia, Y. Bottom-up and top-down approaches to the synthesis of monodispersed spherical colloids of low melting-point metals. Nano Lett. 2004, 4, 2047–2050. [Google Scholar] [CrossRef]
  30. Saiz, E.; Zimmermann, E.A.; Lee, J.S.; Wegst, U.G.K.; Tomsia, A.P. Perspectives on the role of nanotechnology in bone tissue engineering. Dent. Mater. 2013, 29, 103–115. [Google Scholar] [CrossRef] [Green Version]
  31. Gardin, C.; Ferroni, L.; Favero, L.; Stellini, E.; Stomaci, D.; Sivolella, S.; Bressan, E.; Zavan, B. Nanostructured biomaterials for tissue engineered bone tissue reconstruction. Int. J. Mol. Sci. 2012, 13, 737–757. [Google Scholar]
  32. España-Sánchez, B.L.; Cruz-Soto, M.E.; Elizalde-Pena, E.A.; Sabasflores-Benitez, S.; Roca-Aranda, A.; Esquivel-Escalante, K.; Luna-Barcenas, G. Trends in Tissue Regeneration: Bio-nanomaterials. In Tissue Rigeneration; El-Sayed Kaoud, H.A., Ed.; IntechOpen: Rijeka, Croatia, 2018. [Google Scholar]
  33. Nikolova, M.P.; Chavali, M.S. Recent advances in biomaterials for 3D scaffolds: A review. Bioact. Mater. 2019, 4, 271–292. [Google Scholar] [CrossRef]
  34. Zhao, Z.Y.; Liang, L.; Fan, X.; Yu, Z.; Hotchkiss, A.T.; Wilk, B.J.; Eliaz, I. The role of modified citrus pectin as an effective chelator of lead in children hospitalized with toxic lead levels. Altern. Ther. Health Med. 2008, 14, 34–38. [Google Scholar]
  35. Zheng, K.; Balasubramanian, P.; Paterson, T.E.; Stein, R.; MacNeil, S.; Fiorilli, S.; Vitale-Brovarone, C.; Shepherd, J.; Boccaccini, A.R. Ag modified mesoporous bioactive glass nanoparticles for enhanced antibacterial activity in 3D infected skin model. Mater. Sci. Eng. C 2019, 103, 109764. [Google Scholar] [CrossRef]
  36. Chen, F.-M.; Liu, X. Advancing biomaterials of human origin for tissue engineering. Prog. Polym. Sci. 2016, 53, 86–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. John, Ł. Selected developments and medical applications of organic–inorganic hybrid biomaterials based on functionalized spherosilicates. Mater. Sci. Eng. C 2018, 88, 172–181. [Google Scholar] [CrossRef] [PubMed]
  38. Fattahi, P.; Yang, G.; Kim, G.; Abidian, M.R. Biomaterials: A review of organic and inorganic biomaterials for neural interfaces. Adv. Mater. 2014, 26, 1793. [Google Scholar] [CrossRef]
  39. Virlan, M.J.R.; Miricescu, D.; Radulescu, R.; Sabliov, C.M.; Totan, A.; Calenic, B.; Greabu, M. Organic nanomaterials and their applications in the treatment of oral diseases. Molecules 2016, 21, 207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Chellan, P.; Sadler, P.J. The elements of life and medicines. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2015, 373, 20140182. [Google Scholar] [CrossRef] [PubMed]
  41. Kiaie, N.; Aavani, F.; Razavi, M. 2—particles/fibers/bulk. In Development and Evaluation; Razavi, M., Ed.; Woodhead Publishing: Cambridge, UK, 2017; pp. 7–25. [Google Scholar]
  42. Wang, G.; Mostafa, N.Z.; Incani, V.; Kucharski, C.; Uludaĝ, H. Bisphosphonate-decorated lipid nanoparticles designed as drug carriers for bone diseases. J. Biomed. Mater. Res. Part A 2012, 100, 684–693. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, G.; Guo, B.; Wu, H.; Tang, T.; Zhang, B.T.; Zheng, L.; He, Y.; Yang, Z.; Pan, X.; Chow, H.; et al. A delivery system targeting bone formation surfaces to facilitate RNAi-based anabolic therapy. Nat. Med. 2012, 18, 307–314. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, J.; Zhang, H.; Dong, Y.; Jin, Y.; Hu, X.; Cai, K.; Ma, J.; Wu, G. Bi-directionally selective bone targeting delivery for anabolic and antiresorptive drugs: A novel combined therapy for osteoporosis? Med. Hypotheses 2014, 83, 694–696. [Google Scholar] [CrossRef]
  45. Hirsjärvi, S.; Sancey, L.; Dufort, S.; Belloche, C.; Vanpouille-Box, C.; Garcion, E.; Coll, J.-L.; Hindré, F.; Benoît, J.-P. Effect of particle size on the biodistribution of lipid nanocapsules: Comparison between nuclear and fluorescence imaging and counting. Int. J. Pharm. 2013, 453, 594–600. [Google Scholar] [CrossRef] [Green Version]
  46. Nahar, M.; Dutta, T.; Murugesan, S.; Asthana, A.; Mishra, D.; Rajkumar, V.; Tare, M.; Saraf, S.; Jain, N.K. Functional polymeric nanoparticles: An efficient and promising tool for active delivery of bioactives. Crit. Rev. Ther. Drug Carr. Syst. 2006, 23, 259–318. [Google Scholar] [CrossRef] [PubMed]
  47. An, S.Y.; Bui, M.-P.N.; Nam, Y.J.; Han, K.N.; Li, C.A.; Choo, J.; Lee, E.K.; Katoh, S.; Kumada, Y.; Seong, G.H. Preparation of monodisperse and size-controlled poly(ethylene glycol) hydrogel nanoparticles using liposome templates. J. Colloid Interface Sci. 2009, 331, 98–103. [Google Scholar] [CrossRef] [PubMed]
  48. Monteiro, N.; Martins, A.; Reis, R.L.; Neves, N.M. Liposomes in tissue engineering and regenerative medicine. J. R. Soc. Interface 2014, 11, 281–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Qi, R.; Majoros, I.; Misra, A.C.; Koch, A.E.; Campbell, P.; Marotte, H.; Bergin, I.L.; Cao, Z.; Goonewardena, S.; Morry, J.; et al. Folate receptor-targeted dendrimer-methotrexate conjugate for inflammatory arthritis. J. Biomed. Nanotechnol. 2014, 11, 1370–1384. [Google Scholar] [CrossRef]
  50. Duncan, R.; Izzo, L. Dendrimer biocompatibility and toxicity. Adv. Drug Deliv. Rev. 2005, 57, 2215–2237. [Google Scholar] [CrossRef]
  51. De la Riva, B.; Sánchez, E.; Hernández, A.; Reyes, R.; Tamimi, F.; López-Cabarcos, E.; Delgado, A.; vora, C. Local controlled release of VEGF and PDGF from a combined brushite-chitosan system enhances bone regeneration. J. Control. Release 2010, 143, 45–52. [Google Scholar] [CrossRef]
  52. Rampino, A.; Borgogna, M.; Blasi, P.; Bellich, B.; Cesàro, A. Chitosan nanoparticles: Preparation, size evolution and stability. Int. J. Pharm. 2013, 455, 219–228. [Google Scholar] [CrossRef]
  53. Vedakumari, W.S.; Prabu, P.; Sastry, T.P. Chitosan-fibrin nanocomposites as drug delivering and wound healing materials. J. Biomed. Nanotechnol. 2015, 11, 657–667. [Google Scholar] [CrossRef]
  54. Shim, K.; Won, H.S.; Sang, Y.L.; Sang, H.L.; Seong, J.H.; Myung, C.L.; Lee, S.J. Chitosan nano-/microfibrous double-layered membrane with rolled-up three-dimensional structures for chondrocyte cultivation. J. Biomed. Mater. Res. Part A 2009, 90, 595–602. [Google Scholar] [CrossRef]
  55. Poth, N.; Seiffart, V.; Gross, G.; Menzel, H.; Dempwolf, W. Biodegradable chitosan nanoparticle coatings on titanium for the delivery of BMP-2. Biomolecules 2015, 5, 3–19. [Google Scholar] [CrossRef] [Green Version]
  56. Nagarajan, U.; Kawakami, K.; Zhang, S.; Chandrasekaran, B. Fabrication of solid collagen nanoparticles using electrospray deposition. Chem. Pharm. Bull. 2014, 62, 422–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Liu, Y.; Lu, Y.; Tian, X.; Cui, G.; Zhao, Y.; Yang, Q.; Yu, S.; Xing, G.; Zhang, B. Segmental bone regeneration using an rhBMP-2-loaded gelatin/nanohydroxyapatite/fibrin scaffold in a rabbit model. Biomaterials 2009, 30, 6276–6285. [Google Scholar] [CrossRef] [PubMed]
  58. Jahanshahi, M.; Sanati, M.H.; Hajizadeh, S.; Babaei, Z. Gelatin nanoparticle fabrication and optimization of the particle size. Phys. Status Solidi 2008, 205, 2898–2902. [Google Scholar] [CrossRef]
  59. Miller, D.C.; Thapa, A.; Haberstroh, K.M.; Webster, T.J. Endothelial and vascular smooth muscle cell function on poly(lactic-co-glycolic acid) with nano-structured surface features. Biomaterials 2004, 25, 53–61. [Google Scholar] [CrossRef]
  60. Danhier, F.; Ansorena, E.; Silva, J.M.; Coco, R.; Le Breton, A.; Préat, V. PLGA-based nanoparticles: An overview of biomedical applications. J. Control. Release 2012, 161, 505–522. [Google Scholar] [CrossRef]
  61. Pattison, M.A.; Wurster, S.; Webster, T.J.; Haberstroh, K.M. Three-dimensional, nano-structured PLGA scaffolds for bladder tissue replacement applications. Biomaterials 2005, 26, 2491–2500. [Google Scholar] [CrossRef]
  62. Soundrapandian, C.; Mahato, A.; Kundu, B.; Datta, S.; Sa, B.; Basu, D. Development and effect of different bioactive silicate glass scaffolds: In vitro evaluation for use as a bone drug delivery system. J. Mech. Behav. Biomed. Mater. 2014, 40, 1–12. [Google Scholar] [CrossRef] [PubMed]
  63. Baldrighi, M.; Trusel, M.; Tonini, R.; Giordani, S. Carbon Nanomaterials Interfacing with Neurons: An In vivo Perspective. Front. Neurosci. 2016, 10, 250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Gorain, B.; Choudhury, H.; Pandey, M.; Kesharwani, P.; Abeer, M.M.; Tekade, R.K.; Hussain, Z. Carbon nanotube scaffolds as emerging nanoplatform for myocardial tissue regeneration: A review of recent developments and therapeutic implications. Biomed. Pharmacother. 2018, 104, 496–508. [Google Scholar] [CrossRef] [PubMed]
  65. Zhu, Z. An overview of carbon nanotubes and graphene for biosensing applications. Nanomicro Lett. 2017, 9, 25. [Google Scholar] [CrossRef] [Green Version]
  66. Guo, Q.; Shen, X.; Li, Y.; Xu, S. Carbon nanotubes-based drug delivery to cancer and brain. Curr. Med. Sci. 2017, 37, 635–641. [Google Scholar] [CrossRef] [PubMed]
  67. Fahy, E.; Subramaniam, S.; Murphy, R.C.; Nishijima, M.; Raetz, C.R.H.; Shimizu, T.; Spener, F.; van Meer, G.; Wakelam, M.J.O.; Dennis, E.A. Update of the LIPID MAPS comprehensive classification system for lipids. J. Lipid Res. 2009, 50, S9–S14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Vemuri, S.; Rhodes, C.T. Preparation and characterization of liposomes as therapeutic delivery systems: A review. Pharm. Acta Helv. 1995, 70, 95–111. [Google Scholar] [CrossRef]
  69. Subramaniam, S.; Fahy, E.; Gupta, S.; Sud, M.; Byrnes, R.W.; Cotter, D.; Dinasarapu, A.R.; Maurya, M.R. Bioinformatics and systems biology of the lipidome. Chem. Rev. 2011, 111, 6452–6490. [Google Scholar] [CrossRef] [Green Version]
  70. Khan, I.; Saeed, K.; Khan, I. Nanoparticles: Properties, applications and toxicities. Arab. J. Chem. 2019, 12, 908–931. [Google Scholar] [CrossRef]
  71. Carbone, C.; Leonardi, A.; Cupri, S.; Puglisi, G.; Pignatello, R. Pharmaceutical and biomedical applications of lipid-based nanocarriers. Pharm. Pat. Anal. 2014, 3, 199–215. [Google Scholar] [CrossRef]
  72. Bozzuto, G.; Molinari, A. Liposomes as nanomedical devices. Int. J. Nanomed. 2015, 10, 975–999. [Google Scholar] [CrossRef] [Green Version]
  73. Puri, A.; Loomis, K.; Smith, B.; Lee, J.-H.; Yavlovich, A.; Heldman, E.; Blumenthal, R. Lipid-based nanoparticles as pharmaceutical drug carriers: From concepts to clinic. Crit. Rev. Ther. Drug Carr. Syst. 2009, 26, 523–580. [Google Scholar] [CrossRef] [Green Version]
  74. Rai, R.; Alwani, S.; Badea, I. Polymeric nanoparticles in gene therapy: New avenues of design and optimization for delivery applications. Polymers 2019, 11, 745. [Google Scholar] [CrossRef] [Green Version]
  75. Dolatabadi, J.E.N.; Omidi, Y. Solid lipid-based nanocarriers as efficient targeted drug and gene delivery systems. TrAC Trends Anal. Chem. 2016, 77, 100–108. [Google Scholar] [CrossRef]
  76. Mukherjee, S.; Ray, S.; Thakur, R.S. Solid lipid nanoparticles: A modern formulation approach in drug delivery system. Indian J. Pharm. Sci. 2009, 71, 349–358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Tamjidi, F.; Shahedi, M.; Varshosaz, J.; Nasirpour, A. Nanostructured lipid carriers (NLC): A potential delivery system for bioactive food molecules. Innov. Food Sci. Emerg. Technol. 2013, 19, 29–43. [Google Scholar] [CrossRef]
  78. Mouzouvi, C.R.A.; Umerska, A.; Bigot, A.K.; Saulnier, P. Surface active properties of lipid nanocapsules. PLoS ONE 2017, 12, e0179211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Irby, D.; Du, C.; Li, F. Lipid—Drug conjugate for enhancing drug delivery. Mol. Pharm. 2017, 14, 1325–1338. [Google Scholar] [CrossRef] [Green Version]
  80. Talegaonkar, S.; Bhattacharyya, A. Potential of lipid nanoparticles (SLNs and NLCs) in enhancing oral bioavailability of drugs with poor intestinal permeability. AAPS PharmSciTech 2019, 20, 121. [Google Scholar] [CrossRef]
  81. Dave, V.; Tak, K.; Sohgaura, A.; Gupta, A.; Sadhu, V.; Reddy, K.R. Lipid-polymer hybrid nanoparticles: Synthesis strategies and biomedical applications. J. Microbiol. Methods 2019, 160, 130–142. [Google Scholar] [CrossRef]
  82. Bangham, A.D.; Standish, M.M.; Watkins, J.C. Diffusion of univalent ions across the lamellae of swollen phospholipids. J. Mol. Biol. 1965, 13, 238-IN27. [Google Scholar] [CrossRef]
  83. Akbarzadeh, A.; Rezaei-Sadabady, R.; Davaran, S.; Joo, S.; Zarghami, N.; Hanifehpour, Y.; Samiei, M.; Kouhi, M.; Nejati, K. Liposome: Classification, preparation, and applications. Nanoscale Res. Lett. 2013, 8, 102. [Google Scholar] [CrossRef] [Green Version]
  84. Daraee, H.; Etemadi, A.; Kouhi, M.; Alimirzalu, S.; Akbarzadeh, A. Application of liposomes in medicine and drug delivery. Artif. Cells Nanomed. Biotechnol. 2016, 44, 381–391. [Google Scholar] [CrossRef]
  85. Fenske, D.B.; Chonn, A.; Cullis, P.R. Liposomal nanomedicines: An emerging field. Toxicol. Pathol. 2008, 36, 21–29. [Google Scholar] [CrossRef]
  86. Immordino, M.L.; Dosio, F.; Cattel, L. Stealth liposomes: Review of the basic science, rationale, and clinical applications, existing and potential. Int. J. Nanomed. 2006, 1, 297–315. [Google Scholar]
  87. La, W.-G.; Jin, M.; Park, S.; Yoon, H.-H.; Jeong, G.-J.; Bhang, S.; Park, H.; Char, K.; Kim, B.-S. Delivery of bone morphogenetic protein-2 and substance P using graphene oxide for bone regeneration. Int. J. Nanomed. 2014, 9 (Suppl. 1), 107–116. [Google Scholar]
  88. Du, G.-Y.; He, S.-W.; Sun, C.-X.; Mi, L.-D. Bone morphogenic Protein-2 (rhBMP2)-loaded silk fibroin scaffolds to enhance the osteoinductivity in bone tissue engineering. Nanoscale Res. Lett. 2017, 12, 573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Takahashi, M.; Inafuku, K.; Miyagi, T.; Oku, H.; Wada, K.; Imura, T.; Kitamoto, D. Efficient preparation of liposomes encapsulating food materials using lecithins by a mechanochemical method. J. Oleo Sci. 2007, 56, 35–42. [Google Scholar] [CrossRef] [Green Version]
  90. Zhu, C.T.; Xu, Y.Q.; Shi, J.; Li, J.; Ding, J. Liposome combined porous β-TCP scaffold: Preparation, characterization, and anti-biofilm activity. Drug Deliv. 2010, 17, 391–398. [Google Scholar] [CrossRef]
  91. Li, Y.; Bai, Y.; Pan, J.; Wang, H.; Li, H.; Xu, X.; Fu, X.; Shi, R.; Luo, Z.; Li, Y.; et al. A hybrid 3D-printed aspirin-laden liposome composite scaffold for bone tissue engineering. J. Mater. Chem. B 2019, 7, 619–629. [Google Scholar] [CrossRef]
  92. Sou, K.; Inenaga, S.; Takeoka, S.; Tsuchida, E. Loading of curcumin into macrophages using lipid-based nanoparticles. Int. J. Pharm. 2008, 352, 287–293. [Google Scholar] [CrossRef]
  93. Mayer, L.D.; Bally, M.B.; Cullis, P.R. Uptake of adriamycin into large unilamellar vesicles in response to a pH gradient. Biochim. Biophys. Acta Biomembr. 1986, 857, 123–126. [Google Scholar] [CrossRef]
  94. Cullis, P.R.; Hope, M.J.; Bally, M.B.; Madden, T.D.; Mayer, L.D.; Fenske, D.B. Influence of pH gradients on the transbilayer transport of drugs, lipids, peptides and metal ions into large unilamellar vesicles. Biochim. Biophys. Acta Rev. Biomembr. 1997, 1331, 187–211. [Google Scholar] [CrossRef]
  95. Dang, M.; Saunders, L.; Niu, X.; Fan, Y.; Ma, P.X. Biomimetic delivery of signals for bone tissue engineering. Bone Res. 2018, 6, 25. [Google Scholar] [CrossRef]
  96. Kulkarni, M.; Greiser, U.; O’Brien, T.; Pandit, A. Liposomal gene delivery mediated by tissue-engineered scaffolds. Trends Biotechnol. 2010, 28, 28–36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Abbasi, E.; Aval, S.F.; Akbarzadeh, A.; Milani, M.; Nasrabadi, H.T.; Joo, S.W.; Hanifehpour, Y.; Nejati-Koshki, K.; Pashaei-Asl, R. Dendrimers: Synthesis, applications, and properties. Nanoscale Res. Lett. 2014, 9, 247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Majoral, J.-P.; Caminade, A.-M. Dendrimers containing heteroatoms (Si, P, B, Ge, or Bi). Chem. Rev. 1999, 99, 845–880. [Google Scholar] [CrossRef] [PubMed]
  99. Bosman, A.W.; Janssen, H.M.; Meijer, E.W. About dendrimers:  Structure, physical properties, and applications. Chem. Rev. 1999, 99, 1665–1688. [Google Scholar] [CrossRef] [PubMed]
  100. Joshi, N.; Grinstaff, M. Applications of dendrimers in tissue engineering. Curr. Top. Med. Chem. 2008, 8, 1225–1236. [Google Scholar] [CrossRef] [PubMed]
  101. Barrett, T.; Ravizzini, G.; Choyke, P.L.; Kobayashi, H. Dendrimers in medical nanotechnology. IEEE Eng. Med. Biol. Mag. 2009, 28, 12–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Liu, Y.; Tee, J.K.T.; Chiu, G.N.C. Dendrimers in oral drug delivery application: Current explorations, toxicity issues and strategies for improvement. Curr. Pharm. Des. 2015, 21, 2629–2642. [Google Scholar] [CrossRef] [PubMed]
  103. Courtenay, J.C.; Deneke, C.; Lanzoni, E.M.; Costa, C.A.; Bae, Y.; Scott, J.L.; Sharma, R.I. Modulating cell response on cellulose surfaces; tunable attachment and scaffold mechanics. Cellulose 2018, 25, 925–940. [Google Scholar] [CrossRef] [Green Version]
  104. Zhou, L.; Shan, Y.; Hu, H.; Yu, B.Y.; Cong, H. Synthesis and biomedical applications of dendrimers. Curr. Org. Chem. 2018, 22, 600–612. [Google Scholar] [CrossRef]
  105. Opina, A.C.; Wong, K.J.; Griffiths, G.L.; Turkbey, B.I.; Bernardo, M.; Nakajima, T.; Kobayashi, H.; Choyke, P.L.; Vasalatiy, O. Preparation and long-term biodistribution studies of a PAMAM dendrimer G5–Gd-BnDOTA conjugate for lymphatic imaging. Nanomedicine 2014, 10, 1423–1437. [Google Scholar] [CrossRef] [Green Version]
  106. Shadrack, D.M.; Swai, H.S.; Munissi, J.J.E.; Mubofu, E.B.; Nyandoro, S.S. Polyamidoamine dendrimers for enhanced solubility of small molecules and other desirable properties for site specific delivery: Insights from experimental and computational studies. Molecules 2018, 23, 1419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Din, F.U.; Aman, W.; Ullah, I.; Qureshi, O.S.; Mustapha, O.; Shafique, S.; Zeb, A. Effective use of nanocarriers as drug delivery systems for the treatment of selected tumors. Int. J. Nanomed. 2017, 12, 7291–7309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Pooja, D.; Sistla, R.; Kulhari, H. Chapter 7—Dendrimer-drug conjugates: Synthesis strategies, stability and application in anticancer drug delivery. In Design of Nanostructures for Theranostics Applications; Grumezescu, A.M., Ed.; William Andrew Publishing: Norwich, NY, USA, 2018; pp. 277–303. [Google Scholar]
  109. Yiyun, C.; Na, M.; Tongwen, X.; Rongqiang, F.; Xueyuan, W.; Xiaomin, W.; Longping, W. Transdermal delivery of nonsteroidal anti-inflammatory drugs mediated by polyamidoamine (PAMAM) dendrimers. J. Pharm. Sci. 2007, 96, 595–602. [Google Scholar] [CrossRef] [PubMed]
  110. Fu, H.-L.; Cheng, S.-X.; Zhang, X.-Z.; Zhuo, R.-X. Dendrimer/DNA complexes encapsulated functional biodegradable polymer for substrate-mediated gene delivery. J. Gene Med. 2008, 10, 1334–1342. [Google Scholar] [CrossRef]
  111. Grinstaff, M.W. Dendritic macromers for hydrogel formation: Tailored materials for ophthalmic, orthopedic, and biotech applications. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 383–400. [Google Scholar] [CrossRef]
  112. Grinstaff, M. Biodendrimers: New polymeric biomaterials for tissues engineering. Chem. A Eur. J. 2002, 8, 2838–2846. [Google Scholar] [CrossRef]
  113. Boduch-Lee, K.A.; Chapman, T.; Petricca, S.E.; Marra, K.G.; Kumta, P. Design and synthesis of hydroxyapatite composites containing an mPEG−Dendritic Poly(l-lysine) star polycaprolactone. Macromolecules 2004, 37, 8959–8966. [Google Scholar] [CrossRef]
  114. Rajzer, I. Fabrication of bioactive polycaprolactone/hydroxyapatite scaffolds with final bilayer nano-/micro-fibrous structures for tissue engineering application. J. Mater. Sci. 2014, 49, 5799–5807. [Google Scholar] [CrossRef] [Green Version]
  115. Mintzer, M.A.; Grinstaff, M.W. Biomedical applications of dendrimers: A tutorial. Chem. Soc. Rev. 2011, 40, 173–190. [Google Scholar] [CrossRef]
  116. Oliveira, J.M.; Sousa, R.A.; Kotobuki, N.; Tadokoro, M.; Hirose, M.; Mano, J.F.; Reis, R.L.; Ohgushi, H. The osteogenic differentiation of rat bone marrow stromal cells cultured with dexamethasone-loaded carboxymethylchitosan/poly(amidoamine) dendrimer nanoparticles. Biomaterials 2009, 30, 804–813. [Google Scholar] [CrossRef] [Green Version]
  117. Fathi-Achachelouei, M.; Knopf-Marques, H.; Ribeiro da Silva, C.E.; Barthès, J.; Bat, E.; Tezcaner, A.; Vrana, N.E. Use of nanoparticles in tissue engineering and regenerative medicine. Front. Bioeng. Biotechnol. 2019, 7, 113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Jain, S.K. PEGylation: An approach for drug delivery. A review. Crit. Rev. Ther. Drug Carr. Syst. 2008, 25, 403–447. [Google Scholar] [CrossRef]
  119. Yoo, H.S.; Kim, T.G.; Park, T.G. Surface-functionalized electrospun nanofibers for tissue engineering and drug delivery. Adv. Drug Deliv. Rev. 2009, 61, 1033–1042. [Google Scholar] [CrossRef] [PubMed]
  120. Rao, J.P.; Geckeler, K.E. Polymer nanoparticles: Preparation techniques and size-control parameters. Prog. Polym. Sci. 2011, 36, 887–913. [Google Scholar] [CrossRef]
  121. Santo, V.E.; Ratanavaraporn, J.; Sato, K.; Gomes, M.E.; Mano, J.F.; Reis, R.L.; Tabata, Y. Cell engineering by the internalization of bioinstructive micelles for enhanced bone regeneration. Nanomedicine 2015, 10, 1707–1721. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Nicolas, J.; Mura, S.; Brambilla, D.; Mackiewicz, N.; Couvreur, P. Design, functionalization strategies and biomedical applications of targeted biodegradable/biocompatible polymer-based nanocarriers for drug delivery. Chem. Soc. Rev. 2013, 42, 1147–1235. [Google Scholar] [CrossRef]
  123. Fàbregas, A.; Miñarro, M.; García-Montoya, E.; Pérez-Lozano, P.; Carrillo, C.; Sarrate, R.; Sánchez, N.; Ticó, J.R.; Suñé-Negre, J.M. Impact of physical parameters on particle size and reaction yield when using the ionic gelation method to obtain cationic polymeric chitosan-tripolyphosphate nanoparticles. Int. J. Pharm. 2013, 446, 199–204. [Google Scholar] [CrossRef]
  124. Vaculikova, E.; Grunwaldova, V.; Kral, V.; Dohnal, J.; Jampilek, J. Preparation of candesartan and atorvastatin nanoparticles by solvent evaporation. Molecules 2012, 17, 13221–13234. [Google Scholar] [CrossRef]
  125. Chung, J.W.; Lee, K.; Neikirk, C.; Nelson, C.M.; Priestley, R.D. Photoresponsive coumarin-stabilized polymeric nanoparticles as a detectable drug carrier. Small 2012, 8, 1693–1700. [Google Scholar] [CrossRef]
  126. Langer, K.; Anhorn, M.G.; Steinhauser, I.; Dreis, S.; Celebi, D.; Schrickel, N.; Faust, S.; Vogel, V. Human serum albumin (HSA) nanoparticles: Reproducibility of preparation process and kinetics of enzymatic degradation. Int. J. Pharm. 2008, 347, 109–117. [Google Scholar] [CrossRef]
  127. Yu, X.; Trase, I.; Ren, M.; Duval, K.; Guo, X.; Chen, Z. Design of nanoparticle-based carriers for targeted drug delivery. J. Nanomater. 2016, 2016, 1087250. [Google Scholar] [CrossRef] [PubMed]
  128. Nagavarma, B.V.N.; Yadav, H.K.S.; Ayaz, A.; Vasudha, L.S.; Shivakumar, H.G. Different techniques for preparation of polymeric nanoparticles—A review. Asian J. Pharm. Clin. Res. 2012, 5, 16–23. [Google Scholar]
  129. Calzoni, E.; Cesaretti, A.; Polchi, A.; Di Michele, A.; Tancini, B.; Emiliani, C. Biocompatible polymer nanoparticles for drug delivery applications in cancer and neurodegenerative disorder therapies. J. Funct. Biomater. 2019, 10, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Joye, I.J.; McClements, D.J. Biopolymer-based nanoparticles and microparticles: Fabrication, characterization, and application. Curr. Opin. Colloid Interface Sci. 2014, 19, 417–427. [Google Scholar] [CrossRef]
  131. Ezhilarasi, P.N.; Karthik, P.; Chhanwal, N.; Anandharamakrishnan, C. Nanoencapsulation techniques for food bioactive components: A review. Food Bioprocess Technol. 2013, 6, 628–647. [Google Scholar] [CrossRef]
  132. Bennet, D.; Kim, S. Polymer nanoparticles for smart drug delivery. In Application of Nanotechnology in Drug Delivery; Sezer, A.D., Ed.; IntechOpen: Rijeka, Croatia, 2014. [Google Scholar]
  133. Makadia, H.K.; Siegel, S.J. Poly Lactic-co-glycolic acid (PLGA) as biodegradable controlled drug delivery carrier. Polymers 2011, 3, 1377–1397. [Google Scholar] [CrossRef]
  134. Shoichet, M.S. Polymer scaffolds for biomaterials applications. Macromolecules 2010, 43, 581–591. [Google Scholar] [CrossRef]
  135. Priya James, H.; John, R.; Alex, A.; Anoop, K.R. Smart polymers for the controlled delivery of drugs—A concise overview. Acta Pharm. Sin. B 2014, 4, 120–127. [Google Scholar] [CrossRef] [Green Version]
  136. Kashirina, A.; Yao, Y.; Liu, Y.; Leng, J. Biopolymers as bone substitutes: A review. Biomater. Sci. 2019, 7, 3961–3983. [Google Scholar] [CrossRef] [PubMed]
  137. Van Vlierberghe, S.; Dubruel, P.; Schacht, E. Biopolymer-based hydrogels as scaffolds for tissue engineering applications: A review. Biomacromolecules 2011, 12, 1387–1408. [Google Scholar] [CrossRef] [PubMed]
  138. Tang, Z.; He, C.; Tian, H.; Ding, J.; Hsiao, B.S.; Chu, B.; Chen, X. Polymeric nanostructured materials for biomedical applications. Prog. Polym. Sci. 2016, 60, 86–128. [Google Scholar] [CrossRef] [Green Version]
  139. Cheng, R.; Meng, F.; Deng, C.; Klok, H.-A.; Zhong, Z. Dual and multi-stimuli responsive polymeric nanoparticles for programmed site-specific drug delivery. Biomaterials 2013, 34, 3647–3657. [Google Scholar] [CrossRef] [PubMed]
  140. Wu, S.; Liu, X.; Yeung, K.W.K.; Liu, C.; Yang, X. Biomimetic porous scaffolds for bone tissue engineering. Mater. Sci. Eng. R Rep. 2014, 80, 1–36. [Google Scholar] [CrossRef]
  141. Sharma, K.; Mujawar, M.A.; Kaushik, A. State-of-art functional biomaterials for tissue engineering. Front. Mater. 2019, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  142. Tezcaner, A.; Baran, E.; Keskin, D. Nanoparticles based on plasma proteins for drug delivery applications. Curr. Pharm. Des. 2016, 22, 3445–3454. [Google Scholar] [CrossRef]
  143. Kumar, P.; Dehiya, B.S.; Sindhu, A. Comparative study of chitosan and chitosan–gelatin scaffold for tissue engineering. Int. Nano Lett. 2017, 7, 285–290. [Google Scholar] [CrossRef] [Green Version]
  144. Younes, I.; Rinaudo, M. Chitin and chitosan preparation from marine sources. Structure, properties and applications. Mar. Drugs 2015, 13, 1133–1174. [Google Scholar] [CrossRef] [Green Version]
  145. Al-Qadi, S.; Grenha, A.; Carrión-Recio, D.; Seijo, B.; Remuñán-López, C. Microencapsulated chitosan nanoparticles for pulmonary protein delivery: In vivo evaluation of insulin-loaded formulations. J. Control. Release 2012, 157, 383–390. [Google Scholar] [CrossRef]
  146. Berger, J.; Reist, M.; Mayer, J.M.; Felt, O.; Peppas, N.A.; Gurny, R. Structure and interactions in covalently and ionically crosslinked chitosan hydrogels for biomedical applications. Eur. J. Pharm. Biopharm. 2004, 57, 19–34. [Google Scholar] [CrossRef]
  147. Agnihotri, S.A.; Mallikarjuna, N.N.; Aminabhavi, T.M. Recent advances on chitosan-based micro- and nanoparticles in drug delivery. J. Control. Release 2004, 100, 5–28. [Google Scholar] [CrossRef]
  148. Kim, S.K.; Rajapakse, N. Enzymatic production and biological activities of chitosan oligosaccharides (COS): A review. Carbohydr. Polym. 2005, 62, 357–368. [Google Scholar] [CrossRef]
  149. Campos, E.V.R.; Oliveira, J.L.; Fraceto, L.F. Poly(ethylene glycol) and cyclodextrin-grafted chitosan: From methodologies to preparation and potential biotechnological applications. Front. Chem. 2017, 5, 93. [Google Scholar] [CrossRef] [Green Version]
  150. Wang, J.J.; Zeng, Z.W.; Xiao, R.Z.; Xie, T.; Zhou, G.L.; Zhan, X.R.; Wang, S.L. Recent advances of chitosan nanoparticles as drug carriers. Int. J. Nanomed. 2011, 6, 765–774. [Google Scholar]
  151. Levengood, S.L.; Zhang, M. Chitosan-based scaffolds for bone tissue engineering. J. Mater. Chem. B 2014, 2, 3161–3184. [Google Scholar] [CrossRef]
  152. Eap, S.; Keller, L.; Schiavi, J.; Huck, O.; Jacomine, L.; Fioretti, F.; Gauthier, C.; Sebastian, V.; Schwinté, P.; Benkirane-Jessel, N. A living thick nanofibrous implant bifunctionalized with active growth factor and stem cells for bone regeneration. Int. J. Nanomed. 2015, 10, 1061–1075. [Google Scholar]
  153. Shrestha, A.; Kishen, A. The effect of tissue inhibitors on the antibacterial activity of chitosan nanoparticles and photodynamic therapy. J. Endod. 2012, 38, 1275–1278. [Google Scholar] [CrossRef]
  154. Kumar, P.; Dehiya, B.S.; Sindhu, A. Synthesis and characterization of nHA-PEG and nBG-PEG scaffolds for hard tissue engineering applications. Ceram. Int. 2019, 45, 8370–8379. [Google Scholar] [CrossRef]
  155. Kumar, P.; Saini, M.; Dehiya, B.S.; Umar, A.; Sindhu, A.; Mohammed, H.; Al-Hadeethi, Y.; Guo, Z. Fabrication and in-vitro biocompatibility of freeze-dried CTS-nHA and CTS-nBG scaffolds for bone regeneration applications. Int. J. Biol. Macromol. 2020, 149, 1–10. [Google Scholar] [CrossRef]
  156. Bhowmick, A.; Banerjee, S.L.; Pramanik, N.; Jana, P.; Mitra, T.; Gnanamani, A.; Das, M.; Kundu, P.P. Organically modified clay supported chitosan/hydroxyapatite-zinc oxide nanocomposites with enhanced mechanical and biological properties for the application in bone tissue engineering. Int. J. Biol. Macromol. 2018, 106, 11–19. [Google Scholar] [CrossRef]
  157. Yilgor, P.; Tuzlakoglu, K.; Reis, R.L.; Hasirci, N.; Hasirci, V. Incorporation of a sequential BMP-2/BMP-7 delivery system into chitosan-based scaffolds for bone tissue engineering. Biomaterials 2009, 30, 3551–3559. [Google Scholar] [CrossRef] [Green Version]
  158. Mili, B.; Das, K.; Kumar, A.; Saxena, A.C.; Singh, P.; Ghosh, S.; Bag, S. Preparation of NGF encapsulated chitosan nanoparticles and its evaluation on neuronal differentiation potentiality of canine mesenchymal stem cells. J. Mater. Sci. Mater. Med. 2017, 29, 4. [Google Scholar] [CrossRef]
  159. Kumar, P. Nano-TiO(2) doped chitosan scaffold for the bone tissue engineering applications. Int. J. Biomater. 2018, 2018, 6576157. [Google Scholar] [CrossRef] [Green Version]
  160. Koosha, M.; Solouk, A.; Ghalei, S.; Sadeghi, D.; Bagheri, S.; Mirzadeh, H. Electrospun chitosan/gum tragacanth/polyvinyl alcohol hybrid nanofibrous scaffold for tissue engineering applications. Bioinspired Biomim. Nanobiomater. 2019, 9, 1–8. [Google Scholar]
  161. Zugravu, M.; Smith, R.; Reves, B.; Jennings, J.; Cooper, J.; Haggard, W.; Bumgardner, J. Physical properties and in vitro evaluation of collagen-chitosan-calcium phosphate microparticle-based scaffolds for bone tissue regeneration. J. Biomater. Appl. 2012, 28, 566–579. [Google Scholar] [CrossRef] [PubMed]
  162. Wahl, D.; Czernuszka, J. Collagen-hydroxyapatite composites for hard tissue repair. Eur. Cell Mater. 2006, 11, 43–56. [Google Scholar] [CrossRef] [PubMed]
  163. Krafts, K.P. Tissue repair: The hidden drama. Organogenesis 2010, 6, 225–233. [Google Scholar] [CrossRef] [Green Version]
  164. Dong, C.; Lv, Y. Application of collagen scaffold in tissue engineering: Recent advances and new perspectives. Polymers 2016, 8, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Lim, Y.-S.; Ok, Y.-J.; Hwang, S.-Y.; Kwak, J.-Y.; Yoon, S. Marine collagen as a promising biomaterial for biomedical applications. Mar. Drugs 2019, 17, 467. [Google Scholar] [CrossRef] [Green Version]
  166. Khan, R.; Khan, M.H. Use of collagen as a biomaterial: An update. J. Indian Soc. Periodontol. 2013, 17, 539–542. [Google Scholar] [CrossRef]
  167. D’Mello, S.; Atluri, K.; Geary, S.M.; Hong, L.; Elangovan, S.; Salem, A.K. Bone regeneration using gene-activated matrices. AAPS J. 2017, 19, 43–53. [Google Scholar] [CrossRef] [Green Version]
  168. Fujioka-Kobayashi, M.; Schaller, B.; Saulacic, N.; Pippenger, B.E.; Zhang, Y.; Miron, R.J. Absorbable collagen sponges loaded with recombinant bone morphogenetic protein 9 induces greater osteoblast differentiation when compared to bone morphogenetic protein 2. Clin. Exp. Dent. Res. 2017, 3, 32–40. [Google Scholar] [CrossRef] [PubMed]
  169. Nakagawa, T.; Tagawa, T. Ultrastructural study of direct bone formation induced by BMPs-collagen complex implanted into an ectopic site. Oral Dis. 2000, 6, 172–179. [Google Scholar] [CrossRef] [PubMed]
  170. Xie, J.; Baumann, M.J.; McCabe, L.R. Osteoblasts respond to hydroxyapatite surfaces with immediate changes in gene expression. J. Biomed. Mater. Res. Part A 2004, 71, 108–117. [Google Scholar] [CrossRef] [PubMed]
  171. Chan, E.C.; Kuo, S.-M.; Kong, A.M.; Morrison, W.A.; Dusting, G.J.; Mitchell, G.M.; Lim, S.Y.; Liu, G.-S. Three dimensional collagen scaffold promotes intrinsic vascularisation for tissue engineering applications. PLoS ONE 2016, 11, e0149799. [Google Scholar] [CrossRef] [Green Version]
  172. Vázquez, J.J.; Martín-Martínez, E.S. Collagen and elastin scaffold by electrospinning for skin tissue engineering applications. J. Mater. Res. 2019, 34, 2819–2827. [Google Scholar] [CrossRef]
  173. Marques, C.F.; Diogo, G.S.; Pina, S.; Oliveira, J.M.; Silva, T.H.; Reis, R.L. Collagen-based bioinks for hard tissue engineering applications: A comprehensive review. J. Mater. Sci. Mater. Med. 2019, 30, 32. [Google Scholar] [CrossRef]
  174. Hoque, M.E.; Hutmacher, D.W.; Feng, W.; Li, S.; Huang, M.-H.; Vert, M.; Wong, Y.S. Fabrication using a rapid prototyping system and in vitro characterization of PEG-PCL-PLA scaffolds for tissue engineering. J. Biomater. Sci. Polym. Ed. 2005, 16, 1595–1610. [Google Scholar] [CrossRef]
  175. Ba Linh, N.T.; Lee, K.H.; Lee, B.T. Functional nanofiber mat of polyvinyl alcohol/gelatin containing nanoparticles of biphasic calcium phosphate for bone regeneration in rat calvaria defects. J. Biomed. Mater. Res. Part A 2013, 101, 2412–2423. [Google Scholar] [CrossRef]
  176. Elzoghby, A.O. Gelatin-based nanoparticles as drug and gene delivery systems: Reviewing three decades of research. J. Control. Release 2013, 172, 1075–1091. [Google Scholar] [CrossRef]
  177. Santoro, M.; Tatara, A.M.; Mikos, A.G. Gelatin carriers for drug and cell delivery in tissue engineering. J. Control. Release 2014, 190, 210–218. [Google Scholar] [CrossRef] [Green Version]
  178. Won, Y.-W.; Yoon, S.-M.; Sonn, C.H.; Lee, K.-M.; Kim, Y.-H. Nano self-assembly of recombinant human gelatin conjugated with α-tocopheryl succinate for hsp90 inhibitor, 17-AAG, delivery. ACS Nano 2011, 5, 3839–3848. [Google Scholar] [CrossRef] [PubMed]
  179. Olsen, D.; Yang, C.; Bodo, M.; Chang, R.; Leigh, S.; Baez, J.; Carmichael, D.; Perälä, M.; Hämäläinen, E.-R.; Jarvinen, M.; et al. Recombinant collagen and gelatin for drug delivery. Adv. Drug Deliv. Rev. 2003, 55, 1547–1567. [Google Scholar] [CrossRef] [PubMed]
  180. Su, K.; Wang, C. Recent advances in the use of gelatin in biomedical research. Biotechnol. Lett. 2015, 37, 2139–2145. [Google Scholar] [CrossRef] [PubMed]
  181. Nitta, S.K.; Numata, K. Biopolymer-Based nanoparticles for drug/gene delivery and tissue engineering. Int. J. Mol. Sci. 2013, 14, 1629–1654. [Google Scholar] [CrossRef] [Green Version]
  182. Kuijpers, A.J.; Engbers, G.H.M.; Krijgsveld, J.; Zaat, S.A.J.; Dankert, J.; Feijen, J. Cross-linking and characterisation of gelatin matrices for biomedical applications. J. Biomater. Sci. Polym. Ed. 2000, 11, 225–243. [Google Scholar] [CrossRef]
  183. Bigi, A.; Cojazzi, G.; Panzavolta, S.; Rubini, K.; Roveri, N. Mechanical and thermal properties of gelatin films at different degrees of glutaraldehyde crosslinking. Biomaterials 2001, 22, 763–768. [Google Scholar] [CrossRef]
  184. Yue, K.; Trujillo-de Santiago, G.; Alvarez, M.M.; Tamayol, A.; Annabi, N.; Khademhosseini, A. Synthesis, properties, and biomedical applications of gelatin methacryloyl (GelMA) hydrogels. Biomaterials 2015, 73, 254–271. [Google Scholar] [CrossRef] [Green Version]
  185. Wang, H.; Boerman, O.; Sariibrahimoglu, K.; Li, Y.; Jansen, J.; Leeuwenburgh, S. Comparison of micro- vs. nanostructured colloidal gelatin gels for sustained delivery of osteogenic proteins: Bone morphogenetic protein-2 and alkaline phosphatase. Biomaterials 2012, 33, 8695–8703. [Google Scholar] [CrossRef]
  186. Perez, R.; del Valle, S.; Altankov, G.; Ginebra, M.-P. Porous hydroxyapatite and gelatin/hydroxyapatite microspheres obtained by calcium phosphate cement emulsion. J. Biomed. Mater. Res. Part B Appl. Biomater. 2011, 97, 156–166. [Google Scholar] [CrossRef]
  187. Gil, E.S.; Frankowski, D.J.; Spontak, R.J.; Hudson, S.M. Swelling behavior and morphological evolution of mixed gelatin/silk fibroin hydrogels. Biomacromolecules 2005, 6, 3079–3087. [Google Scholar] [CrossRef]
  188. Ulrich, D.; Edwards, S.; Su, K.; Tan, K.; White, J.; Ramshaw, J.; Lo, C.; Rosamilia, A.; Werkmeister, J.; Gargett, C. Human endometrial mesenchymal stem cells modulate the tissue response and mechanical behavior of polyamide mesh implants for pelvic organ prolapse repair. Tissue Eng. Part A 2013, 20, 785–798. [Google Scholar] [CrossRef] [Green Version]
  189. Han, J.; Lazarovici, P.; Pomerantz, C.; Chen, X.; Wei, Y.; Lelkes, P.I. Co-Electrospun blends of PLGA, gelatin, and elastin as potential nonthrombogenic scaffolds for vascular tissue engineering. Biomacromolecules 2011, 12, 399–408. [Google Scholar] [CrossRef] [PubMed]
  190. Ovsianikov, A.; Deiwick, A.; van Vlierberghe, S.; Dubruel, P.; Möller, L.; Dräger, G.; Chichkov, B. Laser fabrication of three-dimensional CAD scaffolds from photosensitive gelatin for applications in tissue engineering. Biomacromolecules 2011, 12, 851–858. [Google Scholar] [CrossRef] [PubMed]
  191. Tielens, S.; Declercq, H.; Gorski, T.; Lippens, E.; Schacht, E.; Cornelissen, M. Gelatin-based microcarriers as embryonic stem cell delivery system in bone tissue engineering:  An in-vitro study. Biomacromolecules 2007, 8, 825–832. [Google Scholar] [CrossRef] [PubMed]
  192. Samal, S.K.; Goranov, V.; Dash, M.; Russo, A.; Shelyakova, T.; Graziosi, P.; Lungaro, L.; Riminucci, A.; Uhlarz, M.; Bañobre-López, M.; et al. Multilayered magnetic gelatin membrane scaffolds. ACS Appl. Mater. Interfaces 2015, 7, 23098–23109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Gentile, P.; Chiono, V.; Carmagnola, I.; Hatton, P.V. An overview of poly(lactic-co-glycolic) acid (PLGA)-based biomaterials for bone tissue engineering. Int. J. Mol. Sci. 2014, 15, 3640–3659. [Google Scholar] [CrossRef] [PubMed]
  194. Meretoja, V.V.; Tirri, T.; Malin, M.; Seppälä, J.V.; Närhi, T.O. Ectopic bone formation in and soft-tissue response to P(CL/DLLA)/bioactive glass composite scaffolds. Clin. Oral Implants Res. 2014, 25, 159–164. [Google Scholar] [CrossRef]
  195. Yang, K.; Wan, J.; Zhang, S.; Zhang, Y.; Lee, S.-T.; Liu, Z. In vivo pharmacokinetics, long-term biodistribution, and toxicology of PEGylated graphene in mice. ACS Nano 2011, 5, 516–522. [Google Scholar] [CrossRef] [PubMed]
  196. Zhang, J.; Zhao, S.; Zhu, Y.; Huang, Y.; Zhu, M.; Tao, C.; Zhang, C. Three-dimensional printing of strontium-containing mesoporous bioactive glass scaffolds for bone regeneration. Acta Biomater. 2014, 10, 2269–2281. [Google Scholar] [CrossRef]
  197. Padmanabhan, J.; Kyriakides, T.R. Nanomaterials, inflammation, and tissue engineering. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2015, 7, 355–370. [Google Scholar] [CrossRef] [Green Version]
  198. Lombardo, D.; Kiselev, M.A.; Caccamo, M.T. Smart nanoparticles for drug delivery application: Development of versatile nanocarrier platforms in biotechnology and nanomedicine. J. Nanomater. 2019, 2019, 3702518. [Google Scholar] [CrossRef]
  199. Harris, L.D.; Kim, B.; Mooney, D.J. Open pore biodegradable matrices formed with gas foaming. J. Biomed. Mater. Res. 1998, 42, 396–402. [Google Scholar] [CrossRef]
  200. Mikos, A.G.; Thorsen, A.J.; Czerwonka, L.A.; Bao, Y.; Langer, R.; Winslow, D.N.; Vacanti, J.P. Preparation and characterization of poly(l-lactic acid) foams. Polymer 1994, 35, 1068–1077. [Google Scholar] [CrossRef]
  201. Zein, I.; Hutmacher, D.W.; Tan, K.C.; Teoh, S.H. Fused deposition modeling of novel scaffold architectures for tissue engineering applications. Biomaterials 2002, 23, 1169–1185. [Google Scholar] [CrossRef]
  202. Zhang, R.; Ma, P.X. Poly (A-hydroxyl acids)/hydroxyapatite porous composites for bone-tissue engineering.I. Preparation and morphology. J. Biomed. Mater. Res. 1999, 44, 446–455. [Google Scholar] [CrossRef]
  203. Perkins, B.L.; Naderi, N. Carbon nanostructures in bone tissue engineering. Open Orthop. J. 2016, 10, 877–899. [Google Scholar] [CrossRef] [Green Version]
  204. Dresselhaus, M.S.; Avouris, P. Introduction to carbon materials research. Carbon Nanotub. 2001, 9, 1–9. [Google Scholar]
  205. Han, Z.J.; Rider, A.E.; Ishaq, M.; Kumar, S.; Kondyurin, A.; Bilek, M.M.M.; Levchenko, I.; Ostrikov, K. Carbon nanostructures for hard tissue engineering. RSC Adv. 2013, 3, 11058–11072. [Google Scholar] [CrossRef]
  206. Eivazzadeh-Keihan, R.; Maleki, A.; de la Guardia, M.; Bani, M.S.; Chenab, K.K.; Pashazadeh-Panahi, P.; Baradaran, B.; Mokhtarzadeh, A.; Hamblin, M.R. Carbon based nanomaterials for tissue engineering of bone: Building new bone on small black scaffolds: A review. J. Adv. Res. 2019, 18, 185–201. [Google Scholar] [CrossRef]
  207. Geetha Bai, R.; Muthoosamy, K.; Manickam, S.; Hilal-Alnaqbi, A. Graphene-based 3D scaffolds in tissue engineering: Fabrication, applications, and future scope in liver tissue engineering. Int. J. Nanomed. 2019, 14, 5753–5783. [Google Scholar] [CrossRef] [Green Version]
  208. Kim, H.; Macosko, C.W. Processing-property relationships of polycarbonate/graphene composites. Polymer 2009, 50, 3797–3809. [Google Scholar] [CrossRef]
  209. Stankovich, S.; Dikin, D.A.; Dommett, G.H.B.; Kohlhaas, K.M.; Zimney, E.J.; Stach, E.A.; Piner, R.D.; Nguyen, S.B.T.; Ruoff, R.S. Graphene-based composite materials. Nature 2006, 442, 282–286. [Google Scholar] [CrossRef] [PubMed]
  210. Bon, S.B.; Valentini, L.; Verdejo, R.; Fierro, J.L.G.; Peponi, L.; Lopez-Manchado, M.A.; Kenny, J.M. Plasma fluorination of chemically derived graphene sheets and subsequent modification with butylamine. Chem. Mater. 2009, 21, 3433–3438. [Google Scholar] [CrossRef] [Green Version]
  211. Valentini, L.; Armentano, I.; Biagiotti, J.; Frulloni, E.; Kenny, J.M.; Santucci, S. Frequency dependent electrical transport between conjugated polymer and single-walled carbon nanotubes. Diam. Relat. Mater. 2003, 12, 1601–1609. [Google Scholar] [CrossRef]
  212. MacDonald, R.A.; Voge, C.M.; Kariolis, M.; Stegemann, J.P. Carbon nanotubes increase the electrical conductivity of fibroblast-seeded collagen hydrogels. Acta Biomater. 2008, 4, 1583–1592. [Google Scholar] [CrossRef]
  213. Armentano, I.; Dottori, M.; Fortunati, E.; Mattioli, S.; Kenny, J.M. Biodegradable polymer matrix nanocomposites for tissue engineering: A review. Polym. Degrad. Stabil. 2010, 95, 2126–2146. [Google Scholar] [CrossRef]
  214. Armentano, I.; Fortunati, E.; Gigli, M.; Luzi, F.; Trotta, R.; Bicchi, I.; Soccio, M.; Lotti, N.; Munari, A.; Martino, S.; et al. Effect of SWCNT introduction in random copolymers on material properties and fibroblast long term culture stability. Polym. Degrad. Stab. 2016, 132, 220–230. [Google Scholar] [CrossRef]
  215. Mihajlovic, M.; Mihajlovic, M.; Dankers, P.Y.W.; Masereeuw, R.; Sijbesma, R.P. Carbon nanotube reinforced supramolecular hydrogels for bioapplications. Macromol. Biosci. 2019, 19, 1800173. [Google Scholar] [CrossRef] [Green Version]
  216. Wu, K.; Tao, J.; Qi, L.; Chen, S.; Wan, W. Intracellular microtubules as nano-scaffolding template self-assembles with conductive carbon nanotubes for biomedical device. Mater. Sci. Eng. C 2020, 113, 11971. [Google Scholar] [CrossRef]
  217. Yang, Y.-L.; Ju, H.-Z.; Liu, S.-F.; Lee, T.-C.; Shih, Y.-W.; Chuang, L.-Y.; Guh, J.-Y.; Yang, Y.-Y.; Liao, T.-N.; Hung, T.-J.; et al. BMP-2 suppresses renal interstitial fibrosis by regulating epithelial–mesenchymal transition. J. Cell. Biochem. 2011, 112, 2558–2565. [Google Scholar] [CrossRef]
  218. Minati, L.; Antonini, V.; Dalla Serra, M.; Speranza, G. Multifunctional branched gold–carbon nanotube hybrid for cell imaging and drug delivery. Langmuir 2012, 28, 15900–15906. [Google Scholar] [CrossRef]
  219. Srikanth, M.; Asmatulu, R.; Cluff, K.; Yao, L. Material characterization and bioanalysis of hybrid scaffolds of carbon nanomaterial and polymer nanofibers. ACS Omega 2019, 4, 5044–5051. [Google Scholar] [CrossRef] [Green Version]
  220. Singh, R.K.; Patel, K.D.; Kim, J.-J.; Kim, T.-H.; Kim, J.-H.; Shin, U.S.; Lee, E.-J.; Knowles, J.C.; Kim, H.-W. Multifunctional hybrid nanocarrier: Magnetic CNTs ensheathed with mesoporous silica for drug delivery and imaging system. ACS Appl. Mater. Interfaces 2014, 6, 2201–2208. [Google Scholar] [CrossRef]
  221. Huang, W.; Zhang, J.; Dorn, H.C.; Geohegan, D.; Zhang, C. Assembly of single-walled carbon nanohorn supported liposome particles. Bioconjug. Chem. 2011, 22, 1012–1016. [Google Scholar] [CrossRef]
  222. Chen, Q.; Zhu, C.; Thouas, G.A. Progress and challenges in biomaterials used for bone tissue engineering: Bioactive glasses and elastomeric composites. Prog. Biomater. 2012, 1, 2. [Google Scholar] [CrossRef] [Green Version]
  223. Navarro, M.; Michiardi, A.; Castaño, O.; Planell, J.A. Biomaterials in orthopaedics. J. R. Soc. Interface 2008, 5, 1137–1158. [Google Scholar] [CrossRef] [Green Version]
  224. European Centre for Disease Prevention and Control: An Agency of the European Union. Available online: http://ecdc.europa.eu (accessed on 1 September 2020).
  225. Wu, X.; Wu, M.; Zhao, J.X. Recent development of silica nanoparticles as delivery vectors for cancer imaging and therapy. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 297–312. [Google Scholar] [CrossRef] [Green Version]
  226. Rosenholm, J.M.; Zhang, J.; Linden, M.; Sahlgren, C. Mesoporous silica nanoparticles in tissue engineering—A perspective. Nanomedicine 2016, 11, 391–402. [Google Scholar] [CrossRef]
  227. Bitar, A.; Ahmad, N.M.; Fessi, H.; Elaissari, A. Silica-based nanoparticles for biomedical applications. Drug Discov. Today 2012, 17, 1147–1154. [Google Scholar] [CrossRef]
  228. Türk, M.; Tamer, U.; Alver, E.; Çiftçi, H.; Metin, A.U.; Karahan, S. Fabrication and characterization of gold-nanoparticles/chitosan film: A scaffold for L929-fibroblasts. Artif. Cells Nanomed. Biotechnol. 2013, 41, 395–401. [Google Scholar] [CrossRef]
  229. Yeh, Y.-C.; Creran, B.; Rotello, V.M. Gold nanoparticles: Preparation, properties, and applications in bionanotechnology. Nanoscale 2012, 4, 1871–1880. [Google Scholar] [CrossRef]
  230. Zuber, A.; Purdey, M.; Schartner, E.; Forbes, C.; van der Hoek, B.; Giles, D.; Abell, A.; Monro, T.; Ebendorff-Heidepriem, H. Detection of gold nanoparticles with different sizes using absorption and fluorescence based method. Sensors Actuators B Chem. 2016, 227, 117–127. [Google Scholar] [CrossRef] [Green Version]
  231. Ito, A.; Kamihira, M. Tissue Engineering Using Magnetite Nanoparticles, 1st ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2011; Volume 104. [Google Scholar]
  232. Xiong, F.; Wang, H.; Feng, Y.; Li, Y.; Hua, X.; Pang, X.; Zhang, S.; Song, L.; Zhang, Y.; Gu, N. Cardioprotective activity of iron oxide nanoparticles. Sci. Rep. 2015, 5, 8579. [Google Scholar] [CrossRef] [Green Version]
  233. Madhumathi, K.; Sampath Kumar, T.S. Regenerative potential and anti-bacterial activity of tetracycline loaded apatitic nanocarriers for the treatment of periodontitis. Biomed. Mater. 2014, 9, 35002. [Google Scholar] [CrossRef]
  234. Choi, B.; Cui, Z.-K.; Kim, S.; Fan, J.; Wu, B.M.; Lee, M. Glutamine-chitosan modified calcium phosphate nanoparticles for efficient siRNA delivery and osteogenic differentiation. J. Mater. Chem. B 2015, 3, 6448–6455. [Google Scholar] [CrossRef] [Green Version]
  235. Ding, T.; Luo, Z.J.; Zheng, Y.; Hu, X.Y.; Ye, Z.X. Rapid repair and regeneration of damaged rabbit sciatic nerves by tissue-engineered scaffold made from nano-silver and collagen type I. Injury 2010, 41, 522–527. [Google Scholar] [CrossRef]
  236. Prabhu, S.; Poulose, E.K. Silver nanoparticles: Mechanism of antimicrobial action, synthesis, medical applications, and toxicity effects. Int. Nano Lett. 2012, 2, 32. [Google Scholar] [CrossRef] [Green Version]
  237. Saravanan, S.; Nethala, S.; Pattnaik, S.; Tripathi, A.; Moorthi, A.; Selvamurugan, N. Preparation, characterization and antimicrobial activity of a bio-composite scaffold containing chitosan/nano-hydroxyapatite/nano-silver for bone tissue engineering. Int. J. Biol. Macromol. 2011, 49, 188–193. [Google Scholar] [CrossRef]
  238. Hirota, M.; Hayakawa, T.; Yoshinari, M.; Ametani, A.; Shima, T.; Monden, Y.; Ozawa, T.; Sato, M.; Koyama, C.; Tamai, N.; et al. Hydroxyapatite coating for titanium fibre mesh scaffold enhances osteoblast activity and bone tissue formation. Int. J. Oral Maxillofac. Surg. 2012, 41, 1304–1309. [Google Scholar] [CrossRef]
  239. Holtorf, H.L.; Jansen, J.A.; Mikos, A.G. Ectopic bone formation in rat marrow stromal cell/titanium fiber mesh scaffold constructs: Effect of initial cell phenotype. Biomaterials 2005, 26, 6208–6216. [Google Scholar] [CrossRef]
  240. Fan, X.; Lin, L.; Messersmith, P.B. Surface-initiated polymerization from TiO2 nanoparticle surfaces through a biomimetic initiator: A new route toward polymer–matrix nanocomposites. Compos. Sci. Technol. 2006, 66, 1198–1204. [Google Scholar] [CrossRef]
  241. Kevin, B.; Chang, Y.; Webster, T.J. Increased chondrocyte adhesion on nanotubular anodized titanium. J. Biomed. Mater. Res. Part A 2008, 88, 561–568. [Google Scholar]
  242. Liao, D.L.; Liao, B.Q. Shape, size and photocatalytic activity control of TiO2 nanoparticles with surfactants. J. Photochem. Photobiol. A Chem. 2007, 187, 363–369. [Google Scholar] [CrossRef]
  243. Rakhmatia, Y.D.; Ayukawa, Y.; Atsuta, I.; Furuhashi, A.; Koyano, K. Fibroblast attachment onto novel titanium mesh membranes for guided bone regeneration. Odontology 2015, 103, 218–226. [Google Scholar] [CrossRef] [PubMed]
  244. Shi, Z.; Huang, X.; Cai, Y.; Tang, R.; Yang, D. Size effect of hydroxyapatite nanoparticles on proliferation and apoptosis of osteoblast-like cells. Acta Biomater. 2009, 5, 338–345. [Google Scholar] [CrossRef] [PubMed]
  245. Webster, T.J.; Ergun, C.; Doremus, R.H.; Siegel, R.W.; Bizios, R. Specific proteins mediate enhanced osteoblast adhesion on nanophase ceramics. J. Biomed. Mater. Res. 2000, 51, 475–483. [Google Scholar] [CrossRef]
  246. Lee, J.; Sieweke, J.H.; Rodriguez, N.A.; Schüpbach, P.; Lindström, H.; Susin, C.; Wikesjö, U.M.E. Evaluation of nano-technology-modified zirconia oral implants: A study in rabbits. J. Clin. Periodontol. 2009, 36, 610–617. [Google Scholar] [CrossRef] [PubMed]
  247. Opalinska, A.; Malka, I.; Dzwolak, W.; Chudoba, T.; Presz, A.; Lojkowski, W. Size-dependent density of zirconia nanoparticles. Beilstein J. Nanotechnol. 2015, 6, 27–35. [Google Scholar] [CrossRef] [PubMed]
  248. Park, Y.K.; Tadd, E.H.; Zubris, M.; Tannenbaum, R. Size-controlled synthesis of alumina nanoparticles from aluminum alkoxides. Mater. Res. Bull. 2005, 40, 1506–1512. [Google Scholar] [CrossRef]
  249. Ravindran Girija, A.; Balasubramanian, S. Theragnostic potentials of core/shell mesoporous silica nanostructures. Nanotheranostics 2019, 3, 1–40. [Google Scholar] [CrossRef]
  250. Tarn, D.; Ashley, C.E.; Xue, M.; Carnes, E.C.; Zink, J.I.; Brinker, C.J. Mesoporous silica nanoparticle nanocarriers: Biofunctionality and biocompatibility. Acc. Chem. Res. 2013, 46, 792–801. [Google Scholar] [CrossRef] [Green Version]
  251. Slowing, I.I.; Vivero-Escoto, J.L.; Wu, C.W.; Lin, V.S.Y. Mesoporous silica nanoparticles as controlled release drug delivery and gene transfection carriers. Adv. Drug Deliv. Rev. 2008, 60, 1278–1288. [Google Scholar] [CrossRef] [PubMed]
  252. Tang, Y.; Zhao, Y.; Wang, X.; Lin, T. Layer-by-layer assembly of silica nanoparticles on 3D fibrous scaffolds: Enhancement of osteoblast cell adhesion, proliferation, and differentiation. J. Biomed. Mater. Res. Part A 2014, 102, 3803–3812. [Google Scholar] [CrossRef] [PubMed]
  253. Zhou, Y.; Quan, G.; Wu, Q.; Zhang, X.; Niu, B.; Wu, B.; Huang, Y.; Pan, X.; Wu, C. Mesoporous silica nanoparticles for drug and gene delivery. Acta Pharm. Sin. B 2018, 8, 165–177. [Google Scholar] [CrossRef] [PubMed]
  254. Anitha, A.; Menon, D.; Koyakutty, M.; Mohan, C.C.; Nair, S.V.; Nair, M.B. Bioinspired composite matrix containing hydroxyapatite–silica core–shell nanorods for bone tissue engineering. ACS Appl. Mater. Interfaces 2017, 9, 26707–26718. [Google Scholar]
  255. Ambekar, R.S.; Kandasubramanian, B. Progress in the advancement of porous biopolymer scaffold: Tissue engineering application. Ind. Eng. Chem. Res. 2019, 58, 6163–6194. [Google Scholar] [CrossRef]
  256. Zhou, P.; Cheng, X.; Xia, Y.; Wang, P.; Zou, K.; Xu, S.; Du, J. Organic/inorganic composite membranes based on poly(l-lactic-co-glycolic acid) and mesoporous silica for effective bone tissue engineering. ACS Appl. Mater. Interfaces 2014, 6, 20895–20903. [Google Scholar] [CrossRef]
  257. Ding, Y.; Li, W.; Correia, A.; Yang, Y.; Zheng, K.; Liu, D.; Schubert, D.W.; Boccaccini, A.R.; Santos, H.A.; Roether, J.A. Electrospun polyhydroxybutyrate/poly(ε-Caprolactone)/sol–gel-derived silica hybrid scaffolds with drug releasing function for bone tissue engineering applications. ACS Appl. Mater. Interfaces 2018, 10, 14540–14548. [Google Scholar] [CrossRef] [Green Version]
  258. Cha, C.; Oh, J.; Kim, K.; Qiu, Y.; Joh, M.; Shin, S.R.; Wang, X.; Camci-Unal, G.; Wan, K.; Liao, R.; et al. Microfluidics-assisted fabrication of gelatin-silica core–shell microgels for injectable tissue constructs. Biomacromolecules 2014, 15, 283–290. [Google Scholar] [CrossRef]
  259. Qiu, K.; Chen, B.; Nie, W.; Zhou, X.; Feng, W.; Wang, W.; Chen, L.; Mo, X.; Wei, Y.; He, C. Electrophoretic deposition of dexamethasone-loaded mesoporous silica nanoparticles onto poly(l-Lactic Acid)/Poly(ε-Caprolactone) composite scaffold for bone tissue engineering. ACS Appl. Mater. Interfaces 2016, 8, 4137–4148. [Google Scholar] [CrossRef]
  260. Jones, J.R. Reprint of: Review of bioactive glass: From Hench to hybrids. Acta Biomater. 2015, 23, S53–S82. [Google Scholar] [CrossRef]
  261. Boccaccini, A.; Maquet, V. Bioresorbable and bioactive polymer/bioglass composites with tailored pore structure for tissue engineering applications. Compos. Sci. Technol. 2003, 63, 2417–2429. [Google Scholar] [CrossRef]
  262. Yu, H.; Peng, J.; Xu, Y.; Chang, J.; Li, H. Bioglass activated skin tissue engineering constructs for wound healing. ACS Appl. Mater. Interfaces 2016, 8, 703–715. [Google Scholar] [CrossRef] [PubMed]
  263. Fernandes, H.R.; Gaddam, A.; Rebelo, A.; Brazete, D.; Stan, G.E.; Ferreira, J.M.F. Bioactive glasses and glass-ceramics for healthcare applications in bone regeneration and tissue engineering. Materials 2018, 11, 2530. [Google Scholar] [CrossRef] [Green Version]
  264. Fiume, E.; Barberi, J.; Verné, E.; Baino, F. Bioactive glasses: From parent 45s5 composition to scaffold-assisted tissue-healing therapies. J. Funct. Biomater. 2018, 9, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Islam, M.T.; Felfel, R.M.; Abou Neel, E.A.; Grant, D.M.; Ahmed, I.; Hossain, K.M.Z. Bioactive calcium phosphate-based glasses and ceramics and their biomedical applications: A review. J. Tissue Eng. 2017, 8, 2041731417719170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. González, P.; Serra, J.; Liste, S.; Chiussi, S.; León, B.; Pérez-Amor, M. Raman spectroscopic study of bioactive silica based glasses. J. Non. Cryst. Solids 2003, 320, 92–99. [Google Scholar] [CrossRef]
  267. Yan, X.; Yu, C.; Zhou, X.; Tang, J.; Zhao, D. Highly ordered mesoporous bioactive glasses with superior in vitro bone-forming bioactivities. Angew. Chem. Int. Ed. 2004, 43, 5980–5984. [Google Scholar] [CrossRef]
  268. Kumar, P.; Dehiya, B.S.; Sindhu, A.; Kumar, V. Synthesis and characterization of nano bioglass for the application of bone tissue engineering. J. Nanosci. Technol. 2018, 4, 471–474. [Google Scholar] [CrossRef]
  269. De Oliveira, A.A.R.; de Souza, D.A.; Dias, L.L.S.; de Carvalho, S.M.; Mansur, H.S.; de Magalhães Pereira, M. Synthesis, characterization and cytocompatibility of spherical bioactive glass nanoparticles for potential hard tissue engineering applications. Biomed. Mater. 2013, 8, 025011. [Google Scholar] [CrossRef]
  270. Wang, W.; Yeung, K.W.K. Bone grafts and biomaterials substitutes for bone defect repair: A review. Bioact. Mater. 2017, 2, 224–247. [Google Scholar] [CrossRef] [PubMed]
  271. Chen, Q.; Boccaccini, A. Poly(D,L-lactic acid) coated 45S5 Bioglass-based scaffolds: Processing and characterization. J. Biomed. Mater. Res. Part A 2006, 77, 445–457. [Google Scholar] [CrossRef] [PubMed]
  272. Ravarian, R.; Zhong, X.; Barbeck, M.; Ghanaati, S.; Kirkpatrick, C.J.; Murphy, C.M.; Schindeler, A.; Chrzanowski, W.; Dehghani, F. Nanoscale chemical interaction enhances the physical properties of bioglass composites. ACS Nano 2013, 7, 8469–8483. [Google Scholar] [CrossRef] [PubMed]
  273. Zeng, Q.; Desai, M.S.; Jin, H.-E.; Lee, J.H.; Chang, J.; Lee, S.-W. Self-healing elastin–bioglass hydrogels. Biomacromolecules 2016, 17, 2619–2625. [Google Scholar] [CrossRef] [PubMed]
  274. Marelli, B.; Ghezzi, C.E.; Barralet, J.E.; Boccaccini, A.R.; Nazhat, S.N. Three-dimensional mineralization of dense nanofibrillar collagen−bioglass hybrid scaffolds. Biomacromolecules 2010, 11, 1470–1479. [Google Scholar] [CrossRef]
  275. Misra, S.K.; Nazhat, S.N.; Valappil, S.P.; Moshrefi-Torbati, M.; Wood, R.J.K.; Roy, I.; Boccaccini, A.R. fabrication and characterization of biodegradable poly(3-hydroxybutyrate) composite containing bioglass. Biomacromolecules 2007, 8, 2112–2119. [Google Scholar] [CrossRef]
  276. Fernandes, J.S.; Martins, M.; Neves, N.M.; Fernandes, M.H.V.; Reis, R.L.; Pires, R.A. Intrinsic antibacterial borosilicate glasses for bone tissue engineering applications. ACS Biomater. Sci. Eng. 2016, 2, 1143–1150. [Google Scholar] [CrossRef] [Green Version]
  277. Xu, Y.; Luong, D.; Walker, J.M.; Dean, D.; Becker, M.L. Modification of poly(propylene fumarate)–bioglass composites with peptide conjugates to enhance bioactivity. Biomacromolecules 2017, 18, 3168–3177. [Google Scholar] [CrossRef]
  278. Pepla, E.; Besharat, L.K.; Palaia, G.; Tenore, G.; Migliau, G. Nano-hydroxyapatite and its applications in preventive, restorative and regenerative dentistry: A review of literature. Ann. Stomatol. 2014, 5, 108–114. [Google Scholar] [CrossRef]
  279. Leventouri, T.; Antonakos, A.; Kyriacou, A.; Venturelli, R.; Liarokapis, E.; Perdikatsis, V. Crystal structure studies of human dental apatite as a function of age. Int. J. Biomater. 2009, 2009, 698547. [Google Scholar] [CrossRef] [Green Version]
  280. Freed, L.E.; Vunjak-Novakovic, G.; Biron, R.J.; Eagles, D.B.; Lesnoy, D.C.; Barlow, S.K.; Langer, R. Biodegradable polymer scaffolds for tissue engineering. Nat. Biotechnol. 1994, 12, 689–693. [Google Scholar] [CrossRef] [PubMed]
  281. Isikli, C.; Hasirci, V.; Hasirci, N. Development of porous chitosan–gelatin/hydroxyapatite composite scaffolds for hard tissue-engineering applications. J. Tissue Eng. Regen. Med. 2011, 6, 135–143. [Google Scholar] [CrossRef] [PubMed]
  282. Klein, C.; Driessen, A.A.; Degroot, K.; Vandenhooff, A. Biodegradation behavior of various calcium-phosphate materials in bone tissue. J. Biomed. Mater. Res. 1983, 17, 769–784. [Google Scholar] [CrossRef] [PubMed]
  283. Rezwan, K.; Chen, Q.Z.; Blaker, J.J.; Boccaccini, A.R. Biodegradable and bioactive porous polymer/inorganic composite scaffolds for bone tissue engineering. Biomaterials 2006, 27, 3413–3431. [Google Scholar] [CrossRef] [PubMed]
  284. Guzmán Vázquez, C.; Piña Barba, C.; Munguía, N. Stoichiometric hydroxyapatite obtained by precipitation and sol gel processes. Rev. Mex. Fis. 2005, 51, 284–293. [Google Scholar]
  285. Li, J.; Lu, X.L.; Zheng, Y.F. Effect of surface modified hydroxyapatite on the tensile property improvement of HA/PLA composite. Appl. Surf. Sci. 2008, 255, 494–497. [Google Scholar] [CrossRef]
  286. Dorozhkin, S.V. Calcium orthophosphate cements for biomedical application. J. Mater. Sci. 2008, 43, 3028–3057. [Google Scholar] [CrossRef]
  287. Narasaraju, T.S.B.; Phebe, D.E. Some physico-chemical aspects of hydroxylapatite. J. Mater. Sci. 1996, 31, 1–21. [Google Scholar] [CrossRef]
  288. Fathi, M.H.; Hanifi, A.; Mortazavi, V. Preparation and bioactivity evaluation of bone-like hydroxyapatite nanopowder. J. Mater. Process. Technol. 2008, 202, 536–542. [Google Scholar] [CrossRef]
  289. Sopyan, I.; Singh, R.; Hamdi, M. Synthesis of nano sized hydroxyapatite powder using sol-gel technique and its conversion to dense and porous bodies. Indian J. Chem. Sect. A 2008, 47, 1626–1631. [Google Scholar]
  290. Kattimani, V.S.; Kondaka, S.; Lingamaneni, K.P. Hydroxyapatite–-Past, present, and future in bone regeneration. Bone Tissue Regen. Insights 2016, 7, S36138. [Google Scholar] [CrossRef] [Green Version]
  291. Woodard, J.R.; Hilldore, A.J.; Lan, S.K.; Park, C.J.; Morgan, A.W.; Eurell, J.A.C.; Clark, S.G.; Wheeler, M.B.; Jamison, R.D.; Wagoner Johnson, A.J. The mechanical properties and osteoconductivity of hydroxyapatite bone scaffolds with multi-scale porosity. Biomaterials 2007, 28, 45–54. [Google Scholar] [CrossRef] [PubMed]
  292. Gao, X.; Song, J.; Ji, P.; Zhang, X.; Li, X.; Xu, X.; Wang, M.; Zhang, S.; Deng, Y.; Deng, F.; et al. Polydopamine-templated hydroxyapatite reinforced polycaprolactone composite nanofibers with enhanced cytocompatibility and osteogenesis for bone tissue engineering. ACS Appl. Mater. Interfaces 2016, 8, 3499–3515. [Google Scholar] [CrossRef] [PubMed]
  293. Jiang, L.; Li, Y.; Xiong, C.; Su, S.; Ding, H. Preparation and properties of bamboo fiber/nano-hydroxyapatite/poly(lactic-co-glycolic) composite scaffold for bone tissue engineering. ACS Appl. Mater. Interfaces 2017, 9, 4890–4897. [Google Scholar] [CrossRef] [PubMed]
  294. Salarian, M.; Xu, W.Z.; Wang, Z.; Sham, T.-K.; Charpentier, P.A. Hydroxyapatite–TiO2-based nanocomposites synthesized in supercritical co2 for bone tissue engineering: Physical and mechanical properties. ACS Appl. Mater. Interfaces 2014, 6, 16918–16931. [Google Scholar] [CrossRef]
  295. Liang, X.; Duan, P.; Gao, J.; Guo, R.; Qu, Z.; Li, X.; He, Y.; Yao, H.; Ding, J. Bilayered PLGA/PLGA-HAp composite scaffold for osteochondral tissue engineering and tissue regeneration. ACS Biomater. Sci. Eng. 2018, 4, 3506–3521. [Google Scholar] [CrossRef]
  296. Fu, S.; Wang, X.; Guo, G.; Shi, S.; Liang, H.; Luo, F.; Wei, Y.; Qian, Z. Preparation and characterization of nano-hydroxyapatite/poly(ε-caprolactone)−poly(ethylene glycol)−poly(ε-caprolactone) composite fibers for tissue engineering. J. Phys. Chem. C 2010, 114, 18372–18378. [Google Scholar] [CrossRef]
  297. Liu, M.; Liu, H.; Sun, S.; Li, X.; Zhou, Y.; Hou, Z.; Lin, J. Multifunctional Hydroxyapatite/Na(Y/Gd)F4:Yb3+,Er3+ composite fibers for drug delivery and dual modal imaging. Langmuir 2014, 30, 1176–1182. [Google Scholar] [CrossRef]
  298. Li, Z.; Liu, Z.; Yin, M.; Yang, X.; Yuan, Q.; Ren, J.; Qu, X. Aptamer-capped multifunctional mesoporous strontium hydroxyapatite nanovehicle for cancer-cell-responsive drug delivery and imaging. Biomacromolecules 2012, 13, 4257–4263. [Google Scholar] [CrossRef]
  299. Wang, L.; Hu, C.; Shao, L. The antimicrobial activity of nanoparticles: Present situation and prospects for the future. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef] [Green Version]
  300. Wu, C.; Lee, C.; Chen, J.; Kuo, S.; Fan, S.; Cheng, C.; Chang, F.; Ko, F. Microwave-assisted electroless deposition of silver nanoparticles onto multiwalled carbon nanotubes. Int. J. Electrochem. Sci. 2012, 7, 4133–4142. [Google Scholar]
  301. Ciobanu, C.S.; Iconaru, S.L.; Coustumer, P.L.; Constantin, L.V.; Predoi, D. Antibacterial activity of silver-doped hydroxyapatite nanoparticles against gram-positive and gram-negative bacteria. Nanoscale Res. Lett. 2012, 7, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Melnikov, O.V.; Gorbenko, O.Y.; Markelova, M.N.; Kaul, A.R.; Atsarkin, V.A.; Demidov, V.V.; Soto, C.; Roy, E.J.; Odintsov, B.M. Ag-doped manganite nanoparticles: New materials for temperature-controlled medical hyperthermia. J. Biomed. Mater. Res. A 2009, 91, 1048–1055. [Google Scholar] [CrossRef] [PubMed]
  303. Bapat, R.A.; Chaubal, T.V.; Joshi, C.P.; Bapat, P.R.; Choudhury, H.; Pandey, M.; Gorain, B.; Kesharwani, P. An overview of application of silver nanoparticles for biomaterials in dentistry. Mater. Sci. Eng. C 2018, 91, 881–898. [Google Scholar] [CrossRef] [PubMed]
  304. Rai, M.; Yadav, A.; Gade, A. Silver nanoparticles as a new generation of antimicrobials. Biotechnol. Adv. 2009, 27, 76–83. [Google Scholar] [CrossRef]
  305. Krishnaraj, C.; Jagan, E.G.; Rajasekar, S.; Selvakumar, P.; Kalaichelvan, P.T.; Mohan, N. Synthesis of silver nanoparticles using Acalypha indica leaf extracts and its antibacterial activity against water borne pathogens. Colloids Surfaces B Biointerfaces 2010, 76, 50–56. [Google Scholar] [CrossRef]
  306. Jung, W.K.; Koo, H.C.; Kim, K.W.; Shin, S.; Kim, S.H.; Park, Y.H. Antibacterial Activity and Mechanism of Action of the Silver Ion in Staphylococcus aureus and Escherichia coli. Appl. Environ. Microbiol. 2008, 74, 2171–2178. [Google Scholar] [CrossRef] [Green Version]
  307. Mihai, M.M.; Dima, M.B.; Dima, B.; Holban, A.M. Nanomaterials for Wound Healing and Infection Control. Materials 2019, 12, 2176. [Google Scholar] [CrossRef] [Green Version]
  308. Burdușel, A.-C.; Gherasim, O.; Grumezescu, A.M.; Mogoantă, L.; Ficai, A.; Andronescu, E. Biomedical applications of silver nanoparticles: An up-to-date overview. Nanomaterials 2018, 8, 681. [Google Scholar] [CrossRef] [Green Version]
  309. Stetciura, I.Y.; Markin, A.V.; Ponomarev, A.N.; Yakimansky, A.V.; Demina, T.S.; Grandfils, C.; Volodkin, D.V.; Gorin, D.A. New surface-enhanced raman scattering platforms: Composite calcium carbonate microspheres coated with astralen and silver nanoparticles. Langmuir 2013, 29, 4140–4147. [Google Scholar] [CrossRef]
  310. Zan, X.; Kozlov, M.; McCarthy, T.J.; Su, Z. Covalently attached, silver-doped Poly(vinyl alcohol) hydrogel films on Poly(l-lactic acid). Biomacromolecules 2010, 11, 1082–1088. [Google Scholar] [CrossRef] [PubMed]
  311. García-Astrain, C.; Chen, C.; Burón, M.; Palomares, T.; Eceiza, A.; Fruk, L.; Corcuera, M.Á.; Gabilondo, N. Biocompatible hydrogel nanocomposite with covalently embedded silver nanoparticles. Biomacromolecules 2015, 16, 1301–1310. [Google Scholar] [CrossRef] [PubMed]
  312. Cao, H.; Zhang, W.; Meng, F.; Guo, J.; Wang, D.; Qian, S.; Jiang, X.; Liu, X.; Chu, P.K. Osteogenesis Catalyzed by Titanium-Supported Silver Nanoparticles. ACS Appl. Mater. Interfaces 2017, 9, 5149–5157. [Google Scholar] [CrossRef] [PubMed]
  313. Patrascu, J.M.; Nedelcu, I.A.; Sonmez, M.; Ficai, D.; Ficai, A.; Vasile, B.S.; Ungureanu, C.; Albu, M.G.; Andor, B.; Andronescu, E.; et al. Composite scaffolds based on silver nanoparticles for biomedical applications. J. Nanomater. 2015, 2015, 587989. [Google Scholar] [CrossRef] [Green Version]
  314. Huang, X.; Qian, W.; El-Sayed, I.H.; El-Sayed, M.A. The potential use of the enhanced nonlinear properties of gold nanospheres in photothermal cancer therapy. Lasers Surg. Med. 2007, 39, 747–753. [Google Scholar] [CrossRef]
  315. Vieira, S.; Vial, S.; Maia, F.R.; Carvalho, M.R.; Reis, R.L.; Granja, P.L.; Oliveira, J.M. Gellan gum-coated gold nanorods: An intracellular nanosystem for bone tissue engineering. RSC Adv. 2015, 5, 77996–78005. [Google Scholar] [CrossRef]
  316. Yuan, H.; Khoury, C.G.; Wilson, C.M.; Grant, G.A.; Bennett, A.J.; Vo-Dinh, T. In vivo particle tracking and photothermal ablation using plasmon-resonant gold nanostars. Nanomed. Nanotechnol. Biol. Med. 2012, 8, 1355–1363. [Google Scholar] [CrossRef] [Green Version]
  317. Zhang, Q.; Uchaker, E.; Candelaria, S.L.; Cao, G. Nanomaterials for energy conversion and storage. Chem. Soc. Rev. 2013, 42, 3127–3171. [Google Scholar] [CrossRef]
  318. Li, H.; Pan, S.; Xia, P.; Chang, Y.; Fu, C.; Kong, W.; Yu, Z.; Wang, K.; Yang, X.; Qi, Z. Advances in the applications of gold nanoparticles in bone tissue engineering. J. Biol. Eng. 2020, 14, 14. [Google Scholar] [CrossRef]
  319. Daniel, M.-C.; Astruc, D. Gold Nanoparticles:  assembly, supramolecular chemistry, quantum-size-related properties, and applications toward biology, catalysis, and nanotechnology. Chem. Rev. 2004, 104, 293–346. [Google Scholar] [CrossRef]
  320. Aldewachi, H.; Chalati, T.; Woodroofe, M.N.; Bricklebank, N.; Sharrack, B.; Gardiner, P. Gold nanoparticle-based colorimetric biosensors. Nanoscale 2018, 10, 18–33. [Google Scholar] [CrossRef] [Green Version]
  321. Morita, M.; Tachikawa, T.; Seino, S.; Tanaka, K.; Majima, T. Controlled synthesis of gold nanoparticles on fluorescent nanodiamond via electron-beam-induced reduction method for dual-modal optical and electron bioimaging. ACS Appl. Nano Mater. 2018, 1, 355–363. [Google Scholar] [CrossRef]
  322. Zhang, Y.S.; Wang, Y.; Wang, L.; Wang, Y.; Cai, X.; Zhang, C.; Wang, L.V.; Xia, Y. Labeling human mesenchymal stem cells with gold nanocages for in vitro and in vivo tracking by two-photon microscopy and photoacoustic microscopy. Theranostics 2013, 3, 532–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Khlebtsov, N.; Bogatyrev, V.; Dykman, L.; Khlebtsov, B.; Staroverov, S.; Shirokov, A.; Matora, L.; Khanadeev, V.; Pylaev, T.; Tsyganova, N.; et al. Analytical and theranostic applications of gold Na-noparticles and multifunctional nanocomposites. Theranostics 2013, 3, 167–180. [Google Scholar] [CrossRef] [PubMed]
  324. Sun, T.-M.; Wang, Y.-C.; Wang, F.; Du, J.; Mao, C.; Sun, C.-Y.; Tang, R.; Liu, Y.; Zhu, J.; Zhu, Y.-H.; et al. Cancer stem cell therapy using doxorubicin conjugated to gold nanoparticles via hydrazone bonds. Biomaterials 2013, 35, 836–845. [Google Scholar] [CrossRef] [PubMed]
  325. Yu, M.; Lei, B.; Gao, C.; Yan, J.; Ma, P.X. Optimizing surface-engineered ultra-small gold nanoparticles for highly efficient miRNA delivery to enhance osteogenic differentiation of bone mesenchimal stromal cells. Nano Res. 2017, 10, 49–63. [Google Scholar] [CrossRef]
  326. Vial, S.; Reis, R.L.; Oliveira, J.M. Recent advances using gold nanoparticles as a promising multimodal tool for tissue engineering and regenerative medicine. Curr. Opin. Solid State Mater. Sci. 2017, 21, 92–112. [Google Scholar] [CrossRef] [Green Version]
  327. Khorasani, A.; Goldberg, M.; Doeven, E.; Littlefair, G. Titanium in biomedical applications—Properties and fabrication: A Review. J. Biomater. Tissue Eng. 2015, 5, 593–619. [Google Scholar] [CrossRef]
  328. Hong, Y.C.; Kim, J.H.; Bang, C.U.; Uhm, H.S. Gas-phase synthesis of nitrogen-doped TiO2 nanorods by microwave plasma torch at atmospheric pressure. Phys. Plasmas 2005, 12, 114501. [Google Scholar] [CrossRef]
  329. Dar, M.I.; Chandiran, A.K.; Gratzel, M.; Nazeeruddin, M.K.; Shivashankar, S.A. Controlled synthesis of TiO2 nanospheres using a microwave assisted approach for their application in dye-sensitized solar cells. J. Mater. Chem. A 2014, 2, 1662–1667. [Google Scholar] [CrossRef]
  330. Jeon, S.; Braun, P.V. Hydrothermal synthesis of er-doped luminescent TiO2 nanoparticles. Chem. Mater. 2003, 15, 1256–1263. [Google Scholar] [CrossRef]
  331. Ramakrishnan, V.M.; Natarajan, M.; Santhanam, A.; Asokan, V.; Velauthapillai, D. Size controlled synthesis of TiO2 nanoparticles by modified solvothermal method towards effective photocatalitic and photovoltaic applications. Mater. Res. 2018, 97, 351–360. [Google Scholar]
  332. Loryuenyong, V.; Angamnuaysiri, K.; Sukcharoenpong, J.; Suwannasri, A. Sol–gel derived mesoporous titania nanoparticles: Effects of calcination temperature and alcoholic solvent on the photocatalytic behavior. Ceram. Int. 2012, 38, 2233–2237. [Google Scholar] [CrossRef]
  333. Maheswari, D.; Venkatachalam, P. Enhanced efficiency and improved photocatalytic activity of 1:1 composite mixture of TiO2 nanoparticles and nanotubes in dye-sensitized solar cell. Bull. Mater. Sci. 2014, 37, 1489–1496. [Google Scholar] [CrossRef]
  334. Cai, Q.; Paulose, M.; Varghese, O.K.; Grimes, C.A. The effect of electrolyte composition on the fabrication of self-organized titanium oxide nanotube arrays by anodic oxidation. J. Mater. Res. 2005, 20, 230–236. [Google Scholar] [CrossRef]
  335. Tang, Y.; Tao, J.; Zhang, Y.; Wu, T.; Tao, H.; Bao, Z. Preparation and characterization of TiO2 nanotube arrays via anodization of titanium films deposited on FTO conducting glass at room temperature. Acta Physico Chimica Sin. 2008, 24, 2191–2197. [Google Scholar] [CrossRef]
  336. Ding, Z.; Hu, X.; Yue, P.; Lu, M.; Greenfield, P. Synthesis of anatase TiO2 supported on porous solids by chemical vapor deposition. Catal. Today 2001, 68, 173–182. [Google Scholar] [CrossRef]
  337. Lee, H.; Song, M.; Jurng, J.; Park, Y.-K. The synthesis and coating process of TiO2 nanoparticles using CVD process. Powder Technol. 2011, 214, 64–68. [Google Scholar] [CrossRef]
  338. Arami, H.; Mazloumi, M.; Khalifehzadeh, R.; Sadrnezhaad, S.K. Sonochemical preparation of TiO2 nanoparticles. Mater. Lett. 2007, 61, 4559–4561. [Google Scholar] [CrossRef]
  339. Guo, J.; Zhu, S.; Chen, Z.; Li, Y.; Yu, Z.; Liu, Q.; Li, J.; Feng, C.; Zhang, D. Sonochemical synthesis of TiO2 nanoparticles on graphene for use as photocatalyst. Ultrason. Sonochem. 2011, 18, 1082–1090. [Google Scholar] [CrossRef]
  340. Nasrollahzadeh, M.; Atarod, M.; Jaleh, B.; Gandomirouzbahani, M. In situ green synthesis of Ag nanoparticles on graphene oxide/TiO2 nanocomposite and their catalytic activity for the reduction of 4-nitrophenol, Congo red and methylene blue. Ceram. Int. 2016, 42, 8587–8596. [Google Scholar] [CrossRef]
  341. Sivaranjani, V.; Philominathan, P. Synthesize of Titanium dioxide nanoparticles using Moringa oleifera leaves and evaluation of wound healing activity. Wound Med. 2016, 12, 1–5. [Google Scholar] [CrossRef]
  342. Goutam, S.P.; Saxena, G.; Singh, V.; Yadav, A.K.; Bharagava, R.N.; Thapa, K.B. Green synthesis of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. Chem. Eng. J. 2018, 336, 386–396. [Google Scholar] [CrossRef]
  343. Paunesku, T.; Rajh, T.; Wiederrecht, G.; Maser, J.; Vogt, S.; Stojićević, N.; Protić, M.; Lai, B.; Oryhon, J.; Thurnauer, M.; et al. Biology of TiO2–oligonucleotide nanocomposites. Nat. Mater. 2003, 2, 343–346. [Google Scholar] [CrossRef] [PubMed]
  344. Paunesku, T.; Vogt, S.; Lai, B.; Maser, J.; Stojićević, N.; Thurn, K.T.; Osipo, C.; Liu, H.; Legnini, D.; Wang, Z.; et al. Intracellular distribution of TiO2-DNA oligonucleotide nanoconjugates directed to nucleolus and mitochondria indicates sequence specificity. Nano Lett. 2007, 7, 596–601. [Google Scholar] [CrossRef] [Green Version]
  345. Cesmeli, S.; Biray Avci, C. Application of titanium dioxide (TiO2) nanoparticles in cancer therapies. J. Drug Target. 2019, 27, 762–766. [Google Scholar] [CrossRef]
  346. Roy, P.; Berger, S.; Schmuki, P. TiO2 Nanotubes: Synthesis and applications. Angew. Chem. Int. Ed. 2011, 50, 2904–2939. [Google Scholar] [CrossRef]
  347. Losic, D.; Aw, M.; Santos, A.; Gulati, K.; Bariana, M. Titania nanotube arrays for local drug delivery: Recent advances and perspectives. Expert Opin. Drug Deliv. 2014, 12, 103–127. [Google Scholar] [CrossRef]
  348. Wang, Q.; Huang, J.-Y.; Li, H.-Q.; Chen, Z.; Zhao, A.Z.-J.; Wang, Y.; Zhang, K.-Q.; Sun, H.-T.; Al-Deyab, S.S.; Lai, Y.-K. TiO2 nanotube platforms for smart drug delivery: A review. Int. J. Nanomed. 2016, 11, 4819–4834. [Google Scholar]
  349. Wu, S.; Weng, Z.; Liu, X.; Yeung, K.W.K.; Chu, P.K. Functionalized TiO2 based nanomaterials for biomedical applications. Adv. Funct. Mater. 2014, 24, 5464–5481. [Google Scholar] [CrossRef]
  350. Tiainen, H.; Monjo, M.; Knychala, J.; Nilsen, O.; Lyngstadaas, S.P.; Ellingsen, J.E.; Haugen, H.J. The effect of fluoride surface modification of ceramic TiO2 on the surface properties and biological response of osteoblastic cells in vitro. Biomed. Mater. 2011, 6, 45006. [Google Scholar] [CrossRef] [PubMed]
  351. Tiainen, H.; Lyngstadaas, S.P.; Ellingsen, J.E.; Haugen, H.J. Ultra-porous titanium oxide scaffold with high compressive strength. J. Mater. Sci. Mater. Med. 2010, 21, 2783–2792. [Google Scholar] [CrossRef] [Green Version]
  352. Tabriz, K.R.; Katbab, A.A. Preparation of modified-TiO2/PLA nanocomposite films: Micromorphology, photo-degradability and antibacterial studies. AIP Conf. Proc. 2017, 1914, 70009. [Google Scholar]
  353. Wu, X.; Liu, X.; Wei, J.; Ma, J.; Deng, F.; Wei, S. Nano-TiO2/PEEK bioactive composite as a bone substitute material: In vitro and in vivo studies. Int. J. Nanomed. 2012, 7, 1215–1225. [Google Scholar]
  354. Eslami, H.; Azimi Lisar, H.; Jafarzadeh Kashi, T.S.; Tahriri, M.; Ansari, M.; Rafiei, T.; Bastami, F.; Shahin-Shamsabadi, A.; Mashhadi Abbas, F.; Tayebi, L. Poly(lactic-co-glycolic acid)(PLGA)/TiO2 nanotube bioactive composite as a novel scaffold for bone tissue engineering: In vitro and in vivo studies. Biologicals 2018, 53, 51–62. [Google Scholar] [CrossRef] [PubMed]
  355. Paušová, Š.; Krýsa, J.; Jirkovský, J.; Prevot, V.; Mailhot, G. Preparation of TiO2-SiO2 composite photocatalysts for environmental applications. J. Chem. Technol. Biotechnol. 2014, 89, 1129–1135. [Google Scholar] [CrossRef]
  356. Liu, S.; Tao, W.; Li, J.; Yang, Z.; Liu, F. Study on the formation process of Al2O3–TiO2 composite powders. Powder Technol. 2005, 155, 187–192. [Google Scholar] [CrossRef]
  357. Omid-Bakhtiari, M.; Nasr-Esfahani, M.; Nourmohamadi, A. TiO2-Bioactive glass nanostructure composite films produced by a sol-gel method: In vitro behavior and UV-enhanced bioactivity. J. Mater. Eng. Perform. 2014, 23, 285–293. [Google Scholar] [CrossRef]
  358. Oktar, F.N. Hydroxyapatite–TiO2 composites. Mater. Lett. 2006, 60, 2207–2210. [Google Scholar] [CrossRef]
  359. Khalid, N.R.; Bilal Tahir, M.; Majid, A.; Ahmed, E.; Ahmad, M.; Khalid, S.; Ahmed, W. TiO2-graphene-based composites: Synthesis, characterization, and application in photocatalysis of organic pollutants. In Micro and Nanomanufacturing Volume II; Jackson, M.J., Ahmed, W., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 95–122. [Google Scholar]
  360. Roguska, A.; Pisarek, M.; Andrzejczuk, M.; Dolata, M.; Lewandowska, M.; Janik-Czachor, M. Characterization of a calcium phosphate–TiO2 nanotube composite layer for biomedical applications. Mater. Sci. Eng. C 2011, 31, 906–914. [Google Scholar] [CrossRef]
  361. Rehman, F.U.; Zhao, C.; Jiang, H.; Wang, X. Biomedical applications of nano-titania in theranostics and photodynamic therapy. Biomater. Sci. 2016, 4, 40–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  362. Chen, J.; Dong, X.; Zhao, J.; Tang, G. In vivo acute toxicity of titanium dioxide nanoparticles to mice after intraperitioneal injection. J. Appl. Toxicol. 2009, 29, 330–337. [Google Scholar] [CrossRef] [PubMed]
  363. Chennell, P.; Feschet-Chassot, E.; Devers, T.; Awitor, K.O.; Descamps, S.; Sautou, V. In vitro evaluation of TiO2 nanotubes as cefuroxime carriers on orthopaedic implants for the prevention of periprosthetic joint infections. Int. J. Pharm. 2013, 455, 298–305. [Google Scholar] [CrossRef] [PubMed]
  364. Kumar, P.; Dehiya, B.S.; Sindhu, A. Bioceramics for hard tissue engineering applications: A review. Int. J. Appl. Eng. Res. 2018, 13, 2744–2752. [Google Scholar]
  365. Hentrich, R.L.; Graves, G.A.; Stein, H.G.; Bajpai, P.K. An evaluation of inert and resorbale ceramics for future clinical orthopedic applications. J. Biomed. Mater. Res. 1971, 5, 25–51. [Google Scholar] [CrossRef] [PubMed]
  366. Christel, P.; Meunier, A.; Heller, M.; Torre, J.P.; Peille, C.N. Mechanical properties and short-term in vivo evaluation of yttrium-oxide-partially-stabilized zirconia. J. Biomed. Mater. Res. 1989, 23, 45–61. [Google Scholar] [CrossRef] [PubMed]
  367. Afzal, M.A.F.; Kesarwani, P.; Reddy, K.M.; Kalmodia, S.; Basu, B.; Balani, K. Functionally graded hydroxyapatite-alumina-zirconia biocomposite: Synergy of toughness and biocompatibility. Mater. Sci. Eng. C 2012, 32, 1164–1173. [Google Scholar] [CrossRef]
  368. Cales, B.; Stefani, Y.; Lilley, E. Long-term in vivo and in vivo aging of a zirconia ceramic used in orthopaedy. J. Biomed. Mater. Res. 1994, 28, 619–624. [Google Scholar] [CrossRef] [PubMed]
  369. Hu, C.Y.; Yoon, T.-R. Recent updates for biomaterials used in total hip arthroplasty. Biomater. Res. 2018, 22, 1–12. [Google Scholar] [CrossRef]
  370. Vagkopoulou, T.; Koutayas, S.O.; Koidis, P.; Strub, J.R. Zirconia in dentistry: Part 1. Discovering the nature of an upcoming bioceramic. Eur. J. Esthet. Dent. 2009, 4, 130–151. [Google Scholar]
  371. Tosiriwatanapong, T.; Singhatanadgit, W. Zirconia-based biomaterials for hard tissue reconstruction. Bone Tissue Regen. Insights 2018, 9. [Google Scholar] [CrossRef]
  372. Aboushelib, M.N.; Shawky, R. Osteogenesis ability of CAD/CAM porous zirconia scaffolds enriched with nano-hydroxyapatite particles. Int. J. Implant Dent. 2017, 3, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  373. Mushahary, D.; Sravanthi, R.; Li, Y.; Kumar, M.J.; Harishankar, N.; Hodgson, P.D.; Wen, C.; Pande, G. Zirconium, calcium, and strontium contents in magnesium based biodegradable alloys modulate the efficiency of implant-induced osseointegration. Int. J. Nanomed. 2013, 8, 2887–2902. [Google Scholar]
  374. Ramaswamy, Y.; Wu, C.; Van Hummel, A.; Combes, V.; Grau, G.; Zreiqat, H. The responses of osteoblasts, osteoclasts and endothelial cells to zirconium modified calcium-silicate-based ceramic. Biomaterials 2008, 29, 4392–4402. [Google Scholar] [CrossRef]
  375. Kim, H.-W.; Kim, H.-E.; Knowles, J.C. Hard-tissue-engineered zirconia porous scaffolds with hydroxyapatite sol–gel and slurry coatings. J. Biomed. Mater. Res. Part B. 2004, 70, 270–277. [Google Scholar] [CrossRef] [Green Version]
  376. Chevalier, J.; Gremillard, L. Zirconia as a biomaterial. Compr. Biomater. 2011, 1, 95–108. [Google Scholar]
  377. Alizadeh, A.; Moztarzadeh, F.; Ostad, S.N.; Azami, M.; Geramizadeh, B.; Hatam, G.; Bizari, D.; Tavangar, S.M.; Vasei, M.; Ai, J. Synthesis of calcium phosphate-zirconia scaffold and human endometrial adult stem cells for bone tissue engineering. Artif. Cells Nanomed. Biotechnol. 2016, 44, 66–73. [Google Scholar] [CrossRef]
  378. Jayakumar, R.; Ramachandran, R.; Sudheesh Kumar, P.T.; Divyarani, V.V.; Srinivasan, S.; Chennazhi, K.P.; Tamura, H.; Nair, S.V. Fabrication of chitin–chitosan/nano ZrO2 composite scaffolds for tissue engineering applications. Int. J. Biol. Macromol. 2011, 49, 274–280. [Google Scholar] [CrossRef]
  379. Teimouri, A.; Ebrahimi, R.; Emadi, R.; Beni, B.H.; Chermahini, A.N. Nano-composite of silk fibroin–chitosan/Nano ZrO2 for tissue engineering applications: Fabrication and morphology. Int. J. Biol. Macromol. 2015, 76, 292–302. [Google Scholar] [CrossRef]
  380. Bermúdez-Reyes, B.; del Refugio Lara-Banda, M.; Reyes-Zarate, E.; Rojas-Martínez, A.; Camacho, A.; Moncada-Saucedo, N.; Pérez-Silos, V.; García-Ruiz, A.; Guzmán-López, A.; Peña-Martínez, V.; et al. Effect on growth and osteoblast mineralization of hydroxyapatite-zirconia (HA-ZrO2) obtained by a new low temperature system. Biomed. Mater. 2018, 13, 035001. [Google Scholar] [CrossRef]
  381. Sa, M.-W.; Nguyen, B.-N.B.; Moriarty, R.A.; Kamalitdinov, T.; Fisher, J.P.; Kim, J.Y. Fabrication and evaluation of 3D printed BCP scaffolds reinforced with ZrO2 for bone tissue applications. Biotechnol. Bioeng. 2018, 115, 989–999. [Google Scholar] [CrossRef] [PubMed]
  382. Bhowmick, A.; Pramanik, N.; Mitra, T.; Gnanamani, A.; Das, M.; Kundu, P.P. Mechanical and biological investigations of chitosan–polyvinyl alcohol based ZrO2 doped porous hybrid composites for bone tissue engineering applications. New J. Chem. 2017, 41, 7524–7530. [Google Scholar] [CrossRef]
  383. Wang, X.; Chen, D.; Cao, L.; Li, Y.; Boyd, B.J.; Caruso, R.A. Mesoporous titanium zirconium oxide nanospheres with potential for drug delivery applications. ACS Appl. Mater. Interfaces 2013, 5, 10926–10932. [Google Scholar] [CrossRef] [PubMed]
  384. Abánades Lázaro, I.; Haddad, S.; Rodrigo-Muñoz, J.M.; Marshall, R.J.; Sastre, B.; del Pozo, V.; Fairen-Jimenez, D.; Forgan, R.S. Surface-functionalization of Zr-Fumarate MOF for selective cytotoxicity and immune system compatibility in nanoscale drug delivery. ACS Appl. Mater. Interfaces 2018, 10, 31146–31157. [Google Scholar] [CrossRef]
  385. Milak, P.; Minatto, F.; De Noni, A., Jr.; Montedo, O. Wear performance of alumina-based ceramics—A review of the influence of microstructure on erosive wear. Cerâmica 2015, 61, 88–103. [Google Scholar] [CrossRef] [Green Version]
  386. Gaber, A.A.A.-A.; Ibrahim, D.M.; Abd-AImohsen, F.F.; El-Zanati, M.M. Synthesis of alumina, titania, and alumina-titania hydrophobic membranes via sol–gel polymeric route. J. Anal. Sci. Technol. 2013, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  387. Elsberg, L.; Moore, M. Total hip replacement: Metal-on-metal systems. In Clinical Performance of Skeletal Prostheses; Hench, L.L., Wilson, J., Eds.; Springer: Dordrecht, The Netherlands, 1996; pp. 57–70. [Google Scholar]
  388. Al-Sanabani, F.A.; Madfa, A.A.; Al-Qudaimi, N.H. Alumina ceramic for dental applications: A review article. Am. J. Mater. Res. 2014, 1, 26–34. [Google Scholar]
  389. Lee, K.-L.; Hsu, H.-Y.; You, M.-L.; Chang, C.-C.; Pan, M.-Y.; Shi, X.; Ueno, K.; Misawa, H.; Wei, P.-K. Highly sensitive aluminum-based biosensors using tailorable fano resonances in capped nanostructures. Sci. Rep. 2017, 7, 44104. [Google Scholar] [CrossRef]
  390. Bartzsch, H.; Glöß, D.; Böcher, B.; Frach, P.; Goedicke, K. Properties of SiO2 and Al2O3 films for electrical insulation applications deposited by reactive pulse magnetron sputtering. Surf. Coat. Technol. 2003, 174, 774–778. [Google Scholar] [CrossRef]
  391. Elisabet, X.-P.; Josep, F.-B.; Josep, P.; Marsal, F.L. Mesoporous alumina as a biomaterial for biomedical applications. Open Mater. Sci. 2015, 2, 13. [Google Scholar]
  392. Mohanty, M. Medical Applications of alumina ceramics. Trans. Indian Ceram. Soc. 1995, 54, 200–204. [Google Scholar] [CrossRef]
  393. Sansone, V.; Pagani, D.; Melato, M. The effects on bone cells of metal ions released from orthopaedic implants. A review. Clin. Cases Miner. Bone Metab. 2013, 10, 34–40. [Google Scholar] [CrossRef] [PubMed]
  394. Piconi, C.; Condo, S.G.; Kosmac, T. Chapter 11—Alumina- and zirconia-based ceramics for load-bearing applications. In Advanced Ceramics for Dentistry; Shen, J.Z., Kosmač, T., Eds.; Butterworth-Heinemann: Oxford, UK, 2014; pp. 219–253. [Google Scholar]
  395. Kandpal, B.C.; kumar, J.; Singh, H. Fabrication and characterisation of Al2O3/aluminium alloy 6061 composites fabricated by stir casting. Mater. Today Proc. 2017, 4, 2783–2792. [Google Scholar] [CrossRef]
  396. Saini, M.; Singh, Y.; Arora, P.; Arora, V.; Jain, K. Implant biomaterials: A comprehensive review. World J. Clin. Cases 2015, 3, 52–57. [Google Scholar] [CrossRef] [PubMed]
  397. Sarhadi, F.; Shafiee Afarani, M.; Mohebbi-Kalhori, D.; Shayesteh, M. Fabrication of alumina porous scaffolds with aligned oriented pores for bone tissue engineering applications. Appl. Phys. A 2016, 122, 390. [Google Scholar] [CrossRef]
  398. Levin, I.; Brandon, D. Metastable alumina polymorphs: Crystal structures and transition sequences. J. Am. Ceram. Soc. 1998, 81, 1995–2012. [Google Scholar] [CrossRef]
  399. Rahmati, M.; Mozafari, M. Biocompatibility of alumina-based biomaterials–A review. J. Cell. Physiol. 2019, 234, 3321–3335. [Google Scholar] [CrossRef]
  400. Ferraz, N.; Hong, J.; Santin, M.; Ott, M. Nanoporosity of alumina surfaces induces different patterns of activation in adhering monocytes/macrophages. Int. J. Biomater. 2010, 2010, 402715. [Google Scholar] [CrossRef]
  401. Gregorová, E.; Pabst, W.; Živcová, Z.; Sedlářová, I.; Holíková, S. Porous alumina ceramics prepared with wheat flour. J. Eur. Ceram. Soc. 2010, 30, 2871–2880. [Google Scholar] [CrossRef]
  402. Bartonickova, E.; Vojtisek, J.; Tkacz, J.; Pořízka, J.; Masilko, J.; Moncekova, M.; Parizek, L. Porous HA/Alumina composites intended for bone-tissue engineering. Mater. Technol. 2017, 51, 631–636. [Google Scholar] [CrossRef]
  403. Leary Swan, E.E.; Popat, K.C.; Desai, T.A. Peptide-immobilized nanoporous alumina membranes for enhanced osteoblast adhesion. Biomaterials 2005, 26, 1969–1976. [Google Scholar] [CrossRef] [PubMed]
  404. Ding, G.; Yang, R.; Ding, J.; Yuan, N.; Zhu, Y. Fabrication of porous anodic alumina with ultrasmall nanopores. Nanoscale Res. Lett. 2010, 5, 1257. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  405. Zaraska, L.; Jaskula, M.; Sulka, G. Porous anodic alumina layers with modulated pore diameters formed by sequential anodizing in different electrolytes. Mater. Lett. 2016, 171, 315–318. [Google Scholar] [CrossRef]
  406. Porta-I-Batalla, M.; Xifré-Pérez, E.; Eckstein, C.; Ferré-Borrull, J.; Marsal, L.F. 3D Nanoporous Anodic Alumina Structures for Sustained Drug Release. Nanomaterials 2017, 7, 227. [Google Scholar] [CrossRef]
  407. Vishal, P. Design and fabrication of ordered mesoporous alumina scaffold for drug delivery of poorly water soluble drug. Austin Ther. 2015, 2, 1–5. [Google Scholar]
  408. Kang, H.J.; Park, S.J.; Yoo, J.B.; Kim, D.J. Controlled drug release using nanoporous anodic aluminum oxide. Solid State Phenom. 2007, 121, 709–712. [Google Scholar] [CrossRef]
  409. Owens, G.J.; Singh, R.K.; Foroutan, F.; Alqaysi, M.; Han, C.-M.; Mahapatra, C.; Kim, H.-W.; Knowles, J.C. Sol–gel based materials for biomedical applications. Prog. Mater. Sci. 2016, 77, 1–79. [Google Scholar] [CrossRef] [Green Version]
  410. Kumar Panda, S.; Dhupal, D.; Kumar Nanda, B. Experimental study on Ca3(PO4)2-Al2O3 bio-ceramic composite using DPSS laser. Mater. Today Proc. 2018, 5, 24133–24140. [Google Scholar] [CrossRef]
  411. Toloue, E.B.; Karbasi, S.; Salehi, H.; Rafienia, M. Evaluation of mechanical properties and cell viability of poly (3-Hydroxybutyrate)-Chitosan/Al2O3 nanocomposite scaffold for cartilage tissue engineering. J. Med. Signals Sens. 2019, 9, 111–116. [Google Scholar]
  412. Venkatesh, N.; Hanumantharaju, H.G.; Aravind, J. A Study of bio-active coating of Al2O3, egg and sea shell powder on Ss316l and Ti-6Al-4V. Mater. Today Proc. 2018, 5, 22687–22693. [Google Scholar] [CrossRef]
  413. Mandracci, P.; Mussano, F.; Rivolo, P.; Carossa, S. Surface treatments and functional coatings for biocompatibility improvement and bacterial adhesion reduction in dental implantology. Coatings 2016, 6, 7. [Google Scholar] [CrossRef] [Green Version]
  414. Zhao, X.; Zhang, W.; Wang, Y.; Liu, Q.; Yang, J.; Zhang, L.; He, F. Fabrication of Al2O3 by anodic oxidation and hydrothermal synthesis of strong-bonding hydroxyapatite coatings on its surface. Appl. Surf. Sci. 2019, 470, 959–969. [Google Scholar] [CrossRef]
  415. Rodríguez, J.P.; Ríos, S.; González, M. Modulation of the proliferation and differentiation of human mesenchymal stem cells by copper. J. Cell. Biochem. 2002, 85, 92–100. [Google Scholar] [CrossRef] [PubMed]
  416. Hu, G. Copper stimulates proliferation of human endothelial cells under culture. J. Cell. Biochem. 1998, 69, 326–335. [Google Scholar] [CrossRef]
  417. Cioffi, N.; Torsi, L.; Ditaranto, N.; Tantillo, G.; Ghibelli, L.; Sabbatini, L.; Bleve-Zacheo, T.; D’Alessio, M.; Zambonin, P.G.; Traversa, E. Copper nanoparticle/polymer composites with antifungal and bacteriostatic properties. Chem. Mater. 2005, 17, 5255–5262. [Google Scholar] [CrossRef]
  418. Gallo, J.; Holinka, M.; Moucha, C.S. Antibacterial surface treatment for orthopaedic implants. Int. J. Mol. Sci. 2014, 15, 13849–13880. [Google Scholar] [CrossRef] [Green Version]
  419. Chaudhri, M.A.; Kemmler, W.; Harsch, I.; Watling, R.J. Plasma copper and bone mineral density in osteopenia: An indicator of bone mineral density in osteopenic females. Biol. Trace Elem. Res. 2009, 129, 94–98. [Google Scholar] [CrossRef]
  420. Das, A.; Sudhahar, V.; Chen, G.-F.; Kim, H.W.; Youn, S.-W.; Finney, L.; Vogt, S.; Yang, J.; Kweon, J.; Surenkhuu, B.; et al. Endothelial antioxidant-1: A Key mediator of copper-dependent wound healing in vivo. Sci. Rep. 2016, 6, 33783. [Google Scholar] [CrossRef] [Green Version]
  421. Nethi, S.K.; Das, S.; Patra, C.R.; Mukherjee, S. Recent advances in inorganic nanomaterials for wound-healing applications. Biomater. Sci. 2019, 7, 2652–2674. [Google Scholar] [CrossRef]
  422. Tripathi, A.; Saravanan, S.; Pattnaik, S.; Moorthi, A.; Partridge, N.C.; Selvamurugan, N. Bio-composite scaffolds containing chitosan/nano-hydroxyapatite/nano-copper–zinc for bone tissue engineering. Int. J. Biol. Macromol. 2012, 50, 294–299. [Google Scholar] [CrossRef]
  423. Azeena, S.; Subhapradha, N.; Selvamurugan, N.; Narayan, S.; Srinivasan, N.; Murugesan, R.; Chung, T.W.; Moorthi, A. Antibacterial activity of agricultural waste derived wollastonite doped with copper for bone tissue engineering. Mater. Sci. Eng. C 2017, 71, 1156–1165. [Google Scholar] [CrossRef] [PubMed]
  424. Ryan, E.J.; Ryan, A.J.; González-Vázquez, A.; Philippart, A.; Ciraldo, F.E.; Hobbs, C.; Nicolosi, V.; Boccaccini, A.R.; Kearney, C.J.; O’Brien, F.J. Collagen scaffolds functionalised with copper-eluting bioactive glass reduce infection and enhance osteogenesis and angiogenesis both in vitro and in vivo. Biomaterials 2019, 197, 405–416. [Google Scholar] [CrossRef] [PubMed]
  425. Alizadeh, S.; Seyedalipour, B.; Shafieyan, S.; Kheime, A.; Mohammadi, P.; Aghdami, N. Copper nanoparticles promote rapid wound healing in acute full thickness defect via acceleration of skin cell migration, proliferation, and neovascularization. Biochem. Biophys. Res. Commun. 2019, 517, 684–690. [Google Scholar] [CrossRef] [PubMed]
  426. Michels, H.; Moran, W.; Michel, J. Antimicrobial properties of copper alloy surfaces, with a focus on hospital-acquired infections. Int. J. Met. 2008, 2, 47–56. [Google Scholar] [CrossRef]
  427. Chen, Z.; Meng, H.; Xing, G.; Chen, C.; Zhao, Y.; Jia, G.; Wang, T.; Yuan, H.; Ye, C.; Zhao, F.; et al. Acute toxicological effects of copper nanoparticles in vivo. Toxicol. Lett. 2006, 163, 109–120. [Google Scholar] [CrossRef] [PubMed]
  428. Khan, H.A.; Sakharkar, M.K.; Nayak, A.; Kishore, U.; Khan, A. Nanoparticles for biomedical applications: An overview. In Nanobiomaterials: Nanostructured Materials for Biomedical Applications, 1st ed.; Narayan, R.B.T.-N., Ed.; Woodhead Publishing: Cambridge, UK, 2017; Chapter 14; pp. 357–384. [Google Scholar]
  429. Colombo, M.; Carregal-Romero, S.; Casula, M.F.; Gutiérrez, L.; Morales, M.P.; Böhm, I.B.; Heverhagen, J.T.; Prosperi, D.; Parak, W.J. Biological applications of magnetic nanoparticles. Chem. Soc. Rev. 2012, 41, 4306–4334. [Google Scholar] [CrossRef]
  430. Sensenig, R.; Sapir, Y.; MacDonald, C.; Cohen, S.; Polyak, B. Magnetic nanoparticle-based approaches to locally target therapy and enhance tissue regeneration in vivo. Nanomedicine 2012, 7, 1425–1442. [Google Scholar] [CrossRef] [Green Version]
  431. Khalil, M.I. Co-precipitation in aqueous solution synthesis of magnetite nanoparticles using iron(III) salts as precursors. Arab. J. Chem. 2015, 8, 279–284. [Google Scholar] [CrossRef] [Green Version]
  432. Malik, M.A.; Wani, M.Y.; Hashim, M.A. Microemulsion method: A novel route to synthesize organic and inorganic nanomaterials: 1st Nano update. Arab. J. Chem. 2012, 5, 397–417. [Google Scholar] [CrossRef] [Green Version]
  433. Daou, T.J.; Pourroy, G.; Bégin-Colin, S.; Grenèche, J.M.; Ulhaq-Bouillet, C.; Legaré, P.; Bernhardt, P.; Leuvrey, C.; Rogez, G. Hydrothermal synthesis of monodisperse magnetite nanoparticles. Chem. Mater. 2006, 18, 4399–4404. [Google Scholar] [CrossRef]
  434. Takai, Z.; Mustafa, M.; Asman, S.; Sekak, K. Preparation and characterization of magnetite (Fe3O4) nanoparticles by sol-gel method. Int. J. Nanoelectron. Mater. 2019, 12, 37–46. [Google Scholar]
  435. Hachani, R.; Lowdell, M.; Birchall, M.; Hervault, A.; Mertz, D.; Begin-Colin, S.; Thanh, N.T.K. Polyol synthesis, functionalisation, and biocompatibility studies of superparamagnetic iron oxide nanoparticles as potential MRI contrast agents. Nanoscale 2016, 8, 3278–3287. [Google Scholar] [CrossRef] [Green Version]
  436. Salazar-Alvarez, G.; Muhammed, M.; Zagorodni, A. Novel flow injection synthesis of iron oxide nanoparticles with narrow size distribution. Chem. Eng. Sci. 2006, 61, 4625–4633. [Google Scholar] [CrossRef]
  437. Hassanjani-Roshan, A.; Vaezi, M.R.; Shokuhfar, A.; Rajabali, Z. Synthesis of iron oxide nanoparticles via sonochemical method and their characterization. Particuology 2011, 9, 95–99. [Google Scholar] [CrossRef]
  438. Aivazoglou, E.; Metaxa, E.; Hristoforou, E. Microwave-assisted synthesis of iron oxide nanoparticles in biocompatible organic environment. AIP Adv. 2017, 8, 48201. [Google Scholar] [CrossRef] [Green Version]
  439. Starowicz, M.; Starowicz, P.; Zukrowski, J.; Przewoźnik, J.; Lemański, A.; Kapusta, C. Electrochemical synthesis of magnetic iron oxide nanoparticles with controlled size. J. Nanopart. Res. 2011, 13, 7167–7176. [Google Scholar] [CrossRef] [Green Version]
  440. Huang, Y.; Zhang, L.; Huan, W.; Liang, X.; Liu, X.; Yang, Y. A study on synthesis and properties of Fe3O4 nanoparticles by solvothermal method. Glas. Phys. Chem. 2010, 36, 325–331. [Google Scholar] [CrossRef]
  441. Wei, D.; Liu, Y.; Cao, L.; Fu, L.; Li, X.; Wang, Y.; Yu, G. A Magnetism-assisted chemical vapor deposition method to produce branched or iron-encapsulated carbon nanotubes. J. Am. Chem. Soc. 2007, 129, 7364–7368. [Google Scholar] [CrossRef]
  442. Majidi, S.; Sehrig, F.Z.; Farkhani, S.M.; Goloujeh, M.S.; Akbarzadeh, A. Current methods for synthesis of magnetic nanoparticles. Artif. Cells Nanomed. Biotechnol. 2016, 44, 722–734. [Google Scholar] [CrossRef]
  443. Karade, V.C.; Waifalkar, P.P.; Dongle, T.D.; Sahoo, S.C.; Kollu, P.; Patil, P.S.; Patil, P.B. Greener synthesis of magnetite nanoparticles using green tea extract and their magnetic properties. Mater. Res. Express 2017, 4, 96102. [Google Scholar] [CrossRef] [Green Version]
  444. Kayal, S.; Bandyopadhyay, D.; Mandal, T.; Ramanujan, R. The flow of magnetic nanoparticles in magnetic drug targeting. RSC Adv. 2011, 1, 238–246. [Google Scholar] [CrossRef]
  445. Safarik, I.; Safarikova, M. Magnetic techniques for the isolation and purification of proteins and peptides. Biomagn. Res. Technol. 2004, 2, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  446. David, R.; Groebner, M.; Franz, W. Magnetic cell sorting purification of differentiated embryonic stem cells stably expressing truncated human CD4 as surface marker. Stem Cells 2005, 23, 477–482. [Google Scholar] [CrossRef] [PubMed]
  447. 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] [PubMed] [Green Version]
  448. Kudr, J.; Haddad, Y.; Richtera, L.; Heger, Z.; Cernak, M.; Adam, V.; Zitka, O. Magnetic nanoparticles: From design and synthesis to real world applications. Nanomaterials 2017, 7, 243. [Google Scholar] [CrossRef]
  449. Zhu, N.; Ji, H.; Yu, P.; Niu, J.; Bajwa, U.; Akram, M.W.; Udego, I.O.; Li, H.; Niu, X. Surface modification of magnetic iron oxide nanoparticles. Nanomaterials 2018, 8, 810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  450. Akbarzadeh, A.; Samiei, M.; Davaran, S. Magnetic nanoparticles: Preparation, physical properties, and applications in biomedicine. Nanoscale Res. Lett. 2012, 7, 144. [Google Scholar] [CrossRef] [Green Version]
  451. Pan, Y.; Du, X.; Zhao, F.; Xu, B. Magnetic nanoparticles for the manipulation of proteins and cells. Chem. Soc. Rev. 2012, 41, 2912–2942. [Google Scholar] [CrossRef]
  452. Baroli, B.; Ennas, M.G.; Loffredo, F.; Isola, M.; Pinna, R.; López-Quintela, M.A. Penetration of Metallic Nanoparticles in Human Full-Thickness Skin. J. Investig. Dermatol. 2007, 127, 1701–1712. [Google Scholar] [CrossRef]
  453. Li, Y.; Ye, D.; Li, M.; Ma, M.; Gu, N. Adaptive materials based on iron oxide nanoparticles for bone regeneration. ChemPhysChem 2018, 19, 1965–1979. [Google Scholar] [CrossRef]
  454. Li, Y.; Liu, Y.-Z.; Long, T.; Yu, X.-B.; Tang, T.; Dai, K.-R.; Tian, B.; Guo, Y.-P.; Zhu, Z.-A. Mesoporous bioactive glass as a drug delivery system: Fabrication, bactericidal properties and biocompatibility. J. Mater. Sci. Mater. Med. 2013, 24, 1951–1961. [Google Scholar] [CrossRef] [PubMed]
  455. Guo, Y.-J.; Wang, Y.-Y.; Chen, T.; Wei, Y.-T.; Chu, L.-F.; Guo, Y.-P. Hollow carbonated hydroxyapatite microspheres with mesoporous structure: Hydrothermal fabrication and drug delivery property. Mater. Sci. Eng. C 2013, 33, 3166–3172. [Google Scholar] [CrossRef] [PubMed]
  456. Mahmoud, E.E.; Kamei, G.; Harada, Y.; Shimizu, R.; Kamei, N.; Adachi, N.; Misk, N.A.; Ochi, M. Cell magnetic targeting system for repair of severe chronic osteochondral defect in a rabbit model. Cell Transplant. 2016, 25, 1073–1083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  457. Gilbertson, L.M.; Zimmerman, J.B.; Plata, D.L.; Hutchison, J.E.; Anastas, P.T. Designing nanomaterials to maximize performance and minimize undesirable implications guided by the Principles of Green Chemistry. Chem. Soc. Rev. 2015, 44, 5758–5777. [Google Scholar] [CrossRef]
  458. Uskoković, V. Entering the era of nanoscience: Time to be so small. J. Biomed. Nanotechnol. 2013, 9, 1441–1470. [Google Scholar] [CrossRef]
  459. Rizvi, S.A.A.; Saleh, A.M. Applications of nanoparticle systems in drug delivery technology. Saudi Pharm. J. 2018, 26, 64–70. [Google Scholar] [CrossRef]
  460. Hasan, A.; Morshed, M.; Memic, A.; Hassan, S.; Webster, T.J.; Marei, H.E.-S. Nanoparticles in tissue engineering: Applications, challenges and prospects. Int. J. Nanomed. 2018, 13, 5637–5655. [Google Scholar] [CrossRef] [Green Version]
  461. Singh, S.K.; Kulkarni, P.P.; Dash, D. Biomedical applications of nanomaterials: An overview. Bionanotechnology 2013, 1–32. [Google Scholar] [CrossRef]
  462. Das, S.; Mitra, S.; Khurana, S.M.P.; Debnath, N. Nanomaterials for biomedical applications. Front. Life Sci. 2013, 7, 90–98. [Google Scholar] [CrossRef]
  463. Ng, C.-T.; Baeg, G.-H.; Yu, L.E.; Bay, C.-N.O. Biomedical applications of nanomaterials as therapeutics. Curr. Med. Chem. 2018, 25, 1409–1419. [Google Scholar] [CrossRef]
  464. Gentile, A.; Ruffino, F.; Grimaldi, G.M. Complex-morphology metal-based nanostructures: Fabrication, characterization, and applications. Nanomaterials 2016, 6, 110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  465. Mourdikoudis, S.; Pallares, R.M.; Thanh, N.T.K. Characterization techniques for nanoparticles: Comparison and complementarity upon studying nanoparticle properties. Nanoscale 2018, 10, 12871–12934. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  466. Jeevanandam, J.; Barhoum, A.; Chan, Y.S.; Dufresne, A.; Danquah, M.K. Review on nanoparticles and nanostructured materials: History, sources, toxicity and regulations. Beilstein J. Nanotechnol. 2018, 9, 1050–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  467. Lin, P.-C.; Lin, S.; Wang, P.C.; Sridhar, R. Techniques for physicochemical characterization of nanomaterials. Biotechnol. Adv. 2014, 32, 711–726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  468. Biener, J.; Wittstock, A.; Baumann, T.F.; Weissmüller, J.; Bäumer, M.; Hamza, A.V. Surface chemistry in nanoscale materials. Materials 2009, 2, 2404–2428. [Google Scholar] [CrossRef]
  469. Williams, D.F. On the mechanisms of biocompatibility. Biomaterials 2008, 29, 2941–2953. [Google Scholar] [CrossRef]
  470. Singh, A.; Elisseeff, J. Biomaterials for stem cell differentiation. J. Mater. Chem. 2010, 20, 8832–8847. [Google Scholar] [CrossRef]
  471. Yang, X.; Li, Y.; Liu, X.; Zhang, R.; Feng, Q. In vitro uptake of hydroxyapatite nanoparticles and their effect on osteogenic differentiation of human mesenchymal stem cells. Stem Cells Int. 2018, 2018, 2036176. [Google Scholar] [CrossRef] [Green Version]
  472. Lavenus, S.; Trichet, V.; Le Chevalier, S.; Hoornaert, A.; Louarn, G.; Layrolle, P. Cell differentiation and osseointegration influenced by nanoscale anodized titanium surfaces. Nanomedicine 2012, 7, 967–980. [Google Scholar] [CrossRef]
  473. Somaiah, C.; Kumar, A.; Mawrie, D.; Sharma, A.; Patil, S.D.; Bhattacharyya, J.; Swaminathan, R.; Jaganathan, B.G. Collagen promotes higher adhesion, survival and proliferation of mesenchymal stem cells. PLoS ONE 2015, 10, e0145068. [Google Scholar] [CrossRef] [Green Version]
  474. Rasouli, R.; Barhoum, A.; Uludag, H. A review of nanostructured surface and materials for dental implants: Surface coating, pattering and functionalization for improved performance. Biomater. Sci. 2018, 6, 1312–1338. [Google Scholar] [CrossRef]
  475. Bertazzo, S.; Zambuzzi, W.F.; Da Silva, H.A.; Ferreira, C.V.; Bertran, C.A. Bioactivation of alumina by surface modification: A possibility for improving the applicability of alumina in bone and oral repair. Clin. Oral Implants Res. 2009, 20, 288–293. [Google Scholar] [CrossRef] [PubMed]
  476. Profeta, A.C.; Huppa, C. Bioactive-glass in oral and maxillofacial surgery. Craniomaxillofac. Trauma Reconstr. 2016, 9, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  477. Crovace, M.C.; Souza, M.T.; Chinaglia, C.R.; Peitl, O.; Zanotto, E.D. Biosilicate®—A multipurpose, highly bioactive glass-ceramic. In vitro, in vivo and clinical trials. J. Non. Cryst. Solids 2016, 432, 90–110. [Google Scholar] [CrossRef]
  478. Kargozar, S.; Montazerian, M.; Fiume, E.; Baino, F. Multiple and promising applications of strontium (Sr)-containing bioactive glasses in bone tissue engineering. Front. Bioeng. Biotechnol. 2019, 7, 161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  479. Baino, F.; Vitale-Brovarone, C. Three-dimensional glass-derived scaffolds for bone tissue engineering: Current trends and forecasts for the future. J. Biomed. Mater. Res. Part A 2011, 97, 514–535. [Google Scholar] [CrossRef] [Green Version]
  480. Tautzenberger, A.; Kovtun, A.; Ignatius, A. Nanoparticles and their potential for application in bone. Int. J. Nanomed. 2012, 7, 4545–4557. [Google Scholar] [CrossRef] [Green Version]
  481. Kumar, P.; Dehiya, B.S.; Sindhu, A. Ibuprofen-loaded CTS/nHA/nBG Scaffolds for the applications of hard tissue engineering. Iran. Biomed. J. 2019, 23, 190–199. [Google Scholar] [CrossRef] [Green Version]
  482. Faraji, A.H.; Wipf, P. Nanoparticles in cellular drug delivery. Bioorganic Med. Chem. 2009, 17, 2950–2962. [Google Scholar] [CrossRef]
  483. Kong, G.; Braun, R.D.; Dewhirst, M.W. Hyperthermia enables tumor-specific nanoparticle delivery: Effect of particle size hyperthermia enables tumor-specific nanoparticle delivery: Effect of particle size 1. Cancer Res. 2000, 60, 4440–4445. [Google Scholar]
  484. Trewyn, B.G.; Slowing, I.I.; Chen, H.; Lin, V.S. Synthesis and functionalization of a mesoporous silica nanoparticle based on the sol-gel process and applications in controlled release. Acc. Chem. Res. 2007, 40, 846–853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  485. Kumari, A.; Singla, R.; Guliani, A.; Yadav, S.K. Nanoencapsulation for drug delivery. EXCLI J. 2014, 13, 265–286. [Google Scholar] [PubMed]
  486. Mahapatro, A.; Singh, D.K. Biodegradable nanoparticles are excellent vehicle for site directed in-vivo delivery of drugs and vaccines. J. Nanobiotechnol. 2011, 9, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  487. Riley, M.K.; Vermerris, W. Recent advances in nanomaterials for gene delivery—A review. Nanomaterials 2017, 7, 94. [Google Scholar] [CrossRef] [Green Version]
  488. Ylä-Herttuala, S. Endgame: Glybera finally recommended for approval as the first gene therapy drug in the European Union. Mol. Ther. 2012, 20, 1831–1832. [Google Scholar] [CrossRef] [Green Version]
  489. Loh, X.J.; Lee, T.-C.; Dou, Q.; Deen, G.R. Utilising inorganic nanocarriers for gene delivery. Biomater. Sci. 2016, 4, 70–86. [Google Scholar] [CrossRef]
  490. Zhao, Y.; Huang, L. Lipid nanoparticles for gene delivery. Adv. Genet. 2014, 88, 13–36. [Google Scholar]
  491. Martin, B.; Sainlos, M.; Aissaoui, A.; Oudrhiri, N.; Hauchecorne, M.; Vigneron, J.-P.; Lehn, J.-M.; Lehn, P. The design of cationic lipids for gene delivery. Curr. Pharm. Des. 2005, 11, 375–394. [Google Scholar] [CrossRef]
  492. Nordling-David, M.M.; Golomb, G. Gene delivery by liposomes. Isr. J. Chem. 2013, 53, 737–747. [Google Scholar] [CrossRef]
  493. Ropert, C. Liposomes as a gene delivery system. Braz. J. Med. Biol. Res. 1999, 32, 163–169. [Google Scholar] [CrossRef] [Green Version]
  494. Zylberberg, C.; Gaskill, K.; Pasley, S.; Matosevic, S. Engineering liposomal nanoparticles for targeted gene therapy. Gene Ther. 2017, 24, 441–452. [Google Scholar] [CrossRef] [PubMed]
  495. Santander-Ortega, M.J.; Lozano, M.V.; Uchegbu, I.F.; Schätzlein, A.G. 6—Dendrimers for gene therapy. In Polymers and Nanomaterials for Gene Therapy; Narain, R., Ed.; Woodhead Publishing: Cambridge, UK, 2016; pp. 113–146. [Google Scholar]
  496. Eliyahu, H.; Barenholz, Y.; Domb, A.J. Polymers for DNA delivery. Molecules 2005, 10, 34–64. [Google Scholar] [CrossRef] [PubMed]
  497. Sharma, M.R.R.; Rekha, C.P. Polymers for gene delivery: Current status and future perspectives. Recent Patents DNA Gene Seq. 2012, 6, 98–107. [Google Scholar]
  498. Zhao, H.; Ding, R.; Zhao, X.; Li, Y.; Qu, L.; Pei, H.; Yildirimer, L.; Wu, Z.; Zhang, W. Graphene-based nanomaterials for drug and/or gene delivery, bioimaging, and tissue engineering. Drug Discov. Today 2017, 22, 1302–1317. [Google Scholar] [CrossRef] [PubMed]
  499. Imani, R.; Mohabatpour, F.; Mostafavi, F. Graphene-based nano-carrier modifications for gene delivery applications. Carbon N. Y. 2018, 140, 569–591. [Google Scholar] [CrossRef]
  500. Dolatabadi, J.E.N.; Omid, Y.O.; Losic, D. Carbon nanotubes as an advanced drug and gene delivery nanosystem. Curr. Nanosci. 2011, 7, 297–314. [Google Scholar] [CrossRef]
  501. Ramos-Perez, V.; Cifuentes, A.; Coronas, N.; de Pablo, A.; Borrós, S. Modification of carbon nanotubes for gene delivery vectors. In Nanomaterial Interfaces in Biology: Methods and Protocols; Bergese, P., Hamad-Schifferli, K., Eds.; Humana Press: Totowa, NJ, USA, 2013; pp. 261–268. [Google Scholar]
  502. Keasberry, N.A.; Yapp, C.W.; Idris, A. Mesoporous silica nanoparticles as a carrier platform for intracellular delivery of nucleic acids. Biochemistry 2017, 82, 655–662. [Google Scholar] [CrossRef]
  503. Mendes, R.; Fernandes, A.R.; Baptista, P.V. Gold nanoparticle approach to the selective delivery of gene silencing in cancer—The case for combined delivery? Genes 2017, 8, 94. [Google Scholar] [CrossRef] [Green Version]
  504. Ding, Y.; Jiang, Z.; Saha, K.; Kim, C.S.; Kim, S.T.; Landis, R.F.; Rotello, V.M. Gold nanoparticles for nucleic acid delivery. Mol. Ther. 2014, 22, 1075–1083. [Google Scholar] [CrossRef] [Green Version]
  505. Majidi, S.; Zeinali Sehrig, F.; Samiei, M.; Milani, M.; Abbasi, E.; Dadashzadeh, K.; Akbarzadeh, A. Magnetic nanoparticles: Applications in gene delivery and gene therapy. Artif. Cells Nanomed. Biotechnol. 2016, 44, 1186–1193. [Google Scholar] [CrossRef]
  506. McBain, S.C.; Yiu, H.H.P.; Dobson, J. Magnetic nanoparticles for gene and drug delivery. Int. J. Nanomed. 2008, 3, 169–180. [Google Scholar]
  507. Giacca, M.; Zacchigna, S. Virus-mediated gene delivery for human gene therapy. J. Control. Release 2012, 161, 377–388. [Google Scholar] [CrossRef] [PubMed]
  508. Domvri, K.; Zarogoulidis, P.; Porpodis, K.; Koffa, M.; Lambropoulou, M.; Kakolyris, S.; Minadakis, G.; Zarogoulidis, K.; Chatzaki, E. Gene therapy in liver diseases: State-of-the-art and future perspectives. Curr. Gene Ther. 2012, 12, 463–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  509. Nienhuis, A.W.; Nathwani, A.C.; Davidoff, A.M. Gene therapy for hemophilia. Mol. Ther. 2017, 25, 1163–1167. [Google Scholar] [CrossRef]
  510. Luo, J.; Sun, M.; Kang, Q.; Peng, Y.; Jiang, W.; Luu, H.; Luo, Q.; Park, J.; Li, Y.; Haydon, R. Gene therapy for bone regeneration. Curr. Gene Ther. 2005, 5, 167–179. [Google Scholar] [CrossRef] [Green Version]
  511. Pensak, M.J.; Lieberman, J.R. Gene therapy for bone regeneration. Curr. Pharm. Des. 2013, 19, 3466–3473. [Google Scholar] [CrossRef]
  512. Shapiro, G.; Lieber, R.; Gazit, D.; Pelled, G. Recent advances and future of gene therapy for bone regeneration. Curr. Osteoporos. Rep. 2018, 16, 504–511. [Google Scholar] [CrossRef]
  513. De Jong, W.H.; Borm, P.J.A. Drug delivery and nanoparticles:applications and hazards. Int. J. Nanomed. 2008, 3, 133–149. [Google Scholar] [CrossRef] [Green Version]
  514. Navya, P.N.; Daima, H.K. Rational engineering of physicochemical properties of nanomaterials for biomedical applications with nanotoxicological perspectives. Nano Converg. 2016, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  515. Sperling, R.A.; Parak, W.J. Surface modification, functionalization and bioconjugation of colloidal inorganic nanoparticles. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2010, 368, 1333–1383. [Google Scholar] [CrossRef]
  516. Liu, C.; Zhang, N. Chapter 13—Nanoparticles in gene therapy: Principles, prospects, and challenges. In Nanoparticles in Translational Science and Medicine; Villaverde, A.B.T.-P., Ed.; Academic Press: Cambridge, MA, USA, 2011; Volume 104, pp. 509–562. [Google Scholar]
  517. Islam, M.; Park, T.-E.; Singh, B.; Maharjan, S.; Firdous, J.; Kang, S.-K.; Yun, C.-H.; Choi, Y.; Cho, C. Major degradable polycations as carriers for DNA and siRNA. J. Control. Release 2014, 193, 74–89. [Google Scholar] [CrossRef] [PubMed]
  518. Wang, Y.; Li, Z.; Weber, T.J.; Hu, D.; Lin, C.-T.; Li, J.; Lin, Y. In Situ Live Cell Sensing of multiple nucleotides exploiting DNA/RNA aptamers and graphene oxide nanosheets. Anal. Chem. 2013, 85, 6775–6782. [Google Scholar] [CrossRef] [PubMed]
  519. Keles, E.; Song, Y.; Du, D.; Dong, W.-J.; Lin, Y. Recent progress in nanomaterials for gene delivery applications. Biomater. Sci. 2016, 4, 1291–1309. [Google Scholar] [CrossRef] [PubMed]
  520. Radu, D.R.; Lai, C.-Y.; Jeftinija, K.; Rowe, E.W.; Jeftinija, S.; Lin, V.S.-Y. A Polyamidoamine dendrimer-capped mesoporous silica nanosphere-based gene transfection reagent. J. Am. Chem. Soc. 2004, 126, 13216–13217. [Google Scholar] [CrossRef] [PubMed]
  521. Palmerston Mendes, L.; Pan, J.; Torchilin, V.P. Dendrimers as nanocarriers for nucleic acid and drug delivery in cancer therapy. Molecules 2017, 22, 1401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  522. Kukowska-Latallo, J.F.; Bielinska, A.U.; Johnson, J.; Spindle, R.; Tomalia, D.A.; Baker, J.R. Efficient transfer of genetic material into mammalian cells using starburst polyamidoamine dendrimers. Proc. Natl. Acad. Sci. USA 1996, 93, 4897–4902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  523. Chen, L.; Deng, H.; Cui, H.; Fang, J.; Zuo, Z.; Deng, J.; Li, Y.; Wang, X.; Zhao, L. Inflammatory responses and inflammation-associated diseases in organs. Oncotarget 2017, 9, 7204–7218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  524. Busscher, H.J.; van der Mei, H.C.; Subbiahdoss, G.; Jutte, P.C.; van den Dungen, J.J.A.M.; Zaat, S.A.J.; Schultz, M.J.; Grainger, D.W. Biomaterial-associated infection: Locating the finish line in the race for the surface. Sci. Transl. Med. 2012, 4, 153rv10. [Google Scholar] [CrossRef] [Green Version]
  525. Buhmann, M.T.; Stiefel, P.; Maniura-Weber, K.; Ren, Q. In vitro biofilm models for device-related infections. Trends Biotechnol. 2016, 34, 945–948. [Google Scholar] [CrossRef] [Green Version]
  526. Lin, G.-L.; McGinley, J.P.; Drysdale, S.B.; Pollard, A.J. Epidemiology and immune pathogenesis of viral sepsis. Front. Immunol. 2018, 9, 2147. [Google Scholar] [CrossRef]
  527. Oliveira, W.F.; Silva, P.M.S.; Silva, R.C.S.; Silva, G.M.M.; Machado, G.; Coelho, L.C.B.B.; Correia, M.T.S. Staphylococcus aureus and Staphylococcus epidermidis infections on implants. J. Hosp. Infect. 2018, 98, 111–117. [Google Scholar] [CrossRef] [PubMed]
  528. Holzapfel, B.M.; Reichert, J.C.; Schantz, J.-T.; Gbureck, U.; Rackwitz, L.; Nöth, U.; Jakob, F.; Rudert, M.; Groll, J.; Hutmacher, D.W. How smart do biomaterials need to be? A translational science and clinical point of view. Adv. Drug Deliv. Rev. 2013, 65, 581–603. [Google Scholar] [CrossRef] [PubMed]
  529. Ng, V. Risk of Disease Transmission With Bone Allograft. Orthopedics 2012, 35, 679–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  530. Jun, I.; Han, H.-S.; Edwards, J.R.; Jeon, H. Electrospun Fibrous scaffolds for tissue engineering: Viewpoints on architecture and fabrication. Int. J. Mol. Sci. 2018, 19, 745. [Google Scholar] [CrossRef] [Green Version]
  531. Gong, T.; Xie, J.; Liao, J.; Zhang, T.; Lin, S.; Lin, Y. Nanomaterials and bone regeneration. Bone Res. 2015, 3, 1–7. [Google Scholar] [CrossRef]
  532. Gopalu, K.; Rangaraj, S.; Venkatachalam, R.; Kannan, N. Influence of ZrO2, SiO2, Al2O3 and TiO2 nanoparticles on maize seed germination under different growth conditions. IET Nanobiotechnol. 2016, 10, 171–177. [Google Scholar]
  533. Dhandayuthapani, B.; Yoshida, Y.; Maekawa, T.; Kumar, D.S. Polymeric scaffolds in tissue engineering application: A review. Int. J. Polym. Sci. 2011, 2011, 290602. [Google Scholar] [CrossRef]
  534. Mozumder, M.S.; Zhu, J.; Perinpanayagam, H. Titania-polymeric powder coatings with nano-topography support enhanced human mesenchymal cell responses. J. Biomed. Mater. Res. Part A 2012, 100, 2695–2709. [Google Scholar] [CrossRef]
  535. Trentler, T.J.; Denler, T.E.; Bertone, J.F.; Agrawal, A.; Colvin, V.L. Synthesis of TiO2 nanocrystals by nonhydrolytic solution-based reactions. J. Am. Chem. Soc. 1999, 121, 1613–1614. [Google Scholar] [CrossRef]
  536. Hussian, H.A.R.A.; Hassan, M.A.M.; Agool, I.R. Synthesis of titanium dioxide (TiO2) nanofiber and nanotube using different chemical method. Optik 2016, 127, 2996–2999. [Google Scholar] [CrossRef]
  537. Shi, H.; Magaye, R.; Castranova, V.; Zhao, J. Titanium dioxide nanoparticles: A review of current toxicological data. Part. Fibre Toxicol. 2013, 10, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  538. Ashkarran, A.A.; Fakhari, M.; Hamidinezhad, H.; Haddadi, H.; Nourani, M.R. TiO2 nanoparticles immobilized on carbon nanotubes for enhanced visible-light photo-induced activity. J. Mater. Res. Technol. 2015, 4, 126–132. [Google Scholar] [CrossRef] [Green Version]
  539. Tsuchiya, H.; Macak, J.M.; Müller, L.; Kunze, J.; Müller, F.; Greil, P.; Virtanen, S.; Schmuki, P. Hydroxyapatite growth on anodic TiO2 nanotubes. J. Biomed. Mater. Res. Part A 2006, 77, 534–541. [Google Scholar] [CrossRef] [PubMed]
  540. Inzunza, D.; Covarrubias, C.; Von Marttens, A.; Leighton, Y.; Carvajal, J.C.; Valenzuela, F.; Diaz-Dosque, M.; Méndez, N.; Martínez, C.; Pino, A.M.; et al. Synthesis of nanostructured porous silica coatings on titanium and their cell adhesive and osteogenic differentiation properties. J. Biomed. Mater. Res. Part A 2013, 102, 37–48. [Google Scholar] [CrossRef]
  541. Assad, M.; Chernyshov, A.; Leroux, M.; Rivard, C. A new porous titanium-nickel alloy: Part 1. Cytotoxicity and genotoxicity evaluation. Biomed. Mater. Eng. 2002, 12, 225–237. [Google Scholar]
  542. Gu, Y.W.; Khor, K.; Pan, D.; Cheang, P. Activity of plasma sprayed yttria stabilized zirconia reinforced hydroxyapatite/Ti–6Al–4V composite coatings in simulated body fluid. Biomaterials 2004, 25, 3177–3185. [Google Scholar] [CrossRef]
  543. Maschhoff, P.M.; Geilich, B.M.; Webster, T.J. Greater fibroblast proliferation on an ultrasonicated ZnO/PVC nanocomposite material. Int. J. Nanomed. 2014, 9, 257–263. [Google Scholar]
  544. Pei, B.; Wang, W.; Dunne, N.; Li, X. Applications of carbon nanotubes in bone tissue regeneration and engineering: Superiority, concerns, current advancements, and prospects. Nanomaterials 2019, 9, 1501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  545. Subbiah, R.; Du, P.; Van, S.Y.; Suhaeri, M.; Hwang, M.P.; Lee, K.; Kwideok, P. Fibronectin-tethered graphene oxide as an artificial matrix for osteogenesis. Biomed. Mater. 2014, 9, 65003. [Google Scholar] [CrossRef] [PubMed]
  546. Zhao, C.; Lu, X.; Zanden, Z.; Liu, J. The promising application of graphene oxide as coating materials in orthopedic implants: Preparation, characterization and cell behavior. Biomed. Mater. 2015, 10, 15019. [Google Scholar] [CrossRef] [PubMed]
  547. Al-Jumaili, A.; Alancherry, S.; Bazaka, K.; Jacob, M.V. Review on the antimicrobial properties of carbon nanostructures. Materials 2017, 10, 1066. [Google Scholar] [CrossRef] [PubMed]
  548. 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]
  549. Guazzo, R.; Gardin, C.; Bellin, G.; Sbricoli, L.; Ferroni, L.; Ludovichetti, F.S.; Piattelli, A.; Antoniac, I.; Bressan, E.; Zavan, B. Graphene-based nanomaterials for tissue engineering in the dental field. Nanomaterials 2018, 8, 349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  550. Vera-Sánchez, M.; Aznar-Cervantes, S.; Jover, E.; García-Bernal, D.; Oñate-Sánchez, R.; Hernández-Romero, D.; Moraleda, J.M.; Collado-González, M.; Rodríguez-Lozano, F.J.; Cenis, J. Silk-fibroin and graphene oxide composites promote human periodontal ligament stem cell spontaneous differentiation into osteo/cementoblast-like cells. Stem Cells Dev. 2016, 25, 1742–1754. [Google Scholar] [CrossRef] [PubMed]
  551. Ghassemi, T.; Shahroodi, A.; Ebrahimzadeh, M.H.; Mousavian, A.; Movaffagh, J.; Moradi, A. Current concepts in scaffolding for bone tissue engineering. Arch. Bone Jt. Surg. 2018, 6, 90–99. [Google Scholar]
  552. Farack, J.; Wolf-Brandstetter, C.; Glorius, S.; Nies, B.; Standke, G.; Quadbeck, P.; Worch, H.; Scharnweber, D. The effect of perfusion culture on proliferation and differentiation of human mesenchymal stem cells on biocorrodible bone replacement material. Mater. Sci. Eng. B 2011, 176, 1767–1772. [Google Scholar] [CrossRef]
  553. Zhang, P.; Wu, T.; Kong, J.-L. In situ monitoring of intracellular controlled drug release from mesoporous silica nanoparticles coated with pH-responsive charge-reversal polymer. ACS Appl. Mater. Interfaces 2014, 6, 17446–17453. [Google Scholar] [CrossRef]
  554. Kim, S.-S.; Ahn, K.-M.; Park, M.S.; Lee, J.-H.; Choi, C.Y.; Kim, B.-S. A poly(lactide-co-glycolide)/hydroxyapatite composite scaffold with enhanced osteoconductivity. J. Biomed. Mater. Res. Part A 2007, 80, 206–215. [Google Scholar] [CrossRef]
  555. Rider, P.; Kačarević, Ž.P.; Alkildani, S.; Retnasingh, S.; Barbeck, M. Bioprinting of tissue engineering scaffolds. J. Tissue Eng. 2018, 9, 2041731418802090. [Google Scholar] [CrossRef] [Green Version]
  556. Park, J.W.; Hwang, S.R.; Yoon, I.-S. Advanced growth factor delivery systems in wound management and skin regeneration. Molecules 2017, 22, 1259. [Google Scholar] [CrossRef] [Green Version]
  557. Kleinman, H.K.; Philp, D.; Hoffman, M.P. Role of the extracellular matrix in morphogenesis. Curr. Opin. Biotechnol. 2003, 14, 526–532. [Google Scholar] [CrossRef] [PubMed]
  558. Lee, E.J.; Kasper, F.K.; Mikos, A.G. Biomaterials for tissue engineering. Ann. Biomed. Eng. 2014, 42, 323–337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  559. Sato, M.; Webster, T.J.; Sato, M.; Webster, T.J. Nanobiotechnology: Implications for the future of nanotechnology in orthopedic applications Nanobiotechnology: Implications for the future of nanotechnology in orthopedic applications. Expert Rev. Med. Devices 2004, 1, 105–114. [Google Scholar] [CrossRef] [PubMed]
  560. Demais, V.; Audrain, C.; Mabilleau, G.; Chappard, D.; Baslé, M.F. Diversity of bone matrix adhesion proteins modulates osteoblast attachment and organization of actin cytoskeleton. Morphologie 2014, 98, 53–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  561. Tran, N.; Webster, T.J. Increased osteoblast functions in the presence of hydroxyapatite-coated iron oxide nanoparticles. Acta Biomater. 2011, 7, 1298–1306. [Google Scholar] [CrossRef] [PubMed]
  562. Smith, L.A.; Ma, P.X. Nano-fibrous scaffolds for tissue engineering. Colloids Surf. B Biointerfaces 2004, 39, 125–131. [Google Scholar] [CrossRef] [PubMed]
  563. Tuzlakoglu, K.; Bolgen, N.; Salgado, A.J.; Gomes, M.E.; Piskin, E.; Reis, R.L. Nano- and micro-fiber combined scaffolds: A new architecture for bone tissue engineering. J. Mater. Sci. Mater. Med. 2005, 16, 1099–1104. [Google Scholar] [CrossRef] [Green Version]
  564. Shin, M.; Yoshimoto, H.; Vacanti, J.P. In vivo bone tissue engineering using mesenchymal stem cells on a novel electrospun nanofibrous scaffold. Tissue Eng. 2004, 10, 33–41. [Google Scholar] [CrossRef]
  565. Li, W.J.; Tuli, R.; Huang, X.; Laquerriere, P.; Tuan, R.S. Multilineage differentiation of human mesenchymal stem cells in a three-dimensional nanofibrous scaffold. Biomaterials 2005, 26, 5158–5166. [Google Scholar] [CrossRef]
  566. Li, W.J.; Tuli, R.; Okafor, C.; Derfoul, A.; Danielson, K.G.; Hall, D.J.; Tuan, R.S. A three-dimensional nanofibrous scaffold for cartilage tissue engineering using human mesenchymal stem cells. Biomaterials 2005, 26, 599–609. [Google Scholar] [CrossRef]
  567. Woo, K.M.; Chen, V.J.; Ma, P.X. Nano-fibrous scaffolding architecture selectively enhances protein adsorption contributing to cell attachment. J. Biomed. Mater. Res. 2003, 67, 531–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  568. Woo, K.M.; Jun, J.H.; Chen, V.J.; Seo, J.; Baek, J.H.; Ryoo, H.M.; Kim, G.S.; Somerman, M.J.; Ma, P.X. Nano-fibrous scaffolding promotes osteoblast differentiation and biomineralization. Biomaterials 2007, 28, 335–343. [Google Scholar] [CrossRef]
  569. Bhattarai, N.; Edmondson, D.; Veiseh, O.; Matsen, F.A.; Zhang, M. Electrospun chitosan-based nanofibers and their cellular compatibility. Biomaterials 2005, 26, 6176–6184. [Google Scholar] [CrossRef] [PubMed]
  570. Zhang, L.; Wang, Z.; Xu, C.; Li, Y.; Gao, J.; Wang, W.; Liu, Y. High strength graphene oxide/polyvinyl alcohol composite hydrogels. J. Mater. Chem. 2011, 21, 10399–10406. [Google Scholar] [CrossRef]
  571. Goenka, S.; Sant, V.; Sant, S. Graphene-based nanomaterials for drug delivery and tissue engineering. J. Control. Release 2014, 173, 75–88. [Google Scholar] [CrossRef] [PubMed]
  572. Yu, M.F.; Lourie, O.; Dyer, M.J.; Moloni, K.; Kelly, T.; Ruoff, R. Strength and breaking mechanism of multiwalled carbon nanotubes under tensile load. Science 2000, 287, 637–640. [Google Scholar] [CrossRef] [Green Version]
  573. Kim, S.-S.; Sun Park, M.; Gwak, S.-J.; Choi, C.; Kim, B.-S. Accelerated bonelike apatite growth on porous polymer/ceramic composite scaffolds in vitro. Tissue Eng. 2006, 12, 2997–3006. [Google Scholar] [CrossRef]
  574. Okamoto, M.; John, B. Synthetic biopolymer nanocomposites for tissue engineering scaffolds. Prog. Polym. Sci. 2013, 38, 1487–1503. [Google Scholar] [CrossRef]
  575. Hench, L.L. Bioceramics: From concept to clinic. J. Am. Ceram. Soc. 1991, 74, 1487–1510. [Google Scholar] [CrossRef]
  576. Wei, B.; Yao, Q.; Guo, Y.; Mao, F.; Liu, S.; Xu, Y.; Wang, L. Three-dimensional polycaprolactone–hydroxyapatite scaffolds combined with bone marrow cells for cartilage tissue engineering. J. Biomater. Appl. 2015, 30, 160–170. [Google Scholar] [CrossRef]
  577. Wei, G.; Ma, P.X. Structure and properties of nano-hydroxyapatite/polymer composite scaffolds for bone tissue engineering. Biomaterials 2004, 25, 4749–4757. [Google Scholar] [CrossRef] [PubMed]
  578. Du, C.; Cui, F.-Z.; Feng, Q.; Zhu, X.D.; Groot, K. Tissue response to nano-hydroxyapatite/collagen composite implants in marrow cavity. J. Biomed. Mat. Res. 1998, 42, 540–548. [Google Scholar] [CrossRef]
  579. Li, Z.; Yubao, L.; Aiping, Y.; Xuelin, P.; Xuejiang, W.; Xiang, Z. Preparation and in vitro investigation of chitosan/nano-hydroxyapatite composite used as bone substitute materials. J. Mater. Sci. Mater. Med. 2005, 16, 213–219. [Google Scholar] [CrossRef] [PubMed]
  580. Jiang, W.G.; Sanders, A.J.; Katoh, M.; Ungefroren, H.; Gieseler, F.; Prince, M.; Thompson, S.K.; Zollo, M.; Spano, D.; Dhawan, P.; et al. Tissue invasion and metastasis: Molecular, biological and clinical perspectives. Semin. Cancer Biol. 2015, 35, S244–S275. [Google Scholar] [CrossRef]
  581. Srinivasan, M.; Rajabi, M.A.; Mousa, S. Chapter 3—Nanobiomaterials in cancer therapy. In Nanobiomaterials in Cancer Therapy; Grumezescu, A.M., Ed.; William Andrew Publishing: Norwich, NY, USA, 2016; pp. 57–89. [Google Scholar]
  582. Hernandez, R.K.; Wade, S.W.; Reich, A.; Pirolli, M.; Liede, A.; Lyman, G.H. Incidence of bone metastases in patients with solid tumors: Analysis of oncology electronic medical records in the United States. BMC Cancer 2018, 18, 44. [Google Scholar] [CrossRef] [Green Version]
  583. Adjei, I.M.; Temples, M.N.; Brown, S.B.; Sharma, B. Targeted nanomedicine to treat bone metastasis. Pharmaceutics 2018, 10, 205. [Google Scholar] [CrossRef] [Green Version]
  584. Serafini, A.N. Therapy of metastatic bone pain. J. Nucl. Med. 2001, 42, 895–906. [Google Scholar]
  585. Auffinger, B.; Morshed, R.; Tobias, A.; Cheng, Y.; Ahmed, A.U.; Lesniak, M.S. Drug-loaded nanoparticle systems and adult stem cells: A potential marriage for the treatment of malignant glioma? Oncotarget 2013, 4, 378–396. [Google Scholar] [CrossRef]
  586. Brannon-Peppas, L.; Blanchette, J.O. Nanoparticle and targeted systems for cancer therapy. Adv. Drug Deliv. Rev. 2004, 56, 1649–1659. [Google Scholar] [CrossRef]
  587. Chu, K.F.; Dupuy, D.E. Thermal ablation of tumours: Biological mechanisms and advances in therapy. Nat. Rev. Cancer 2014, 14, 199–208. [Google Scholar] [CrossRef]
  588. Nichols, J.W.; Bae, Y.H. EPR: Evidence and fallacy. J. Control. Release 2014, 190, 451–464. [Google Scholar] [CrossRef] [PubMed]
  589. Navya, P.N.; Kaphle, A.; Srinivas, S.P.; Bhargava, S.K.; Rotello, V.M.; Daima, H.K. Current trends and challenges in cancer management and therapy using designer nanomaterials. Nano Converg. 2019, 6, 23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Composition of natural bone.
Figure 1. Composition of natural bone.
Nanomaterials 10 02019 g001
Figure 2. Design of bioactive materials based on top-down and bottom-up approaches.
Figure 2. Design of bioactive materials based on top-down and bottom-up approaches.
Nanomaterials 10 02019 g002
Figure 3. Fabrication of liposome loaded scaffold.
Figure 3. Fabrication of liposome loaded scaffold.
Nanomaterials 10 02019 g003
Figure 4. Schematic depiction of the role of tissue-engineered scaffolds in gene delivery [96] (adapted with permission from Elsevier © 2009).
Figure 4. Schematic depiction of the role of tissue-engineered scaffolds in gene delivery [96] (adapted with permission from Elsevier © 2009).
Nanomaterials 10 02019 g004
Figure 5. Structure of a typical dendrimer–drug conjugate (Reused with permission from Elsevier [108].)
Figure 5. Structure of a typical dendrimer–drug conjugate (Reused with permission from Elsevier [108].)
Nanomaterials 10 02019 g005
Figure 6. Bioactive/drug molecules loaded polymeric nanoparticles.
Figure 6. Bioactive/drug molecules loaded polymeric nanoparticles.
Nanomaterials 10 02019 g006
Figure 7. Prospective applications of chitosan.
Figure 7. Prospective applications of chitosan.
Nanomaterials 10 02019 g007
Figure 8. Biomedical applications of collagen.
Figure 8. Biomedical applications of collagen.
Nanomaterials 10 02019 g008
Figure 9. Biomedical and pharmaceutical uses of gelatin.
Figure 9. Biomedical and pharmaceutical uses of gelatin.
Nanomaterials 10 02019 g009
Figure 10. (A) Fabrication of a multilayered magnetic gelatin scaffold. (B) Magnetic gelatin membranes with increasing MNPs concentration from left to right as well as a representation of the properties of the magnetic gradient. (Adapted with permission from © 2015 American Chemical Society [192]).
Figure 10. (A) Fabrication of a multilayered magnetic gelatin scaffold. (B) Magnetic gelatin membranes with increasing MNPs concentration from left to right as well as a representation of the properties of the magnetic gradient. (Adapted with permission from © 2015 American Chemical Society [192]).
Nanomaterials 10 02019 g010
Figure 11. Structure of various allotropes of carbon (adapted with permission from Royal Society of Chemistry [205]). In the figure, “BCB” stands for “benzocyclobutene”.
Figure 11. Structure of various allotropes of carbon (adapted with permission from Royal Society of Chemistry [205]). In the figure, “BCB” stands for “benzocyclobutene”.
Nanomaterials 10 02019 g011
Figure 12. Schematic of a multifunctional mesoporous silica nanoparticle showing possible core/shell design, surface modifications, and multiple types of cargos. (Adapted with permission from © 2013 American Chemical Society [250]).
Figure 12. Schematic of a multifunctional mesoporous silica nanoparticle showing possible core/shell design, surface modifications, and multiple types of cargos. (Adapted with permission from © 2013 American Chemical Society [250]).
Nanomaterials 10 02019 g012
Figure 13. Biomedical applications of hydroxyapatite and glass ceramics.
Figure 13. Biomedical applications of hydroxyapatite and glass ceramics.
Nanomaterials 10 02019 g013
Figure 14. Schematic procedure for the fabrication of a PMMA-bioglass class II hybrid (Adapted with permission from © 2013 American Chemical Society [272]).
Figure 14. Schematic procedure for the fabrication of a PMMA-bioglass class II hybrid (Adapted with permission from © 2013 American Chemical Society [272]).
Nanomaterials 10 02019 g014
Figure 15. Schematic illustration of preparation and evaluation of tHA/PCL composite nanofibers. (Adapted with permission from © 2016 American Chemical Society [292].)
Figure 15. Schematic illustration of preparation and evaluation of tHA/PCL composite nanofibers. (Adapted with permission from © 2016 American Chemical Society [292].)
Nanomaterials 10 02019 g015
Figure 16. Schematic diagram representing mechanisms of action of Ag nanoparticles for antibacterial action (adapted with permission from Elsevier © 2018 [303]). (A) AgNP diffusion and uptake into the bacterial cell. (B) Destabilization of ribosomes. (C) Enzyme interaction. (D) Interruption of electron transfer chain. (E) Reactive oxygen species (ROS). (F) DNA damage. (G) Cell death.
Figure 16. Schematic diagram representing mechanisms of action of Ag nanoparticles for antibacterial action (adapted with permission from Elsevier © 2018 [303]). (A) AgNP diffusion and uptake into the bacterial cell. (B) Destabilization of ribosomes. (C) Enzyme interaction. (D) Interruption of electron transfer chain. (E) Reactive oxygen species (ROS). (F) DNA damage. (G) Cell death.
Nanomaterials 10 02019 g016
Figure 17. Scheme representing the importance of introducing GNPs in tissue engineering and the regenerative medicine (TERM) realm [326]. (Adapted with permission from Elsevier © 2017).
Figure 17. Scheme representing the importance of introducing GNPs in tissue engineering and the regenerative medicine (TERM) realm [326]. (Adapted with permission from Elsevier © 2017).
Nanomaterials 10 02019 g017
Figure 18. Schematic representation of the many fields of applications of nanostructuredTiO2.
Figure 18. Schematic representation of the many fields of applications of nanostructuredTiO2.
Nanomaterials 10 02019 g018
Figure 19. Schematic showing the Zr-fumarate structure with preferred properties of a metal–organic framework (MOF)-based drug delivery device (adapted with permission from © 2018 American Chemical Society [384]).
Figure 19. Schematic showing the Zr-fumarate structure with preferred properties of a metal–organic framework (MOF)-based drug delivery device (adapted with permission from © 2018 American Chemical Society [384]).
Nanomaterials 10 02019 g019
Figure 20. Schematic illustration of the mUCNPs-based redox-stimuli responsive drug delivery system for tumor diagnosis and synergetic chemo-phototherapy (adapted with permission from Elsevier © 2017 [428]).
Figure 20. Schematic illustration of the mUCNPs-based redox-stimuli responsive drug delivery system for tumor diagnosis and synergetic chemo-phototherapy (adapted with permission from Elsevier © 2017 [428]).
Nanomaterials 10 02019 g020
Figure 21. Surface modified magnetic nanoparticle.
Figure 21. Surface modified magnetic nanoparticle.
Nanomaterials 10 02019 g021
Figure 22. Applications of nanobiomaterials in the biomedical field.
Figure 22. Applications of nanobiomaterials in the biomedical field.
Nanomaterials 10 02019 g022
Figure 23. Basic principle procedures for tissue engineering. (Adapted with permission from ©2019 American Chemical Society [255]).
Figure 23. Basic principle procedures for tissue engineering. (Adapted with permission from ©2019 American Chemical Society [255]).
Nanomaterials 10 02019 g023
Figure 24. Drugs and biomolecules loaded scaffold for tissue engineering.
Figure 24. Drugs and biomolecules loaded scaffold for tissue engineering.
Nanomaterials 10 02019 g024
Figure 25. Fundamental steps of gene delivery by nanocarriers (orange spheres). (Adapted with permission from ©2016 Royal Society of Chemistry [489]).
Figure 25. Fundamental steps of gene delivery by nanocarriers (orange spheres). (Adapted with permission from ©2016 Royal Society of Chemistry [489]).
Nanomaterials 10 02019 g025
Figure 26. Schematic diagram for a possible route in the use of dendrimers as gene delivery vectors.
Figure 26. Schematic diagram for a possible route in the use of dendrimers as gene delivery vectors.
Nanomaterials 10 02019 g026
Figure 27. Desired properties of an ideal scaffold.
Figure 27. Desired properties of an ideal scaffold.
Nanomaterials 10 02019 g027
Figure 28. Engineered organic and inorganic nanobiomaterials for hard tissue engineering applications.
Figure 28. Engineered organic and inorganic nanobiomaterials for hard tissue engineering applications.
Nanomaterials 10 02019 g028
Figure 29. Bone metastasis management through combination of therapies.
Figure 29. Bone metastasis management through combination of therapies.
Nanomaterials 10 02019 g029
Figure 30. Schematic representation of different types of nanomaterials and their drug-loaded conjugates employed in cancer therapy. (Adapted with permission from ©2019 Springer [589]).
Figure 30. Schematic representation of different types of nanomaterials and their drug-loaded conjugates employed in cancer therapy. (Adapted with permission from ©2019 Springer [589]).
Nanomaterials 10 02019 g030
Table 1. Types of organic nanobiomaterials with their applications.
Table 1. Types of organic nanobiomaterials with their applications.
Types of NanomaterialsSize (nm)ApplicationsReferences
Lipid<100Nanocarriers for anticancer
drug doxorubicin
Osteoblastic bone formation
Osteoporosis treatment
[42,43,44,45]
Liposome>25High encapsulation of hydrophilic
drug (drug delivery)
Growth factor delivery
Therapeutic gene delivery
Used as a template
[46,47,48]
Dendrimers<10Multidrug delivery system[46,49,50]
Chitosan20–200Nano/microparticles or fiber-based scaffolds
Drug delivery
Support chondrocyte adhesion
Implant coating
[51,52,53,54,55]
CollagenDrug Delivery
Scaffolds
[56]
Gelatin<200Bone scaffold systems formation
Drug-loaded gelatin nanoparticles (DGNPs)
Promote cell growth
[24,57,58]
Poly(lactic-co-glycolic acid) PLGA100–250Drug delivery
Scaffold system
Nanostructured Film
Enhanced cell attachment and growth
[59,60,61,62]
Carbon Nanotubes20–100Drug delivery
Biosensing
Mechanically improved scaffold fabrication
Enhanced rat brain neuron response
[63,64,65,66]
Table 2. Types of inorganic nanomaterials with their applications.
Table 2. Types of inorganic nanomaterials with their applications.
Types of NanomaterialsSize (nm)ApplicationsReferences
Nano Silica10–100Composite-based scaffold
Bio-imaging
Drug delivery
Enhanced osteogenic differentiation
[225,226,227]
Gold nanostructured materials5–50Bioinorganic hybrid nanostructures
Thin film scaffold
Bio-imaging
[228,229,230]
Magnetic nanomaterials and nanoparticles10Drug and gene delivery
Improved cell adhesion
Cell tracking
[21,231,232]
Bioactive Glasses20–500Improved scaffolds performance
Drug and gene delivery
[233,234]
Silver nanoparticles1–100Tissue repair and regeneration
Antibacterial action
[235,236,237]
Nanostructured Titanium<300Nano tubular anodized titanium
Improved mechanical properties
Enhanced chondrocyte adhesion
Support osteoblast adhesion and proliferation
Orthopedic coating
[238,239,240,241,242,243]
Hydroxyapatite20–80
~200–500
Enhanced osteoblast functioning
Increase bone apatite formation
[244,245]
Zirconia nanoparticles<100Enhanced osteointegration
Antibacterial implants formation
[246,247]
Alumina nanoparticles<80Enhanced bone cells adhesion and proliferation
Calcium phase deposition
[245,248]
Copper nanoparticles<100Antimicrobial implant fabrication[25]
Table 3. Mineral composition of hydroxyapatite, bone, and teeth.
Table 3. Mineral composition of hydroxyapatite, bone, and teeth.
TypesCaPCa/PTotal Inorganic (%)Total Organic (%)Water (%)
HA39.618.51.67100--
Dentine35.116.91.61702010
Bone34.815.21.71652510
Enamel36.517.11.63971.51.5

Share and Cite

MDPI and ACS Style

Kumar, P.; Saini, M.; Dehiya, B.S.; Sindhu, A.; Kumar, V.; Kumar, R.; Lamberti, L.; Pruncu, C.I.; Thakur, R. Comprehensive Survey on Nanobiomaterials for Bone Tissue Engineering Applications. Nanomaterials 2020, 10, 2019. https://doi.org/10.3390/nano10102019

AMA Style

Kumar P, Saini M, Dehiya BS, Sindhu A, Kumar V, Kumar R, Lamberti L, Pruncu CI, Thakur R. Comprehensive Survey on Nanobiomaterials for Bone Tissue Engineering Applications. Nanomaterials. 2020; 10(10):2019. https://doi.org/10.3390/nano10102019

Chicago/Turabian Style

Kumar, Pawan, Meenu Saini, Brijnandan S. Dehiya, Anil Sindhu, Vinod Kumar, Ravinder Kumar, Luciano Lamberti, Catalin I. Pruncu, and Rajesh Thakur. 2020. "Comprehensive Survey on Nanobiomaterials for Bone Tissue Engineering Applications" Nanomaterials 10, no. 10: 2019. https://doi.org/10.3390/nano10102019

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