Next Article in Journal
Excitonic Enhancement and Excited Excitonic States in CsPbBr3 Halide Perovskite Single Crystals
Next Article in Special Issue
Improvement of the Surface Properties of Polyether Ether Ketone via Arc Evaporation for Biomedical Applications
Previous Article in Journal
Selective Removal of Hexavalent Chromium by Novel Nitrogen and Sulfur Containing Cellulose Composite: Role of Counter Anions
Previous Article in Special Issue
A Biocompatibility Study of Plasma Nanocoatings onto Cobalt Chromium L605 Alloy for Cardiovascular Stent Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Multifunctional Bioactive Coatings for Metallic Bone Implants

1
Department of Material Science and Technology, University of Ruse “A. Kanchev”, 8 Studentska Str., 7017 Ruse, Bulgaria
2
Medical and Biological Research Lab., “Roumen Tsanev” Institute of Molecular Biology, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Materials 2023, 16(1), 183; https://doi.org/10.3390/ma16010183
Submission received: 9 December 2022 / Revised: 19 December 2022 / Accepted: 21 December 2022 / Published: 25 December 2022
(This article belongs to the Special Issue Nanostructured Materials for Biomedical Applications)

Abstract

:
To fix the bone in orthopedics, it is almost always necessary to use implants. Metals provide the needed physical and mechanical properties for load-bearing applications. Although widely used as biomedical materials for the replacement of hard tissue, metallic implants still confront challenges, among which the foremost is their low biocompatibility. Some of them also suffer from excessive wear, low corrosion resistance, infections and shielding stress. To address these issues, various coatings have been applied to enhance their in vitro and in vivo performance. When merged with the beneficial properties of various bio-ceramic or polymer coatings remarkable bioactive, osteogenic, antibacterial, or biodegradable composite implants can be created. In this review, bioactive and high-performance coatings for metallic bone implants are systematically reviewed and their biocompatibility is discussed. Updates in coating materials and formulations for metallic implants, as well as their production routes, have been provided. The ways of improving the bioactive coating performance by incorporating bioactive moieties such as growth factors, osteogenic factors, immunomodulatory factors, antibiotics, or other drugs that are locally released in a controlled manner have also been addressed.

1. Introduction

With the global aging and obesity of the population, the necessity of applying orthopedic implants has increased rapidly. For improving the quality of a patient’s life, durability of a minimum of 15–20 years for older persons and more than 20 years for younger people is expected for a bioimplant [1]. Orthopedic implants are synthetic tools designed to provide biological support to damaged tissues or organs in the living organism and restore physiological functions. Metallic materials were first used as first-generation bone implants due to their physical and mechanical properties and inert nature. Still, among all manufactured implants, about 70 to 80% are produced from bio-metallics. However, the failure of metallic implants mostly originates at the tissue–implant interface due to poor bonding leading to the formation of a non-adherent layer of movement at this interface [2]. A study of almost 338,000 procedures for knee revisions demonstrated that aseptic loosening accounted for 20.4%, infectious factors (septic loosening) for approximately 20.3% of implant failure [3]. The aseptic loosening of the implants is a slow process that develops over the years. It can be related to the release of allergens by metals, mechanical incompatibility and occurrence of stress shielding effect on bone, poor wear and corrosion resistance leading to the formation of wear debris and release of toxic corrosive ions which can trigger inflammatory reactions [4]. Early and late (more than 24 months) implant infections are caused by different pathogens (such as Staphylococcus aureus) that rapidly proliferate and secrete extracellular polymers to form biofilm. In the presence of tiny gaps within the bone and prosthesis interface, particles and/or pathogens can accumulate and hinder the direct contact between the graft and bone [5]. All these problems are related mostly to the surface of the metallic implant. To minimize revision surgery, numerous approaches including oral intake of antibiotics, selection of implant material, engineering of modified implants, etc., have been developed. Simultaneously, by modifying the chemistry, morphology and structure of the metal implant surface, different properties than in bulk can be obtained. Therefore, a way of preventing implant failure and loss is applying coatings that improve the mechanical properties and corrosion resistance of materials and promote osseointegration by stimulating osteoblast recruitment and differentiation to achieve bone formation, as well as removing or killing pathogens, thus preventing biofilm formation. Coatings based on stimuli-triggered therapeutic strategies with improved drag bioavailability and fast on-site bone healing effects have been also proposed [6]. In that way, the performance and service life of dental and orthopedic implants can be substantially improved. Focusing on enhancing various shortcomings of metal implant surfaces, biocompatible coatings can offer a unique combination of properties that best functionalize the implants. For that reason, this review outlines the main materials used for metal implant manufacturing and some general coating techniques and focuses on recent trends in the design and performance of biomedical coatings for metallic implants used for orthopedic and dental applications. Some techniques for the reduction of infections and advanced combinatorial treatment are also described together with the associated drawbacks of the coated systems.

2. Bone Regeneration

The main objective for applying a biocompatible coating is to enhance the osteoconductive, osteo-inductive and osteogenic performance of the metal implant surface. Osteo-conductivity is the ability of the coating to act as a scaffold for extracellular bone matrix formation where osteoblasts can adhere and proliferate, while osteo-inductivity refers to the ability of the surface to stimulate the differentiation of precursor (stem) cells into osteoblasts. Osteogenic coatings alloy osteoblasts to produce calcium nodules for calcification of the collagen matrix of the newly formed bone structure [7].

2.1. Molecular Mechanisms

Mammalian bone can be found in two forms: woven and lamellar. While woven bone is characterized by a random distribution of collagen fibrils and mineral crystals due to its fast deposition, occurring during the development and repair of fractures, lamellar bone is more precisely arranged, being formed much more slowly [8]. Consequently, woven bone is mechanically much less stable than lamellar bone, thus being replaced by lamellar bone during remodeling.
The principal role of bone is skeletal support, providing body mobility and the protection of organs. For example, the jaw bones—maxilla and mandible—hold the teeth in place and transmit chewing forces from the muscles of mastication to the teeth. Bone architecture is highly adapted to the structural needs of high strength at minimum weight. The way to accomplish this is to form dense, highly mineralized cortical bone and porous trabecular bone, also called cancellous bone. Bone is also responsible for blood production, mainly in the bone marrow and bone acts as a storage medium for minerals such as calcium and phosphates, as well as proteins such as growth factors, which are deposited in the matrix and later released to responding cells [3]. The trabecular bone is the site of hematopoiesis, involved in fat and iron storage. Ca2+ homeostasis also occurs in trabecular bone and hematopoietic activity is limited to the bone marrow [9].
The bone is composed of a hierarchically organized composite material of fibrous type 1 collagen, non-collagenous proteins forming the osteoid and intercalated hydroxyapatite crystals. Although bone seems relatively static, it is a living and dynamic tissue consisting of three main cell types, osteoblasts, osteoclasts and osteocytes, as well as bone lining cells, which cover the bone surface and are marginally involved in bone remodeling [10]. Usually, osteoclasts and osteoblasts travel the bone surface, whereas osteocytes are embedded in the calcified bone matrix throughout the entire bone mass. Through canals (canaliculi), osteocytes remain connected to neighboring cells, thus forming a 3D cellular network in the bone structure.
Osteoblasts originate from pluripotent mesenchymal stem cells. Apart from differentiating into osteoblasts, these cells can differentiate into chondroblasts (forming cartilage), fibroblasts, adipocytes, myoblasts and other cell types [11]. The morphology of osteoblasts is dependent on the stage of functional differentiation. They attach to the extracellular matrix and other cells via transmembrane proteins, such as integrins, connexins and cadherins, which enable them to react to metabolic and mechanical stimuli [12,13]. The main task of mature osteoblasts is to synthesize proteins, such as collagen I and other glycoproteins such as osteopontin, osteocalcin and growth factors.
Mineralization of new hydroxyapatite is initiated and regulated through the formation of matrix vesicles (MV). By supplying sufficient calcium and phosphate ions, osteoblasts release MV from the cell surface into the extracellular matrix. Further mineral accumulation in the MV is achieved through ion channels and transporters leading to initial nucleation and calcification in the MV. Finally, needle-like hydroxyapatite crystals are released from the MV into the extravesicular fluid, these crystals subsequently being incorporated into the collagen fibril network [14,15]. Some osteoblasts become incorporated into the matrix and turn into immature osteocytes embedded in the fully mineralized and matured matrix [16,17].
Today, it is known that osteocytes are essential in both mechano-sensation and mechano-transduction [18,19]. Over 80% of all cells in the bone are osteocytes, which originate from highly specialized and fully differentiated osteoblasts, as mentioned above. Any microcracks in the bone can induce osteocyte apoptosis, which activates remodeling [20]. Bone-resorbing osteoclasts and bone-forming osteoblasts at the cellular level execute bone remodeling. Osteoclasts resorb existing bone and leave behind a resorption pit. They are followed by osteoblasts building new bone through collagen deposition and mineralization of calcium phosphate. An active osteoclast can resorb around 200,000 µm3 per day and this compares to 15–20 days of bone formation by seven to ten generations of osteoblasts [21]. In the bone-remodeling cycle, local communication and molecular interactions between osteocytes, osteoclasts and osteoblasts play a crucial role, together with the signaling system, involving RANKL (receptor activator of nuclear factor κB ligand), PTH (parathyroid hormone) and transforming growth factors (TGFs) [22].

2.2. Cells and Their Microenvironment

Every cell is embedded into a microenvironment—the extracellular matrix (ECM). This matrix consists of several proteins, such as fibronectin (Fn), collagen, elastin and laminin. The other main components are glycosaminoglycans (GAGs) which normally covalently link to proteins. These macromolecules are secreted locally and assembled into an organized meshwork by the cells. The cell’s microenvironment can be divided into geometrical, mechanical and chemical components. The geometry is determined by the cell shape, the mechanics by the rigidity of the microenvironment and the chemistry by the extracellular molecules with which the cell interacts. These properties of the local cell microenvironment regulate cell behavior in concert with autocrine and paracrine soluble or matrix-bound signaling molecules [23,24].
Typical amino acid sequences of the matrix proteins mediate cell adhesion. In general, when biomaterials are functionalized with peptides of this type, the peptides are adsorbed to the surface or are covalently linked to them with the use of different chemical reactions and surface modifications. Typical sequences involved in cell adhesion are Arginine-Glycine-Aspartic acid (RGD), Pro-His-Ser-Arg-Asn (PHSRN), Lys-Arg-Ser-Arg (KRSR), Asp-Gly-Glu-Ala (DGEA), etc. [25]. The well-known RGD sequence is a domain that occurs in many matrix proteins [26]. It is recognized by a range of cell types by various types of integrin receptors [27]. The PHSRN sequence has a synergetic effect with RGD, crucial for optimal integrin binding [28]. KRSR—a heparin-binding domain of fibronectin and collagen—plays an essential role in the selective adhesion of osteoblasts. It also inhibits the adhesion of fibroblasts, endothelial and epithelial cells [29,30]. It is recognized by the non-integrin receptor-cell-membrane heparin sulfate proteoglycan. DGEA is a typical sequence for type I collagen recognized by α2β1 integrins [31]. Similarly, as was done with the bone morphogenic protein-2 collagen binding domain (BMP2-CBD), these bioactive sequences can be designed and produced with additional domains to direct their attachment to various biomaterials, such as collagen or hydroxyapatite.
Angiogenesis is another important factor needed for an appropriate bone healing procedure. New blood vessel formation is essential for supplying the cells, nutrients and oxygen and for removing waste products. Angiogenesis is a complicated process involving various growth factors, ECM molecules and cell types. Among others, osteopontin is a phosphoprotein that participates in bone metabolism, mediates inflammatory responses and plays a crucial role in angiogenesis [32]. Recently, an osteopontin-derived peptide Ser-Val-Val-Tyr-Gly-Leu-Arg (SVVYGLR) has been identified as having pro-angiogenic properties, supporting the adhesion and migration of endothelial cells [33]. Moreover, it enables bone marrow stromal cell differentiation into endothelial cells [34].
Much effort has been invested in engineering implant surfaces that mimic and contain native ECM proteins. However, an adsorbed mixture of proteins with random unfolding, orientation and conformational states presents a divergence from natural, intentionally arranged protein layers. Immobilizing entire native proteins to surfaces can provide many functions because of the various domains within the molecule-engineered cellular environments [35]. An alternative approach is using peptides instead of complete proteins. Peptides based on the primary structure of the receptor-binding domain of an entire protein, such as RGD from fibronectin, aim to target specific cellular adhesion and interactions [35].
Normal tissue cells are dependent on anchoring to a substrate for survival. However, tissue can have a wide range of stiffness, with the brain in the soft range (about 1 kPa), bone in the hard range (about 100 kPa) and many others in between. Therefore, a cell experiences different tissue stiffness in vivo. Physical properties of tissue are also altered at a wound healing site and can change during disease progression [36,37,38]. The initial cellular response to mechanical signals occurs in seconds to minutes (Figure 1). A cell in culture for several days can be expected to experience countless stimulus–response cycles. The mechanism behind mechano-sensing is complex. Focal adhesions, focal complexes, integrins and cytoskeleton have been identified to play a key role in the mechano-sensing cells’ ability [39,40]. All these processes are essential for implant integration.
The clinical success of implant treatment is based on the concept of osseointegration, first described by Brånemark [41], today defined as “the formation of a direct interface between an implant and bone, without intervening soft tissue” [42]. Osseointegration is the antithesis of the fibrous encapsulation of a foreign-body response, where an implant or any other foreign material inserted in the body is encapsulated in fibrous tissue. Describing osseointegration is not possible without highlighting the eminent role of the implant materials and their surface [43]. In addition to obvious factors, such as cleanliness and sterility of the implant surfaces, which are indispensable, surface roughness and surface energy influence osseointegration. It has been shown in vitro, in vivo and clinically that rough implant surfaces show better osseointegration properties than smooth surfaces. There are other clinical fields in which the concept of osseointegration is applied, such as orthopedics, where it is named osteointegration. These areas will not be discussed further here but are outlined elsewhere [44,45]. The role of implant materials, preparations and surface modifications are described in the following text.

3. Requirements and Generations of Materials Used for Bone Implants

The requirements for implant materials can be both general and specific (Figure 2). The general requirements for biomaterials are related to (a) their clinical performance—to use a metal or alloy as a biomaterial, it should be biocompatible. This means that it must not be rejected and should avoid harmful effects on the host while working as a medical therapy for a certain response in the living organism [46], and (b) manufacturing criteria, including easy fabrication with complex geometry and affordable price.
Simultaneously, the biomaterial should comply with important specific requirements that determine its performance in a particular application. As seen in Figure 2, the most important implant-specific properties can be divided into (a) bulk, (b) chemical and (c) textural properties [47]. The bulk properties can be subdivided into mechanical properties and degradability. The mechanical properties of a biomaterial define its response under the influence of various forces. These properties are critical to consider, since during healing the implant bears much ununiform stress and load [48]. To avoid stress at localized points (stress shielding effect), the modulus of elasticity of both bone and implant should not show significant differences. When the biomaterial is used as a scaffold, its mechanical properties must be boosted in line with the degradability. The degradation rate that is dependent on different chemical and physical processes such as dissolution, phase transformations, etc., should correspond to the increase in new tissue formation. In bone implants, the degradation rate should be lower than 0.5 mm per year [49].
The chemical properties of metal biomaterials are linked to their chemical activity and corrosion resistance. Various factors such as a change in oxygen level or pH value in the bio environment or diffusion of ions may contribute to corrosion. The corrosion behavior of the material is dependent on its relative crystallinity, crystalline size and textural properties. The latter ensures the interaction with the bio-environment, tissue and cells and aids the healing processes. There is much evidence that the surface characteristics of the implants affect cell attachment, proliferation, migration and differentiation [50]. Some authors [50] divided the textural factors affecting the biocompatibility into two groups: (a) surface chemical properties such as hydrophilicity, free energy, polarity, crystallinity, electrostatic interactions and mobility of the surface functional groups; (b) surface physical properties such as surface topography, roughness, density, thickness and adhesion of the layer. Theoretically, the interaction of the cell with the outermost surface layer equals about 0.1-nm thickness [51]. The micro-surface roughness is important in providing contact guidance for filopodia attachment, while nanoscale roughness influences protein absorption that aids in cell attachment to protein through the integrin [52]. The wettability influences the conformation of the absorbed proteins on the surface [50] and, together with a higher percentage of crystallinity (over 70% [53] and optimal porosity level of about 30% for titanium scaffolds [54], the biocompatibility properties of the material are enhanced. To stimulate bone ingrowth, the minimal pore size should range between 100 and 150 μm [55,56].
Still, it is hard for a single metallic material to fulfill all general and specific requirements for biomaterials for implant application. Although metals and their alloys (first generation of biomaterials) meet many of the desired properties for biomaterials, their interfacial bonding with the surrounding tissue varies from poor to virtually absent [57]. Given the improved formation of adherent layer and reduced movement at the bone-implant interface, by surface modification of this first generation of biomaterials their specific properties have been upgraded. In contrast to the first generation of bioinert materials that form fibrous tissue capsules, the second generation of biomaterials focuses on the development of bioactive (generating hydroxyapatite layer on their surface) or biodegradable (bio-absorbable in a progressive manner) materials that promote specific cellular responses [58]. The third generation of biomaterials is designed to be temporary 3D scaffolds that stimulate tissue regeneration at a molecular level, angiogenesis and nutrient supply [59]. The fourth generation of biomaterials is based on personalized interaction with tissue and cellular processes depending on four requirements: receptivity, activity, autonomy and inertia [60].

4. Metals and Alloys for Bone Replacement

Metals and alloys are generally used for implants where high load-bearing capacity and strength are required. Compared to other biomaterials used for hard tissue replacement such as ceramics and polymers, metals display cost effectiveness, durability, mechanical strength and appropriate corrosion resistance. However, most of the alloys release metal ions in the surrounding tissue that enters the blood circulation and, in higher concentrations, the ions accumulate in the liver and spleen which can lead to cytotoxicity and organ failure upon prolonged exposure. They also suffer from impaired tissue growth because of inadequate attachment of the implant, higher risk of infections and slower healing time, compared to ceramic biomaterials [57]. Currently used metals and alloys in orthopedic metal-based implants consist of titanium-based alloys, cobalt-based alloys, magnesium-based alloys and stainless steels.

4.1. Stainless Steel

Stainless steel (SS) containing about 18% Cr and 8% Ni such as 316L and 304 is the most commonly used alloy for removable orthopedic devices such as screws, pins, bone plates, medullary nails, etc., because of its low price [61]. The presence of Cr allows the formation of a strong adherent Cr2O3 layer that favors corrosion resistance. Despite easy fabrication and low-cost, stainless-steel implants suffer from the stress shielding effect (because of their high elastic modulus of around 200 GPa), the occurrence of allergic reactions due to ion release [62], and are susceptible to pitting corrosion attacks. To reduce these adverse effects, small quantities of Mo (2–4 wt%), to improve the corrosion resistance and strengthen the alloy, have been added to the steel. Additionally, Ni-free SS with high corrosion resistance and biocompatibility has been developed [63,64]. However, compared with Ti-based alloy, the density (7.9 g/cm3) is higher while the osteointegration, biocompatibility and corrosion resistance of SS implants are lower which causes fewer implant success rates [64].

4.2. Co-Cr Alloys

Compared with stainless steel, Co-based alloys are characterized by superior strength. Despite their difficult fabrication, Co-based alloys are more resilient to corrosion and wear and display better biocompatibility [65]. Alloyed with Cr and Mo, some Co-based alloys like Co-Cr-Mo and Co-Ni-Cr-Mo are specially used for implants in the hip, knee and shoulder prosthesis [66]. Compared with other metallic implants, Co-based alloys display a better elastic modulus, stiffness and high density (8.3 g/cm3) [67]. However, the elastic modulus and ultimate tensile strength of Co-based alloys (200–250 GPa and 400–1000 GPa, respectively) are about 10 times higher than those of human bones which can result in a stress shielding effect. Co-based alloys are also not ideal for joint and bearing surfaces because of their sub-par frictional properties [68]. Additionally, their alloying elements such as Ni, Cr and Mo were proven to be toxic when released from the implant surface in the body fluid during corrosion processes. Excessive accumulation of these elements in organs such as kidneys, liver, lungs and blood cells can trigger their damage [65].

4.3. Ti and Ti Alloys

Ti and its alloys have an advantage over steel and Co-based alloys because of having low density (4.5 g/cm3), low weight and biocompatibility. Pure Ti has the advantage of being very highly corrosion resistant because of the formation of TiO2. Titanium alloys show better mechanical properties in comparison to pure Ti. About 50% of the total usage of titanium alloys falls on Ti6Al4V alloy because of its excellent formability, structural stability, weight-to-strength ratio and corrosion resistance [69]. However, together with Ti6Al7Nb, Ti6Al4V alloy suffers from high friction and low wear resistance [70]. Though proven to be highly biocompatible and corrosion resistant, during long-term implantation, α+β structured Ti6Al4V and Ti6Al7Nb alloys also suffer from the release of toxic alloying elements, risk of stress shielding and high cost [71]. The presence of Al and V in the Ti alloys results in a release of toxic ions in the body milieu under the physiological environment which triggers adverse health problems [72]. Hence, scientists focus on β structured alloys containing Nb, Zr and Ta that substitute Al and V in the composition. Such alloys are characterized by good mechanical properties that can be tuned by heat treatment, better formability, lower elastic modulus (down to around 50 GPa) and higher wear and corrosion resistance at the expense of a higher price [73].

4.4. Mg Alloys

In contrast to the above-mentioned alloys, Mg-based alloys are used for the production of biodegradable orthopedic implants that eliminate the need to be retrieved via a second surgery. The revision surgery is discomforting and expensive for the patients and can lead to possible complications and infections [74]. Mg and its alloys are among the lightest structural materials, with densities very similar to that of human cortical bone (1.75 g/cm3) [75]. Additionally, Mg is an essential element in the construction of soft tissue and bone. Mg and its alloys such as Mg-Ca-Zn, Mg-Zn-Ca-Mn with enhanced mechanical and corrosion properties have been applied for making bone screws and pins, wires, sutures, surgical clips, etc. [76]. Except for low density and lightweight, Mg-based alloys have excellent mechanical properties, elastic modulus close to that of bone and good biocompatibility [77]. A challenge in using Mg and its alloys is their high degradation rate in vivo via electrochemical mechanisms of dissolution. Besides, a decrease in mechanical properties due to corrosion, alkalization and hydrogen generation in the surrounding tissue is observed which can inhibit implant tissue interaction and promote tissue necrosis [78]. To overcome this shortcoming, scientists develop coatings or surface treatment procedures to postpone the start of degradation [79].
A comparison of some mechanical properties of the main metallic materials used for orthopedic implants is presented in Figure 3.
Although metallic implants are classified as the first generation of biomaterials, they evolve and are currently in large-scale use. Because of the wide variety of surface treatments modifying their roughness, wettability, topography, etc., the quality of bone-to-implant anchorage can be substantially improved.

5. Coating Methods

For the successful modification and functionalization of metallic biomaterials, it is necessary to develop techniques for changing their surface composition, morphology and structure without compromising their mechanical properties. Various coating materials such as metals, alloys, ceramics, glass, polymers, organic/inorganic hybrids, biomolecules and composites have been used to achieve bioactivity. The deposition process has to offer a high degree of flexibility and sufficient adhesion to the substrate, since it has to meet various demands for specific biomedical applications. To produce the coatings, high energy is applied to the system to achieve deposition through melting or vaporization and subsequent solidification or accelerated flow of particles toward the metallic substrate [80]. The resultant structure of the coatings can be divided into amorphous and polycrystalline, depending on the preparation conditions and the deposited material. Since the structure of the bone is microcrystalline while the amorphous phases are more prone to dissolution, the stable crystalline phases are preferred for biomedical applications.
Presently, coatings can be fabricated in various ways. The techniques can be divided into chemical, thermochemical, electrochemical, physical, electrophysical and 3D printing. The incorporation of ingredients by micro- and nanoencapsulation techniques are also promising options for controlled drug delivery.

5.1. Chemical Methods

The sol-gel method is usually used for the synthesis of oxide materials. It includes the preparation of precursor solution, deposition of “sol” (formation of a colloid suspension) onto the substrate and heat treatment for densification or “gel” (viscous or solid material) formation. The precursor materials can be both inorganic salts in an aqueous solution or metal alkoxides in organic solvents. The deposition of the sol is carried out via electrodeposition, spin- or dip-coating techniques (Figure 4). The technique is frequently used because of its simplicity, cost-effectiveness and ease of incorporation of active ingredients (such as growth factors recombinant human bone morphogenetic protein-2 (rhBMP2) [81] and ability to adjust coatings with different thicknesses to substrates with various shapes [82]. The abundance of chemical and physical parameters that influence the structural properties of the coatings renders the possibility of effective control over their structure and homogeneity. The obtained coatings can be porous which increases the surface area in contact with tissue and fluids, but the adhesion ability of the coatings to the metal surfaces is low due to thermal mismatch. To enhance adhesion as well as density, sol-gel coatings can be sintered at high temperatures (more than 1000 °C) or an intermediate layer can be formed [83,84]. Because of the long processing time, the need for post-sintering and the cracking of the final coating due to shrinkage, the application of the sol-gel route is limited in industry.
Sol-gel techniques are also used for the production of hydrogels and aerogels with a large specific area and high porosity at moderate temperatures. Hydrogels are formed from hydrophilic polymers such as polyvinyl alcohol (PVA), polyacrylic acid (PAA), polyglycols, hyaluronic acid, collagen, gelatin and dextran, with elastic and swelling properties, while aerogels are obtained when wet gels (usually silica based) are super-critically dried and the liquid phase is replaced by a gaseous phase. Thus, aerogels with high porosity, low density and thermal conductivity are formed [85]. Polysaccharides such as methoxyl pectin and xanthan incorporating indomethacin, are also used to prepare aerogel coatings on SS substrate after supercritical CO2 drying [86]. However, hydrogels have low adhesion to a metal surface which imposes physical or chemical treatment that bonds or crosslinks the hydrogel to the surface by chemical or physical bonds [87].
The electroless deposition method, also known as plating, chemical bath deposition, solution growth technique, or controlled precipitation, is mostly used to prepare metal oxide films at lower temperatures. During the process, metal ions are complexed by ligands by controlling the pH and then the substrate is immersed under a temperature range of 60 to 100 °C to form, usually, metal hydroxide films. After that, the hydroxide film is transferred to the oxide by annealing. This technique is simple, unexpensive and suitable for complex geometries. The process is used for the deposition of the biodegradable and biocompatible iron coating on pure Mg to increase its corrosion resistance [88], as well as to produce antibacterial ZnO and hybrid ZnO/Ag on biodegradable Mg-Ca alloy [89]. Although reliable, electroless deposition is a lengthy process that requires the chemical bath to be replenished often, while the coatings may show imperfections such as microcavities with columnar structure and pitting marks [90].
Employing heterogeneous nucleation and crystal growth of coating, biomimetic deposition of films with bonelike properties can be obtained [91]. The coatings are prepared, initially pretreated either with acidic or alkaline solution metallic substrates by immersion in simulated body fluid (SBF) at body temperature (37 °C) and physiological pH (7.4). After several weeks, CaP-based coatings with different Ca/P ration and crystallinity on the substrates are deposited [92]. This method allows additional bioactive components such as osteogenic agents to be co-precipitated and incorporated into the structure of the coating [93]. To enhance the mechanical properties, carbon nanotubes can be incorporated into the coating structure by adding them to the SBF solution [94]. However, this method requires a long processing time and maintenance of ion concertation, while the coatings suffer from poor bonding strength.
The wet-chemical methods also include common techniques for covalent bonding of target molecules onto metal (titanium) surfaces using phosphonic acids [95] or catechol [96] to create organofunctional anchors to graft bioactive polymers with different architectures. This kind of bonding counts on the covalent connection of end-functionalized polymers that polymerize in situ, initiated from the surface. Thus, demanding diffusion of monomers towards the propagating radicals is required. For that reason, Chouirfa and co-authors developed a technique based on catechol surface modification designed to enable bioactive ionic polymers bearing thiol end groups to be attached using a thiolene click reaction [96].
To produce coatings with high purity and crystallinity at low setting-up costs, spray pyrolysis techniques can be used. They employ processes held at atmospheric pressure in the air at ambient conditions. The setup consists of generators (such as ultrasonic, electrospray, centrifugal, pressure and pneumatic atomization) that produce mist from a spraying solution containing the precursor material [97]. The mist is transported to the surface of the substrate which is heated to achieve pyrolysis of the solution, thus forming a solid coating on the top. The synthesis of nano-coatings can be flame-, plasma-, microwave-, or laser-assisted [98]. However, post-deposition annealing may be necessary to obtain the required coating characteristics [99]. Depending on the solvent and deposition parameters, solid or molten metal moieties may vaporize upon approaching the substrate and then the coating may grow via a true chemical vapor deposition (CVD) reaction. Recently, the method has been used for the production of Zirconia-incorporated bioactive glass films on pure Ti substrates [100], thin oxide films [101] and ultra-porous HAp network on Ti alloy discs [102] by using flame spray pyrolysis. It is a one-range synthesis process with a high scalability and production rate that is capable of producing coatings of a wide range of material compositions on nonplanar implants.

5.2. Thermochemical Methods

Thermal spraying is a process where molten or semi-molten micrometer-sized particles are sprayed onto a substrate at a high speed. Depending on the heating source, the spraying process can be divided into electric arc, flame, plasma, kinetic spraying, etc. while the process can be held both at atmospheric pressure or in a vacuum. The principal of coating deposition by using a plasma spraying technique is illustrated in Figure 5. Thermal spraying is a favorable process to deposit bio-ceramics, bio-glasses [103], polymer [104] or composite coatings such as hydroxyapatite (HAp)/Ti [105] on metal implant materials. Due to the high impact velocity, the coating material particles are flattened in the form of lamellae or splats. Although the process is cost-effective with a high deposition rate and the coatings’ surfaces are micro rough, the obtained films are usually porous, inhomogeneous, amorphous with low adhesion strength, and prone to coating spallation due to compressive residual stresses [80]. Though the increased surface roughness is considered as an enhancing factor in the implant–bone interaction, the roughness should be in a certain range. The process is less suitable for temperature-sensitive materials, biological molecules and drug release applications due to the high process temperatures. Additionally, mechanically bonded coatings with different bond strengths are poorly accepted for biomedical applications [93]. To minimize the above-mentioned disadvantages of the high-temperature thermal processes, cold spraying, where the particles are heated below the melting temperature and accelerated by carrier gas toward the substrate, has also been conducted for biomedical applications [106]. The heated particles participate in the successful deposition of coatings such as HAp/graphene [107] or Ta [108] on metal substrates while retaining their crystal structure even after deposition.
Another high-temperature technique is hot isostatic pressing in which high isostatic pressure at high temperature is used to obtain dense ceramic coatings. The application of uniform pressure all over the system allows for removing the shape limitations of the substrate. For instance, this method is used for the deposition of HAp coating at a relatively low temperature (135 °C) and density close to that of bone [109].
The hydrothermal crystallization method is a single- or multistep process for coating deposition from solution onto a metal substrate at different temperatures (up to about 200 °C), various pH and high autogenous pressure in a hydrothermal reactor. During the single-step process, due to heating, the pressure increases in the reactor and coatings deposit on the sample surface in a supercritical environment. The multistep process is applied to already existing coatings to form a targeted phase. A single-step deposition method of the CaP layer is used for improving the corrosion properties of Mg-based alloys [110], while a multistep process was applied for the deposition of HAp on either Mg alloy or Ti6Al4V [111]. Due to its simplicity and cost-effectiveness, the hydrothermal method is applied in mass production [112].
Chemical vapor deposition (CVD) methods form films through a chemical reaction between the incoming species on the metallic surface. The chemical reaction that occurs during the CVD processes depends on the volatile precursor transported via the vapor phase and the by-products that together form a non-volatile solid through a chemical reaction. Liquid or gas source materials are used during the various variations of the CVD methods such as plasma-enhanced CVD (PECVD), metal-organo-chemical vapor deposition (MOCVD), low-pressure CVD (LPCVD), pulsed CVD, spray, chemical bath, etc. [113]. The energy used to drive the chemical reaction can be laser, photon, or temperature. Differently structured films (amorphous, polycrystalline, epitaxial and uniaxial) with a high degree of purity can be deposited. The main disadvantage is related to the limited components that are volatile at nearly ambient temperatures, the generation of toxic by-products, or the use of toxic precursors [114].
Pulsed CVD, also known as atomic layer deposition (ALD), is based on self-limiting reactions between two gaseous precursors and allows for the deposition of thin oxide films layer-by-layer at relatively low temperatures and moderate pressure. The ALD method was used for the deposition of amorphous and homogenous TiO2 layers [115] or to enhance the functionality of nanotubular Ti surfaces [116]. The thermodynamic complex processes that occur during CVD depend on temperature, pressure, flow rate concentration of chemical species and reactor geometry [117]. The process is resourceful for producing simple and complex compounds in the form of coatings with uniform thickness and controlled properties, but requires expensive equipment.

5.3. Electrochemical Methods

Electrochemical methods produce good quality films at a low price without the need for expensive equipment, while the waste is limited to the solution. They depend on the solution composition, pH value, viscosity, etc.
By electrophoretic deposition (EPD), an electric field is applied between two electrodes and various charged particles, such as fine powders, colloids, or macromolecules that are dispersed or suspended in the liquid, move toward the oppositely charged electrode thus leading to accumulation (by coagulation and precipitation) on the deposition electrode (substrate) (Figure 6). This technique can be employed in the production of homogenous bio-ceramic HAp films [118] on metallic substrates as well as polyacrylic acid films containing nanotubes and oxide particles [119]. The process requires a short process time and gives the ability to produce interconnected porous coatings with controlled thicknesses up to 2 mm [120]. The EPD technique also allows the incorporation of bioactive ingredients such as antibiotics [121] and zoledronate (medication used to treat several bone diseases) [122]. The main issues are related to the coating adhesion, occurrence of cracks and difficult scalability [123].
During electrochemical deposition, the substrate is placed as a cathode at either constant or pulse current/voltage. The latter gives more possibilities of controlling the properties of the coating by the process parameters. Good shape conformity, uniform coating thickness, short processing time and low processing temperature (usually room temperature) are the main advantages of electroplating techniques. The method allows the synthesis of metallic [124], ceramic [125], polymer [126] and composite [127] coatings. The major disadvantages include inadequate bonding to the surface and the development of stresses in the coatings [128]. The critical factor for adhesion is surface roughness. For ceramics that cannot dissociate from ions, their deposition on the metal base is difficult and the obtained surfaces are often non-homogeneous and non-stoichiometric [129].
Besides the implementation of coatings, another electrochemical strategy is changing the topography by forming a passive layer on the surface when exposed to oxygen. One of the most effective techniques is anodization. Thus, self-assembled titania nanotubes (TNTs) can be fabricated on the surface of Ti and its alloys. By varying the applied potential, time, electrolyte temperature, pH, concentration, etc., the characteristics (diameter size, branched tube, double-walled, double-layered, etc.) of TNTs can change [130] which are known to affect osteoblast cell adhesion and proliferation [131]. The TNTs can be tailored with drug-loaded properties because of their open-ended tubular shape. The loading procedure can be achieved by different conjugation strategies (for growth factors such as bone morphogenic protein-2 (BMP-2) [132] or using vacuum-freezing technologies for drugs such as antibiotics [133]. Despite these advantages, the mechanical stability of TNTs under load-bearing conditions as well as their cytotoxicity caused by the released ions or debris is still under consideration.
Among the electrodeposition methods, plasma electrochemical oxidation (PEO), also called micro-arc oxidation (MAO), is a technique developed on anodic oxidation that brings out a rough surface rich in biological elements due to high-voltage oxidation in electrolytes containing different compounds. The formation of MAO coating occurs both due to the oxidation of the substrate and the incorporation of compounds or particles from the electrolyte (Figure 7). It is a hybrid of conventional electrolysis and an atmospheric plasma process to produce well-adherent coatings on different valve metals. The advantage of MAO is that it can produce highly microporous oxide layers on implant surfaces, which is important for anchoring the bone [134]. The electrolyte may be silicate, phosphate, aluminate, sodium tetraborate, or phytic acid to produce various coatings. For example, using calcium acetate and sodium phosphate crystalline TiO2/Hap, coatings on Ti substrates have been prepared [135], though HAp produced by MAO is usually of low crystallinity. The HAp crystallinity and surface roughness can be enhanced by increasing the voltage and duration of the MOA process, which subsequently triggers the appearance of cracks and failures within the coating. An alternative method to increase crystallinity is to apply subsequent hydrothermal treatment [136], ultrasonic assistance [137] or immersion in SBF [134]. The MAO process is stable and reliable, with good reputability and environmental friendliness [138]. Besides, owing to its ability to form highly adherent microporous coatings, its simplicity and its cost-effectiveness, MAO is becoming an indispensable technique for the generation of functional nanostructured coatings. However, the pores can serve as channels for corrosive electrolyte ingression. Besides, in the most common Ti alloy (Ti6Al4V), a large amount of mixed oxide aluminum titanate is produced [139].

5.4. Physical and Electrophysical Methods

The deposition process during the PVD processes takes place in a vacuum chamber. The deposition of coatings by gas-phase processes such as evaporation and sputtering produces a thin film with good purity and structural properties. The transport of free species emitted by the target and accelerated towards the substrate allow controlling of their electric charge and kinetic energy. PVD techniques are environmentally friendly, the coatings have superior mechanical biocompatible properties and they can be applied to temperature-sensitive substrates [140]. However, these methods suffer from some drawbacks such as high processing and equipment costs, the need for a large number of material targets, gaseous waste treatment, strict instrumentation requirements and the inability of drug release applications.
Vacuum thermal evaporation is a simple PVD technique that is usually used for obtaining amorphous thin films. The material is resistively heated below 1500 °C in a vacuum condition until vapor pressure is produced. During the process of evaporation of the covering material, a potential difference to the substrate is applied at a medium or high-vacuum level. Commonly evaporated films of Ti, TiO2 and calcium phosphate are used to tailor the chemical and topographical properties of biomedical materials [141]. During electron beam evaporation, target material is hit and vaporized under vacuum by an intensive beam of electrons. A magnetic field is also applied to focus and change the beam trajectory. Large categories of coatings, such as amorphous and crystalline metals, oxides and molecular materials, can be produced [142]. Similarly, during laser beam evaporation a laser beam is used to ablate the material for depositing the coating and produce the plume with neutral or ionized species. The morphology of the film is affected by the substrate temperature, while the process is characterized by a fast deposition rate and compatibility with inert gases and oxygen for oxide film deposition. Several variations of laser beam evaporation such as pulse laser deposition [143] and pulse plasma deposition [144] for the deposition of ceramic films are also currently investigated. However, problems due to high temperatures and lack of uniformity occur. An electric arc at high current and low voltage that passes across the surface of the target is used to vaporize the material during cathodic arc deposition. The metal atoms become highly ionized, which makes the process suitable for the formation of thick films, but the released microparticles from the target can make the coating susceptible to corrosion processes.
Sputtering is an etching process involving the backward scattering of solid surface atoms upon bombardment by energetic ions to obtain glow discharge in the residual gas (reactive or inert) in the vacuum chamber. The sputtering process depends on the bombardment of the ions released from the discharge that hit the cathode to liberate particles (mostly neutrals) with high kinetic energy (Figure 8). The latter strike the substrate or anode to form a dense film [145]. Common sputtering systems adopt radiofrequency (RF), direct current (DC), magnetron diode, or ion beam sputtering. Although the deposition rate is lower than that of evaporation, the deposited films have a uniform dense structure, good adhesion to the substrate, high purity, outstanding adhesion to the substrate and composition similar to that of the starting material. Compounds such as oxides, nitrides and carbides can be formed by flowing reactive gases as well as composite materials such as Ag-HAp, Ag-SiC-HAp, Sr-HAp, Ti-Si-N, etc. [80]. However, it is hard to deposit uniform coatings on implants with complex shapes, as well as films with higher thicknesses and roughness. Even more importantly, the length of the process and its cost limit the sputtered coatings’ application.
The ion implantation technique is a process in which ions extracted from plasma by an extraction system accelerate in the form of a beam with high energy towards a substrate. Because of the small cross-sectional area of the beam, the beam must be rotated to achieve uniform implantation on flat large surfaces, whereas for complex geometries sample rotation is required which limits its size [141]. Plasma immersion ion implantation and deposition allows simultaneous and sequential ion implantation and deposition by combining conventional plasma and ion beam technologies. Under negative sample bias, positive ions are accelerated from the plasma while the plasma sheath around the sample governs the ion implantation process. In that way, advanced treatment of devices with complex shapes and large implantation areas for shorter process time is carried out [146]. This process was used to improve the biocompatibility and bioactivity of Ti6Al4V alloy by the formation of surface TiN and TiC layer [147], as well as to improve the corrosion resistance of magnesium alloys by adding metallic elements such as Zr [148] and metalloids like oxygen [149]. The method provides a strong adhesive bonding of the coating to the surface. Its main limitations are the formation of amorphous coatings and expensive costs [150].
Laser cladding is another electrophysical technique that uses a high-power density laser beam to melt the surface of the material which rapidly cools down to form cladding layers with changed surface structure and properties [151]. This process was used for the deposition of bioactive glass onto ultra-fine-grained Ti substrates [152], where the laser beam is focused onto the metallic substrate to melt it, while bioactive glass in powder form is delivered by inert gas. Due to the localized heating of the laser beam, the properties of the substrate are slightly changed but the homogenization of the microstructure in the cladded coating improves the mechanical properties and realizes metallurgical bonding with the substrate [153]. Although substantially roughened, the surface can exhibit pores and cracks.
Another electrophysical method is the electrospinning technique for the production of fibers with nano- or microscale diameters. Through a high-voltage supplier, a liquid droplet of a material solution is dispensed out of a nozzle of a syringe under a high voltage. At a critical voltage, the droplet becomes charged and stretches due to repulsion and surface tension [154]. The tip of the syringe act as an electrode while the collector is the counter electrode. Attracted by the counter electrode, the stretched droplet moves to the collector, while the solvent evaporates triggering the solidification of the material. Although polymers such as polycaprolactone (PCL) and polylactic-co-glycolic acid (PLGA) are typical starting materials for the production of fine fibers, metals, ceramics and composites can be also used [155]. The resulting filaments display the potential to serve as bioactive coatings due to the high porosity, the implementation of biocompatible and biodegradable materials and the incorporation of active ingredients. For example, loaded with antibiotics or other drugs, the coatings undergo initial burst release within the first few days and, later, slow controller release which can last more than a month [153,154,155,156,157,158].

5.5. Additive Manufacturing Methods

3D printing is a technology that constructs three-dimensional objects following a computer-aided design (CAD) or three-dimensional (3D) scanners in a layer-by-layer method [159]. The technique offers flexibility in terms of materials used, obtaining complex geometric shapes, time efficacy and low production costs [160]. Among various 3D printing technologies, the most commonly investigated methods for the development of bioactive coatings for orthopedic implants are extrusion-based methods such as fused deposition modeling (FDM) and semi-solid extrusion (SSE). During the FDM, the printed thermoplastic materials are softened and melted using a heated nozzle, built up layer-by-layer and, after solidifying, the filaments form a 3D coating on the metal surface. In SSE, semi-solids are, layer-by-layer, extruded using air pressure or screw gear rotation through a nozzle to form a 3D coating. A combination of FDM and hot melt extrusion is also used for the production of amorphous materials with active ingredients. Initially, the thermoplastic material and the active ingredients are mixed by pumping them at a temperature above the glass transition temperature with a rotating screw and after that FDM is printed to obtain the desired structures [161]. Additively manufactured composite coatings printed from cellulose nanofibril suspension, alginate and carboxymethylcellulose and further loaded with antibiotic clindamycin were deposited on SS 316 LVM and Ti6Al4V substrates [162]. Although not extensively explored for the production of bioactive coatings, 3D printing technologies are promising candidates for the development of personalized shelled systems for orthopedic applications or porous scaffolds for controlled drug release [163,164].

5.6. Combined Methods

The diverse applications of coatings have led to combining various conventional and unconventional assembly methods for the deposition of bioactive films. For example, in the layer-by-layer deposition method, the substrate is covered by oppositely charged constituents attached by electrostatic attraction, hydrophobic interactions, hydrogen bonds, or van-der-Waals forces by different methods [165]. They include the dip method of immersing in polymer or colloid solution followed by the spin method and spray method that allow faster layer deposition [166]. Recently, electromagnetic assembly applying electric current to electrodeposit the layers, fluidic assembly in which the substrate is fixed in the fluidic channels (tubes/capillaries) and the coating is deposited by the gradual movement of polyelectrolyte and rinse solutions by pressure or vacuum; as well as ink-jet printing, 3D bioprinting, inorganic-organic hybrid assembly, etc. have been developed [167]. In this way, layer-by-layer obtained coatings offer flexibility in production at mild temperatures, implementation of a wide range of materials and the ability for functionalization with peptides, drugs, etc. [168,169]. The number of layers and the incorporating method determines the kinetic profile of active substance release.
Table 1 summarizes some of the important advantages and disadvantages of the discussed deposition techniques for the deposition of bioactive coatings.

6. Bioactive Coatings

Surface modifications of metallic implants have been extensively explored to hinder a range of adverse effects such as long-term stability, low biomechanical properties, lack of biocompatibility, restriction of implant surface corrosion, post-surgery infections, etc. Therefore, the development and design of biometallics rely on surface modification since, by applying appropriate coatings, the surface properties can be tailored and improved. In connection with this, two approaches for surface modification have been applied: (a) direct coating deposition on the unmodified metal substrate; (b) initial substrate modification by grinding, sand-blasting, etching, or other treatment and deposition of overlaying coating. In the second case, improved surface roughness parameters for synergy of both textural properties and mechanical interlocking of coating are obtained.
The main requirements for a selection of coating material include (a) adequate mechanical reliability, adhesion strength and fracture toughness to withstand the applied forces; (b) corrosion resistance in the body fluid environment; (c) biocompatibility and lack of toxicity, allergic or other undesirable effects or inflammation [171]. Depending on their performance in the organism, the biomedical coatings can be subdivided into three main groups: bioinert, bioactive and bioresorbable [78]. In contrast to bioinert (such as Al2O3 and ZrO2), bioactive coatings refer to biomaterials that can stimulate the surrounding tissue and cells to regenerate around the exogenous graft and to release bioactive molecules for elimination the post-operative complications [172]. Absorbable (bioresorbable) coatings are designed to degrade in the human body via an electrochemical mechanism of dissolution and then metabolized by cells and tissue [173]. Recently published research works addressing osseointegration of inorganic bioactive coatings are given in the next paragraphs.

6.1. Inorganic Coatings

Various research groups focus on the development of inorganic coatings for biomedical applications because of their stability in the body environment, good mechanical properties, corrosion and wear resistance. To be called “bioactive”, these coatings should have surface-located functional groups that in an aqueous solution create conditions favoring heterogeneous mineral nucleation and growth on the surface. However, some of these inorganic coatings have shown disadvantages, including the release of toxic ions, cytotoxicity, lack of biodegradability, low bonding strength, etc. Some recent studies on the wide range of such inorganic coating materials are reviewed in the next sections.

6.1.1. Nitrides

Various transitional metal nitride coatings such as TiN, ZrN, NbN, TaN, etc. have been studied as protective films against wear and corrosion of medical metal surfaces of the prosthesis and surgical implants [174]. Nitride films have a high melting point, chemical resistance to oxidation and acceptable adhesion [175]. Titanium nitride (TiN) films are often used in industry because of their high surface hardness and chemical properties. For biomedical applications, TiN was found to be well tolerated by tissue due to its inertness [176]. On orthopedic implants, nitride coatings protect the surface against wear and act as a diffusion barrier, preventing ion release from the metal to the body fluids [177]. Compared to other nitrides, TiN shows better biological properties for orthopedic applications [178]. However, because of dissimilarities in the hardness of TiN coating and substrate, plastic deformation at the coating/substrate interface occurs resulting in the formation of flakes, defects in the coating and fracture [179]. Compared with CVD-deposited TiN coatings, PVD utilizes a higher deposition rate and delivers improved bonding strength [180]. To improve the adhesion strength and wear performance “hard” nano-TiN and “soft” Ti4N3−x transitional phase with variable composition was prepared by DC magnetron sputtering on Ti6Al4V alloy [181]. The coating showed excellent bonding and wear resistance because of the match of the mechanical properties at the substrate/coating interface together with good biocompatibility. Except for higher resistance to plastic deformation and improved wear behavior than the bare Ti20Nb13Zr (TNZ) alloy, TiN coating deposited by the cathodic arc PVD process indicated better corrosion resistance in SBF than TNZ [182]. Compared to bare Ti substrate, TiN coating showed approximately eight times more corrosion resistance and 4 times more wear resistance [183]. For that reason, sputtered TiN was used as an interior layer between HAp films and pure Ti rendering the controllability of thin film structural properties because of the improvement in bonding, strong fatigue resistance and better mechanical performance of the HAp layer [184].
Because the release of metal ions such as nickel, cobalt and chromium may cause serious problems in joint replacement due to metal hypersensitivity, considerable attention was also paid to PVD deposited ternary nitrides such as TiNbN that can act as a surface coat to hide the metal beneath affording an immuno-privileged state [185]. Copper-doped TiN (TiCuN) deposited by axial magnetic field enhanced arc ion plating has proved to own excellent corrosion resistance, wear resistance and good antibacterial properties and displayed no cytotoxic effect on human umbilical vein endothelial cells [186]. Compared with TiN coating, TiCuN promoted mRNA expression of endothelial nitric oxide synthase (eNOS) and vascular endothelial growth factor (VEGF), enhancing cell migration and angiogenesis ability. Similarly, ZrN/Cu coating deposited on SS and Ti substrate enhanced their wear- and corrosion resistance in SBF and their antibacterial activity by using ion release and a contact killing mechanism [187].
Similar to other transitional metals, Ta has a high affinity to nitrogen and forms inert TaN. TaN is also a commonly used material for the production of hard thin coatings because of its chemical inertness, corrosion resistance [188] and biocompatibility [189]. Mendizabal et al. discovered that the highest corrosion resistance was observed for Ta and low nitrogen content TaNx films (lower than at30%) when deposited on AISI 316L steel by modulated pulse power magnetron sputtering [190]. The excessive amount of nitrogen on the film worsened its corrosion protection. Deposited on Ti substrate, the magnetron sputtered TaN film exhibited stronger bonding properties than TiN, and optimal compressive performance [191]. Deposited on Mg alloy by reactive magnetron sputtering, TaN exhibited a 95-fold decrease in corrosion current density in SBF solution compared to uncoated Mg-Y-Re alloy [192]. Additionally, the incorporation of Ag and Cu in TaN nanocomposite films gave the condensates anti-bacterial and anti-wear properties [193].
In contrast to crystalline TiN and TaN, amorphous SiN was also shown to be biocompatible and slow dissolving in a water-based solution [194,195]. It also has high hardness (up to 26 GPa) and Young modulus (up to 212 GPa), low wear rates, acts as a barrier for ion release from the metal and generates biocompatible wear particles [196]. However, some challenges such as inadequate adhesion and high dissolution rate that can reduce scratch resistance and cause premature coating failure are still faced by these coatings [197]. By alloying SiN with Fe and C in the Si-Fe-C-N system [198] or with Nb and Cr in the Si-Nb-Cr-N system [199], better adhesion, optimized dissolution rate and ion release can be obtained without compromising the adherence and morphology of MC3T3 or L929 cells. Overall, the composite coatings have low surface roughness, high hardness and elastic modulus and no evident cytotoxicity as opposed to SiN controls and CoCrMo alloy. With the increase of Cr between 4 to 6 at% the release of Si, Cr and Nb ions and the dissolution rate of the coating reduced, while the cell viability was reduced.

6.1.2. Oxides

TiO2 is an important material in biomedical applications, since it has good mechanical properties, antibacterial and catalytic activity and long-term stability under photo- and chemical corrosion [200]. TiO2 can promote the formation of bone-like apatite or calcium phosphate on its surface when soaked in SBF solution which makes it suitable for bone reconstruction and replacement [201]. Moreover, it was found that the formation of TiO2 coating by anodization on the surface of Ti substrates was an effective method to reduce the temperature rise of the implant during microwave diathermy treatment that would provide a potential rehabilitation solution to internal fixation of bone fracture [202]. Similar to TiO2, tantalum oxide (Ta2O5) could facilitate the formation of bone-like apatite and stimulate the rapid attachment of bone and soft tissue [189]. Tantalum oxide produced by reactive magnetron sputtering was able to enhance both early-stage corrosion resistance and in vitro biocompatibility of Mg alloy [203]. Ta2O5 coating with 12.5 at% Ag deposited by a twin-gun magnetron sputtering system exhibited both improved antibacterial effects against S. aureus and good skin fibroblast cell cellular biocompatibility [204].
Grafting metal ions and compounds is also a common method to improve the osteogenic ability of oxide coatings, since various metal ions (Ca2+, Sr2+, Mg2+, etc.) have been demonstrated to possess the property of enhancing osseointegration [205]. An illustration of the influence of different metallic ions on various processes involved in bone regeneration is presented in Figure 9. Zhao and co-authors used MAO to produce Mn-TiO2 coatings on the Ti surface that showed good biocompatibility and osteogenic properties while the Mn2+ release from the coating promoted surface biomineralization [206]. TiO2 nanotubes produced by anodizing and loaded with Sr combined with icariin (ICA) showed a better effect on cell adhesion, proliferation and higher mineralization activity than pure Ti and TiO2 coatings. Furthermore, in osteoporotic rats, more bone was formed around the implants loaded with Sr and ICA [207]. Similarly, Y-doped TiO2 coatings on Ti6Al4V produced by PEO demonstrated good biocompatibility on osteoblastic precursor cells and fibroblast cells with increasing doping concentration of yttrium and excellent antibacterial activity against E. coli and S. aureus [208]. MAO processed Fe2O3/TiO2 composite coating on Ti implants with sensitivity to the micro-magnetic field because of super-para-magnetism, which was able to enhance fibroblast response including proliferation, phenotype and extracellular collagen secretion by increasing the amount of Fe2O3 NPs [209]. Compared to pure TiO2 coatings, Fe2O3 (4.41 wt% Fe)/TiO2 composite coatings reduced bacterial growth by 60% and efficiently prevented recession and inflammatory reaction of soft tissue. By immobilizing anionic polypeptides on the surface of TiO2 nanospike coating by coordination, Gao et al. demonstrated that the obtained film was able to kill pathogenic bacteria, inhibit biofilm formation for up to 2 weeks and promote the formation of HA on the surface [210].
TiO2 can not only be used to promote osseointegration but, also as a photosensitizer. As a stable photocatalyst, TiO2 produces ROS to kill bacteria under UV radiation but UV light is harmful to the body. For that reason, Nagay et al. prepared N- and Bi-codoped TiO2 coating on Ti by PEO that produced ROS to kill microorganisms under visible light [211]. By embedding silver (Ag) and zinc (Zn) nanoparticles into a 3D printed porous titanium oxide layer, the surface promoted the release of Ag+ and Zn2+ which favored antibacterial effect and osteogenesis, respectively [212]. Moreover, this synergetic effect was able to reduce the toxicity of Ag to the host cell. In contrast to bare Ti substrate and undoped TiO2, Ag-doped TiO2 coatings produced by sol-gel technology enhanced the corrosion resistance of Ti in SBF solution [213]. Nanoclusters of Ag incorporated in silica coatings obtained by RF co-sputtering technique displayed both good adhesion on steel substrate and antibacterial activity against S. aureus [214].
Aiming at modifying the surface of TiO2 to generate reactive oxygen species (ROS) to eradicate bacteria under near-infrared (NIR) light, different photosensitizers can be used. Chai et al. synthesized hydrothermally produced MoSe2 nanosheets on the surface of porous MAO-prepared TiO2 coatings and covered them with chitosan by electrostatic bonding to improve biocompatibility [215]. Under NIR irradiation because of the synergistic effect of hyperthermia and ROS generation, the coatings demonstrated excellent in vivo and in vitro antibacterial properties against S. mutans, whereas chitosan improved hydrophilicity and biocompatibility of the hybrid coating, promoting osseointegration even in the presence of infection under NIR light. Han and co-authors chose MoS2 with a broad spectral response to modify the surface of composite collagen/polydopamine/TiO2 coatings on Ti implants produced by MAO and hydrothermal treatment [216]. Under the combined action of photothermal and photodynamic therapy, the biofilm of S. aureus was quickly eradicated while the collagen promoted the adhesion and proliferation of osteoblasts. TiO2 nano-shovel/quercetin/L-arginine coatings doped with ytterbium (Yb) and erbium (Er) exhibited the production of ROS under near-infrared II light irradiation that could kill bacteria. At the same time, ROS catalyzed the release of nitrogen oxide (NO) free radicals from L-arginine which promoted angiogenesis and osseointegration [217]. The electrons and hole complexes generated by TiO2 reduced the photocatalytic properties, whereas the nano-shovel structure and quercetin that was coupled to the surface by organo-silanes, promoted the differentiation of bone marrow stem cells (BMSCs). Li et al. produced thermosensitive chitosan-glycerin-hydroxypropyl methylcellulose hydrogel (CGHH) to layer the top of simvastatin-loaded TiO2 nanotubes [218]. At 37 °C, the CGHH was found to be in a sol state which facilitated the controlled release of simvastatin to enhance MC3T3-E1 cell differentiation. The results of subcutaneous infection animal models indicated that CGHH had almost no antibacterial activity but, at high temperatures caused by infection, GCHH transitioned into a gel state and released a large amount of glycerin that induced acute inflammatory reaction and antibacterial activity against S. aureus and E. coli.
TiO2 can be used as a sound sensitizer to produce ROS by ultrasound-triggered electron-hole separation. Applying photoacoustic therapy sulfur-doped titanium oxide (S-TiO2-x) to titanium implants endowed the implant with good sonodynamic and photothermal properties [219]. Under NIR irradiation and ultrasound, the killing rate of S. aureus was equal to 99.995% after 15 min of exposure while the coating displayed good stability after soaking in water for 6 months.
The principle weakness of bio-ceramics originates from the low mechanical strength that makes them inappropriate for load-bearing application. When combined with metallic implants and bio-ceramic films, their mechanical properties are preserved while the integration with the bone is improved. However, the metal–ceramic interface accumulates residual stresses causing delamination at the interface.

6.1.3. Oxynitrides

To combine superior mechanical properties such as hardness and adhesion to the substrate and enhanced corrosion resistance, oxynitrides of transitional metals have been developed. Moreover, transitional metal nitride-oxide coatings are interesting materials because of their low degree of dissolution, corrosion resistance and inertness in body fluids [220]. The oxygen addition in cathodic arc deposited TiN decreased the grain size and enhanced the formation of a passive layer resulting in superior corrosion resistance in aggressive H2O2 augmented saline solution [221]. Sputtered TiNxOy coatings with chemical composition ranging from TiN to TiO2 deposited on microroughened titanium plates showed a significantly high level of bioactivity as compared to bare Ti substrates (1.2 up to 1.4 fold increase in cell proliferation) that made them biocompatible over a broad range of compositions [222]. Similar results were reported for TiNxOy coatings deposited on roughened SS [223] and CoCr alloy during the first two weeks of healing [224]. Therefore, in addition to the enhanced wear resistance of TiNxOy coatings, they can “isolate” the substrate metal from the bone and accelerate the effect on the growth of bone cells. For that reason, our research group synthesized gradient TiN/TiO2 coatings by a cathodic arc (for TiN) and glow discharge deposition (for TiO2) on an initial surface treated by electron beam Ti6Al4V alloy [225]. The initial electron beam treatment (EBT) of the alloy not only roughened the surface of the alloy forming regular grooves and heights but also enhanced the hardness of the substrate, thus generating a smooth gradient in stiffness from the substrate to the coating. Because of the decreased grain sizes and increased number of defects in the substrate, the EBT lowers the heat conductivity (λ) of the surface. Therefore, due to trapping heat near the surface, this initial substrate treatment not only improved the adhesion of TiN/TiO2 coating to the Ti6Al4V alloy but also triggered reorientation in the micro-volumes of the nitride and rutile to anatase ratio of the oxide (Figure 10), thus decreasing its surface hardness and bringing it closer to that of trabecular bone and human teeth and decreasing the elastic modulus mismatch between the bone and implant.
We observed similar reorientation of the micro volumes of the nitride and oxide and a decrease in microhardness after EBT for the magnetron sputtered TiN/TiO2 coating deposited on both Ti5Al4V alloy [226] and Co-Cr alloy [227]. The TiN/TiO2 coatings displayed adhesion bonding between the nitride and oxide layers [228] with substantially improved tribological performance and corrosion resistance as opposed to the Ti-Al-V substrate. Compared to the magnetron sputtered TiN/TiO2 coatings, those deposited by cathodic arc deposition (CAD) and glow discharge oxidation showed better human osteoblast-like cell (MG63) adhesion, viability and bone mineralization activity on both polished and EBT Ti6Al4V samples (Figure 11). This is because, in contrast to the less rough magnetron sputtered (MS) TiN/TiO2 coatings, on the surface of CAD and glow discharge oxidized coatings there are many surface elements for focal adhesion in the coating, such as droplet phase particles, disoriented crystallographic planes at the tips of oxide crystals, defects such as pores, etc., all of which support cell movement and proliferation. Simultaneously, cells cultured on the grooved surface with a smaller channel spacing (AR500) tended to have a stronger orientation along the groove axis compared to the AR850 surface with greater groove spacing. The results show that the deposited TiN/TiO2 coating on the micro-rough EBM surface stimulates and accelerates cell differentiation [225].
Besides, magnetron-sputtered TiON and ZrON films on SS 316L substrates indicated a drastic reduction of bacterial adhesion (P. aeruginosa) as well as inhibition of biofilm formation at different time durations [229]. The bactericide activity of TiON and TiON-Ag sputtered films under visible light irradiation was reported by Rtimi et al. [230]. They stated that compared to TiON film which inactivated bacteria within 2 h, TiON-Ag coatings with Ag concentration below the cytotoxicity level showed faster and repetitive inactivation of E. coli. Similarly, ZrO2-Ag and ZrON-Ag coatings had lower bacterial retention while ZrON-Ag with porous structure and 11.8 at% Ag possessed the best antibacterial performance against S. aureus and A. actinomycetemcomitans together with excellent human gingival fibroblast (HGF) cell compatibility [231]. For magnetron-sputtered ZrON-Cu coatings, only the presence of CuO species caused bactericidal activity against S. epidermidis while Cu2+ ion release did not influence the antibacterial properties of the coating [232].

6.1.4. Carbon-Based Coatings

There are several types of carbon-based materials that are used for biomedical applications: (a) amorphous carbon nanostructures (diamond-like carbon (DLC), graphite-like carbon (GLC), pyrolytic carbon); (b) nanocrystalline diamond (NCD); (c) graphene and its derivates. Recent studies indicate that these carbon materials have exceptional biocompatibility, stability and mechanical properties [233,234]. Amorphous carbon with high sp3 content is referred to as diamond-like carbon (DLC) while higher sp2 content yields materials closer to graphite (graphite-like carbon). However, in contrast to graphite, the latter has higher hardness and high corrosion resistance. Pyrolytic carbon is also an amorphous carbon allotrope with dominating sp2 bonding. It is conventionally produced by CVD from gaseous hydrocarbon precursors. Depending on the process conditions, pyrolytic carbon coatings can have isotropic, granular, lamellar, columnar, etc., structures. Although mainly used for heart valve protection [235], pyrolytic coatings have been applied for the replacement of small joints such as knuckles, wrist joints and proximal interphalangeal joints [236].
DLC-based coatings are considered promising for bioimplant application because they have excellent mechanical properties, a low coefficient of friction and good wear resistance. For that reason, applied as coatings on Ti substrates by a CVD technique, the DLC film substantially improved the nano-hardness and tribological performance, decreasing the coefficient of friction by one order of magnitude [237,238]. The in vivo behavior of PVD-deposited DLC coatings on Ti substrates indicated no inflammatory reactions, confirming its good biocompatibility [239]. However, some disadvantages such as high internal stress, low toughness and high sensitivity to ambient conditions can be observed for a single layer of DLC coating that can explain the high revision rates of single-layered DLC-coated orthopedic joints [240]. Compared to TiN-coated joint prosthesis, DLC coatings demonstrated lower wear resistance [241]. To address this problem, multilayered coatings on Ti-6Al-4V alloy consisting of (a) alternating Zr and ZrN sublayers responsible for corrosion resistance and load carrying capacity, (b) overlaying Zr/DLC composite film for enhanced adhesion and reduced fatigue residual stresses and (c) top N-doped DLC to reduce friction and enhance, have been designed [242]. The resultant coatings showed a decreased coefficient of friction by more than 50% and two to three times increased hardness than that of bare Ti substrate. Except for a substantial decrease in wear, the middle layer improved the delamination strength, which is low in single DLC coatings. Additionally, fluorinated DLC coatings also exhibited good antibacterial properties against E. coli and S. aureus by decreasing their counts from 2.4 × 104 and 2.54 × 104 to less than 20, in contrast to two orders of magnitude growth of bacteria in the control groups [243]. At the same time, no substantial difference in cytotoxicity between the groups was observed confirming good biocompatibility of the coating. To improve the biocompatibility of magnetron-sputtered DLC coatings, Si-doping was also applied. Deposited on Ti6Al7Nb alloy, the addition of silicon up to 14–22 at% to the DLC coatings had a very positive effect on the proliferation and viability of endothelial cells [244]. Increasing the Si content resulted in a rise in the hydrophilic character of the coating, film hardness by up to 40% and reduced colonization by E. coli bacteria compared to the uncoated substrate [245]. Wachesk et al. deposited hybrid DLC coatings incorporating TiO2 nanoparticles by plasma-enhanced CVD on AISI 316 and implanted them in CF1 mice peritoneum [246]. The in vivo results showed that the presence of TiO2 nanoparticles enhanced healing activity and reduced the inflammatory reactions increasing DLC biocompatibility. However, a major concern with DLC coatings is their instability in an aqueous environment, which promotes delamination of the coatings [247].
Nanodiamonds possess a high surface-area-to-volume ratio together with good biocompatibility and bioactivity [248]. Additionally, diamonds were reported to have high wear resistance and low friction coefficient, which make them ideal for protective layers [249]. Nanodiamond coatings also show high surface roughness, hydrophobicity of the surface, high stability, superior electrochemical properties and biocompatibility [250,251]. On metallic substrates nanocrystalline diamond (NCD) coatings behaved as well-adhering and highly cohesive films [252]. It was found that the cell performance on NCD films depended on surface atoms or chemical groups [253]. For example, on micropatterned NCD films, human dental stem cells adhered and grew preferentially on O-terminated domains rather than on H-terminated areas [254]. The low boron (100 to 1000 ppm of B) doping of NCD films was also found to support cell proliferation and early osteogenic differentiation of MG63 cells because of the increased electroconductivity of the doped films [255]. A similar effect of enhanced attachment and spreading of MG63 cells was observed for composite apatite-nanodiamond coatings compared to pure SS and apatite coatings without nanodiamonds [256]. The authors explained the observed effect of the increased adsorption of fibronectin on the composite coatings. Simultaneously, Medina and co-authors observed that NCD coatings were able to establish a chemical bond with the cell wall or membrane of Gram-negative P. aeruginosa bacteria, thus hindering the bacterial adhesion and colonization of the surface [257].
Two-dimensional (2D) allotropes of carbon–graphene, a single atom thick layer of sp2 carbon and related graphene oxide (GO) and reduced graphene oxide (rGO) are innovative materials in the medical sector because of their unique biological properties. Graphene oxide is an oxygenated derivate of graphene with abundant functional groups on planes and edges allowing desirable dispersion behavior in aqueous media [107]. rGO consists of fewer oxygen-containing groups because of interactions with reducing agents [258]. The chemical properties of rGO resemble those of pristine graphene [259]. Many scientific works report the ability of rGO to promote osteogenic stem cell differentiation [260,261]. Graphene can be directly grown on metallic surfaces such as Ti6Al4V [262] and Mg [263] to improve bioactivity and corrosion resistance. It is also popular material with antibacterial, antifouling and hemo-compatible properties but the layered structure of graphene nanosheets limits its benefits and advantages [264]. For that reason, graphene nanoplatelets with improved biocompatibility and effectiveness for biomedical devices have been introduced [265]. In such a form, graphene is usually combined with natural or synthetic biopolymers to enhance the osteogenic potential and mechanical properties of the coating. For example, by using electrophoretic deposition, Suo et al. deposited GO/chitosan/HAp coatings on Ti that showed higher bonding strength to the substrate than HAp, GO/HAp and chitosan/HAp coatings and significantly enhanced cell–coating interactions in vitro and osseointegration in vivo [261,262,263,264,265,266]. Simultaneously, the fracture toughness of HAp rose by 200% by including only 1 wt% rGO [267]. Graphene-based materials have powerful antimicrobial properties and inhibit bacterial colonization. For example, Agarwalla and co-authors [268] tested graphene coatings on Ti against P. aerugimosa, E. faecalis, S. mutans and C. albicans and found that, when repeated twice, the film reduced biofilm formation due to the hydrophobicity of graphene. Similarly, functionalized GO nanocomposite with Ag NPs showed excellent antimicrobial activity against E. coli and S. aureus [269]. Despite these impressive properties, there is still a concern about the biodegradability of nanodiamonds and graphene in the organism. Additionally, in vitro studies with GO nanomaterials indicated the generation of ROS, DNA damage and mitochondrial disturbance [270].

6.1.5. Calcium Phosphates and Hydroxyapatite

Calcium phosphate ceramic coatings are extensively used to boost the biocompatibility of metal implants because of their superior adaptation to in vivo conditions. Bioactivity properties are varied according to the type of calcium phosphates. Both calcium phosphate types, HAp and tricalcium phosphate (TCP) have different crystallinity, stability, solubility, ion release and mechanical properties. The crystallinity is affected by the Ca/P ratio and a higher amount of Ca2+ or PO43- can trigger amorphous phase formation such as dicalcium phosphate dihydrate, CaHPO4.H2O and Ca3(PO4)2 [271]. In contrast to highly crystalline HAp, calcium phosphate-based coatings have high solubility in an aqueous medium that reduces coating stability and can cause implant loosening. Because of its similar properties to the inorganic composition of hard tissue such as bone and teeth [272], HAp (Ca10(PO4)6(OH)2, Ca/P = 1.67) is frequently employed as bioactive material. HAp has shown exceptional biocompatibility, osteo-inductivity, osteoconductive and bioactivity [273]. By releasing calcium and phosphate ions, HAp enhances bone regeneration and promotes mineralization [274]. By covering the metal biomaterial, HAp coating helps in maintaining stability fixed to the bone while minimizing the side effects of ion release in the bio-environment. The enhanced osteoconductive properties of HAp coatings can be attributed to the bone-like apatite chemistry of the coating (Na+, Mg2+, CO32−, Ca2+ and PO43−) and reduced degradation rate that allows a balance between ion release by the coating and ion absorption by the tissue during the bone formation [275].
However, due to their ceramic nature, highly porous or highly crystalline HAp coatings can show low mechanical properties (very brittle, with low flexibility), poor adhesion to the metal surfaces and low corrosion resistance, which makes them inappropriate for load-bearing applications. The difference in the thermal expansion coefficient of metallic alloys and HAp results in residual thermal stresses which can promote cracking or delamination of the coatings. The corrosion resistance also depends on the deposition method. Sankar et al. compared the corrosion behavior of HAp coatings obtained by EPD and pulse laser deposition (PLD) method and found that EPD coating had lower corrosion protection than PLD films due to the formation of denser and pore-free structures [276]. Dispersion strengthening by introducing a second phase to its microstructure such as other ceramics, carbon nanotubes or other compounds is deemed to overcome the poor mechanical properties of these coatings. For example, the addition of TiO2 to HAp (20–80 wt%) coating produced by High-Velocity Oxygen Fuel (HVOF) spraying on Ti6Al4V alloy delayed the HAp dissolution and increased the coating stability in Hank’s solution [277]. Similarly, by introducing TiO2 to fiber HAp by the EPD process it was discovered that the pores of the HAp coating produced from suspension with 50 and 75 wt% fiber HAp can be efficiently infiltrated and filled with TiO2 nanoparticles which increase the corrosion resistance of the coatings in SBF solution [278]. Evcin et al. [279] produced a series of HAp/Al2O3, HAp/B2O3 coatings on Ti6Al4V alloys by the HVOF method and found that increasing the amount of Al2O3 and B2O3 in HAp increased the adhesion strength and wettability.
The addition of specific trace elements to HAp like Zn, Mg, Cu, Si, Sr, Mn and F can also have a role in bone regeneration. For example, Zn was found to increase alkaline phosphatase activity and stimulate bone formation by osteoblasts [280]. For that reason, Zhou and co-authors produced Zn-doped HAp coating on ZK60 magnesium alloy by one-pot hydrothermal method with nano-whisker structure and showed that the film had a higher corrosion resistance compared to HAp coatings, promoting adhesion and differentiation of rat bone marrow mesenchymal stem cells at Zn concentration of 5% and obvious antibacterial activity [281]. Si-doped CaP was deposited on AZ31 magnesium alloy and the osteoblast cytocompatibility, evaluation showed that Si ion played a vital role in the nucleation and growth of apatite thus influencing the biological metabolism of osteoblast cells [282]. Pure Mg demonstrated an antimicrobial effect because of the increase in pH by degradation, while F is a basic element in bones. F can stimulate the differentiation of mesenchymal stem cells into osteoblasts, induce bone formation and promote the nucleation of HAp [283]. Comparing F-doped HAp, Mg-doped HAp and Mg/F-doped HAp coatings deposited on Ti substrate by pulsed laser deposition, it was found that Mg-F-HAp coating better promoted the transformation of apatite-like to HAp phase due to the synergistic effect of Mg and F. The porous 3D structure of the coatings enhanced the viability of rBMSCs, especially for Mg-F-HAp coatings, where a regulated biodegradable rate and good cellular proliferation were observed [284]. Sr ions were also found to increase bone-to-implant contact by osteoblastic cell proliferation, accelerate bone matrix synthesis and inhibit bone resorption [285]. Consequently, Sr-doped HAp coatings prepared on Mg-4Zn substrates by electrochemical deposition showed better corrosion resistance, improved protein adsorption and initial adhesion of mesenchymal stem cells, as well as improved osteogenic differentiation compared to HAp coatings [286]. Similarly, Ca-Sr-P coatings with dense crystalline structure deposited on biodegradable Mg alloy by chemical immersion method demonstrated improved corrosion resistance, higher bone formation and better osteointegration around the coating than the Mg alloy after 4 weeks of implantation in a rabbit model [287]. The biocompatible properties of HAp coatings were also enhanced by imparting antibacterial properties by incorporating silver. For example, in the presence of F in Ag-F-HAp coatings developed on Ti substrate by sol-gel method with a silver concentration of 0.3 wt%, the crystalline size and pores in the coating decreased whereas the antibacterial activity against E. coli bacteria increased with the amount of F [288].
The incorporation of polymers in HAp coating structures was also found to have a positive effect on the ceramic coating properties. For example, electro-phoretically deposited HAp-CaSiO3-chitosan composite coatings that were made porous by heat treatment at 700 °C in a controlled atmosphere indicated improved corrosion resistance and bioactivity in SBF compared to Ti substrate [289]. Similarly, biomimetically deposited Ce-doped HAp/collagen coatings on initially alkali-thermal oxidation pretreated Ti substrate showed good antibacterial activity against both E. coli and S. aureus, being more effective against E. coli [290]. By electrostatic interaction, the negatively charged surface absorbed positively charged collagen and negatively charge HAp that incorporated Ce ions in its lattice.
The general requirements for the properties of HAp coatings are listed in Table 2, although the relatively high thickness of HAp coatings, the ease of delamination of the film from the base metal and the possibility of the coating fracture occurrence can reduce the functional performance of the implant. The debris may cause inflammation in the host body.

6.1.6. Bioactive Glasses

The bioactive ceramic materials include silica, calcium, phosphorous and sodium ions (glass composition CaO-SiO2-P2O5-Na2O) that are released when the bioactive glass interacts with cells and the bio-environment, leading to fast bone growth. The 45S5 bio-glass (45 wt% SiO2, 24.5 wt% CaO, 24.5 wt% Na2O, 6 wt% P2O5) has shown the most effective bioactive properties, namely class A bioactivity, allowing it to bond to soft and hard tissues [292]. Similar to many other ceramic materials, bioactive glasses can be produced as particles with micron and nano-size, or fibers, 3D scaffolds, mesoporous coatings, or monoliths [293]. Compared to other biomaterials, bioactive glasses can make possible better integration between the metal implant and the growing tissue because of their significant bioactive behavior [276]. For example, when comparing the in vivo efficacy of CaO-MgO-SiO2-based bioactive glass-ceramic on Ti6Al4V alloy (deposited by atmospheric plasma spraying) with HAp-coated samples implanted in New Zealand rabbits, the significant growth of new bone confirmed the superior biological activity of bio-glass coatings in treating load-bearing bone defects [294].
Although showing excellent bioactivity, because of their semicrystalline or amorphous structure, bio-glasses can exhibit poor mechanical strength, low tensile strength, fatigue resistance, elastic modulus and corrosion resistance especially as regards porous coatings. The porosity formation of coatings obtained by plasma spraying occurred due to the evaporation of volatile Na2O and P2O5 [295]. To overcome this disadvantage, composites with metal oxides such as ZrO2, TiO2, Al2O3 or graphene and its derivates were made [296]. These could improve the thermal, electrical and strength properties of bioactive glasses [297]. Additionally, except for excellent bioactivity, such bioactive glass composite coatings exhibit improved antibacterial activity, angiogenic properties and corrosion resistance [298]. For example, one-dimensional bioactive glass nanorods of 45S5 composition produced by sol-gel process and hybridized with reduced graphene oxide sheets (rGO), following different methods for developing composites such as constant stirring, sonification and simultaneous reduction in GO–bio-glass composite, showed better results in bioactivity, hemocompatibility, cell proliferation and antibacterial activity as compared to pure bioactive glass nanorods [299]. Similarly, electro-phoretically deposited bioactive glass-rGO hybrid thin films (2 μm thickness) deposited on TiO2 nanotubes with a diameter of around 100 nm were advantageous in antibacterial activity, hemocompatibility and MG-63 cell proliferation [300].
Bioglass composites with additions of metals, metal oxides and HAp also demonstrated promising bioactive properties. For example, laser process bio-glass coatings reinforced with Ti on Ti substrate with excellent coating interfacial characteristics, improved hardness, corrosion protection and in vitro wear resistance, also indicating better cell-material interaction than bare -Ti [301]. However, when comparing HAp-based HAp/Ag coatings with bio-glass-Ag bio-composite coatings on NiTi alloy, both deposited by sol-gel method, higher corrosion resistance and adhesion strength were found for HAp/Ag coatings [302]. Often, there are challenges in creating a good adhesion between the glass topcoat and the metal substrate due to low metal–amorphous ceramic interface bonding and the formation of cracks [303]. A solution to the problem can be the utilization of a polymeric matrix to create a nano-composite coating. These are characterized by low processing temperature and elimination of the sintering process if required. Among various polymers, chitosan, a natural polymer, is often used because of its biodegradability, biocompatibility, osteo=conductivity and antimicrobial properties [304]. Recently, Alaei and co-authors produced chitosan-bioactive glass nanocomposite coatings that provided significant corrosion protection to Mg alloy and controlled biological properties [305]. A new family of chitosan-based composite coatings incorporating HAp–bio-glass and different concentrations of Fe2O3 particles was electro-phoretically deposited on Ti13Nb13Zr alloy [306]. All Fe2O3-containing coating formulations showed favorable hemocompatibility, better surface properties, improved corrosion resistance and better cytocompatibility with MG63 cells as opposed to bare alloy and Hap–bio-glass coatings. Similarly, Mn-modified bio-glass/alginate nanostructure composites deposited on SS 316L by electrophoretic deposition demonstrated that the increased manganese in bio-glass had a positive effect on corrosion resistance in SBF and improved bioactivity [307].

6.2. Organic Coatings

Recently, the interest in applying polymer materials as coatings has increased substantially because of their easy fabrication, affordable price, low toxicity, corrosion resistance and eco-friendly nature. Polymers show low strength and elastic moduli as compared to ceramics and metals and are not used for load-bearing applications. They can be both non-biodegradable and biodegradable with complete degradation over time. However, except for low mechanical properties, another issue faced by polymers is their inadequate degradation rate and inflammatory reaction which limitations prevent them from being widely used as biomaterials for hard tissue coatings [57].

6.2.1. Synthetic Polymers

The most commonly used synthetic polymers for periosteum development are polylactic L-lactic-co-glycolic acid (PLGA), polyurethanes (PU), polyethylene glycol, (PEG), polycaprolactone (PCL) and poly L-lactic acid (PLLA), and polymethylmethacrylate (PMMA) [308]. Synthetic polymers are usually hydrophobic and possess no antibacterial activity. They can also deteriorate the adhesion of bone cells [309] restricting their widespread application in the medical sector. To improve the biological performance of polymer coatings, composite systems based on biocompatible polymers modified with various compounds or particles are often used [310]. For example, combining PU, which is usually used in medicine because of its favorable mechanical properties and high biocompatibility, with 0.25 wt% graphene (used as an antibacterial agent) and 2 wt% β-TCP (as a bioactive component) in dip coatings on Ti implants gave positive cell response in normal human osteoblast (NHOst) cells and effective antibacterial activity in contrast to the other examined composites with higher graphene content [309].
Similarly, PMMA is characterized by high thermal and chemical stability, biocompatibility and advanced mechanical properties [311]. Extensive studies on PMMA composites containing HAp, metal oxides and bio-glass showed that they are attractive for surface modification of biomedical implants because of their high biocompatibility, bioactivity and antimicrobial properties [312]. High molecular PMMA composite coatings with TiO2, Al2O3, HAp, bio-glass and Hap–bio-glass also provided enhanced corrosion protection compared to pure PMMA coatings [313].
Conductive polymer coatings have been also used as coatings on hard implants. Among them, poly-pyrrole (PPy) which has good biocompatibility, is frequently examined for biomedical applications [314]. However, once fabricated, pristine PPy has a rigid, brittle and insoluble nature [315]. To overcome this shortcoming, composites with various additives have been developed. For example, a composite coating of PPy with ZnO was developed by Guo et al. to protect the biodegradable Mg alloys from fast decomposition and to impart cyto-compatible and antibacterial properties [316]. Multifunctional composite coatings of PPy with pectin and 10 wt% gentamicin deposited on TiNbZr substrate demonstrated effective antibacterial performance, lower corrosion rate, controlled degradation because of the slow release of gentamicin and improved biocompatibility [317].
A synthetic polymer matrix was also used to augment HAp and bio-glass coatings and improve their mechanical strength. For example, nanostructured HAp was incorporated in polyetheretherketone (PEEK) to form PEEK–HAp composite coating which was deposited by EPD on 316 SS and heat treated at 375 °C to densify the coating and enhance the adhesion to the substrate [318]. In contrast to the as-deposited film where the HAp covered the PEEK and stimulated bioactivity, after heat treatment the HAp became encapsulated in PEEK and reduced bioactivity. Both adhesion strength and bioactivity were dependent on PEEK/HAp ratio. The increased amount of HAp caused improved bioactivity and reduced adhesion strength. Biodegradable PCL coatings on 316L SS containing 10 wt% gelatin (GE) and 3 wt% bio-glass showed drastically improved corrosion resistance and significant apatite formation as opposed to only PCL/GE coatings [319]. The bio-glass-containing composites also revealed increased MG63 cell viability compared to PCL/GE coatings while the results in an animal model (New Zealand white rabbits) demonstrated no inflammation and granulation, endothelial swelling, fibrotic tissue or other toxic effects.
More merits can be offered by biodegradable and resorbable polymers than non-degradable ones in terms of low levels of possible infections and implant rejection. However, the degradation of polyesters such as polylactic acid (PLA) and polyglycolic acid (PGA) and their co-polymers is known to create an acidic environment. The process can trigger host tissue response and foreign body reactions during degradation, as well as moderate cytotoxic reactions [320]. For that reason, cationic polymers with low toxicity such as poly(glycidyl methacrylate) or PGMA can be used. When Ti implants were functionalized with PGMA coupled with quaternized polyethyleneimine (bactericidal agent) and alendronate with high affinity to bone minerals, the obtained coating inhibited bacterial infections and promoted osseointegration in the late stages [321]. However, synthetic polymers do not have signaling sequences that are naturally present in biological polymers such as collagen, fibrinogen or fibronectin.

6.2.2. Polymeric Gels (Natural Polymers)

Various natural polymers such as chitosan, silk fibroin, collagen, etc., have been used in the production of bioactive coatings. Chitosan, representing the de-acetylated derivative of chitin, is considered to have osteoconductive properties. Coatings of chitosan/heparinized GO deposited by layer-by-layer technique on Mg alloy showed that substantial endothelial cell adhesion and proliferation were promoted [322]. Chitosan-Mg composite dip-coatings on Mg-Gd alloys also showed a higher amount of newly formed bone in rabbits [323]. Similarly, improved cell adhesion and proliferation of osteoblasts were observed on coated AZ31D alloy with bioactive carboxymethyl chitosan by immersion treatment [324]. As cationic macromolecule, chitosan can bind to the negatively charged cell membrane of bacteria and display, albeit weak, antibacterial properties. To impart a stronger bactericidal effect of the chitosan-containing coatings, films of chitosan and hyaluronic acid (HA) on rough Ti substrate were designed to release β-amino acid-based peptidomimetic antimicrobial peptide [325]. The layer-by-layer prepared coating showed a strong chemical cross-linking of chitosan with HA films which caused prolonged β-peptide retention that selectively prevented S. aureus biofilm formation for up to 24 days and remained its bactericidal properties after being challenged sequentially five times with S. aureus inoculum over 18 days. Simultaneously, no significant cytotoxicity on osteoblast precursor cell line derived from mouse (MC3T3-E1) compared to uncoated and film-coated controls without β-peptide was observed. Such a novel localized delivery approach that can maintain long-term antibacterial properties is promising for the development of coated medical devices prone to biofilm-associated infections. However, the adhesion and durability of chitosan coatings might raise some concerns [326].
Polydopamine, the final oxidation product of dopamine or other catecholamines, was found to form layers with an adjustable thickness (from a few to about 100 nm [327]) with good adhesion and high cell affinity [328]. This fact was confirmed by the study of Peng and co-authors who demonstrated an enhanced osteogenic differentiation on Zn-containing polydopamine films on AZ31 magnesium alloy together with improved osteogenesis and osteointegration in Sprague-Dawley rats after 8 weeks post-implantation [329]. A hybrid coating consisting of hydrothermally grown ZnO nanorods on Ti modified with polydopamine and covalently immobilized Arg-Gly-Aspartic acid-Cys (RGDC) peptide promoted cytocompatibility, new bone tissue formation and osteointegration between the implant and the new bone even in the presence of injected bacteria, or demonstrated simultaneous osseointegration and infection prevention [330]. Polymer coating produced via reversible addition-fragmentation chain transfer polymerization from glutamic acid and dopamine metha-crylamide was immobilized on Ti substrate by catechol pendants on the polymer chain [331]. Besides promoting mineral deposition, the coating was found to promote osteoblast adhesion and proliferation. Dopamine-silver loaded coating prepared at different pH values (4, 7 and 10) and different Ag+ concentrations (0.01 and 0.1 mg/mL) showed that the pH10/0.1 group displayed osteogenesis in the bacterial environment due to the great antibacterial properties and promoted mineralization activity [332]. To reduce the well-known cytotoxicity of Ag, Guo et al. prepared Poly-L-lysin (PLL)/sodium alginate/PLL self-assembled coating loaded with nano-silver on Ti that effectively inhibited the adhesion of bacteria [333]. At the same time, the PLL/SA/PLL coating induced mineralization in SBF and improved cytocompatibility and reduced cytotoxicity. Similarly, by double chelation of dopamine and chitosan, a hybrid coating consisting of HAp/dopamine/chitosan and nano-silver achieved a long-term release of silver and a continuous bacteriostatic effect [334]. This effect was accompanied by substantial osteogenic potential demonstrated in both in vitro and in vivo tests.
Silk fibroin consists of light and heavy chains and hydrophobically linked glycoprotein P25 that are all crosslinked to form a complex with antiparallel beta-sheets [335]. Because of the formation of β-sheets, silk fibroin scaffolds have better mechanical properties than collagen and chitosan but they are still insufficient compared with bone tissue [329,330,331,332,333,334,335,336]. In the form of hybrid coatings on WE43 magnesium alloy consisting of an inner layer of Mg(OH)2 produced by anodization, a middle layer of HAp formed by EDP and an outer silk fibroin layer deposited by spin coating, the surface modification was shown to improve not only corrosion resistance but also cell attachment, viability and proliferation [337]. To increase the osteogenic capacity and mechanical properties of silk fibroin, besides different organic and inorganic components, surface modifications by bioactive moieties that form hybrid films can be applied. For example, blends of silk fibroin/chitosan/rGO were fabricated by solvent casting method as films whose hydrophilicity, swelling and degradability decreased with increasing silk fibroin content, whereas the tensile strength increased [338]. The cell behavior of the G-292 cell demonstrated promoted osteogenic performance by increasing chitosan content while the increase in rGO reduced the porosity and tensile strength. The optimum result corresponded to SF:CS:rGO equal to 84:7:9 weight ratio.

6.2.3. ECM Proteins/Cell Coatings

Synthetic or natural multifunctional peptides can be used as coating materials on metallic grafts because introducing organic molecules that contain functional fragments can stimulate the interaction with proteins of the extracellular matrix (ECM). The organic part of ECM consists of collagen type I fibrils embedded in the amorphous substance of glycosaminoglycans and different bone proteins. Since ECM components actively participate in the regulation of cellular processes and interactions, the modification of the implant surface with components of the ECM is an attractive approach. Collagen is known to enhance tissue regeneration of bone, tendon, ligaments and vascular and connective tissue [339]. Collagen type I coating extracted from rat tail and deposited on Mg-Zr-Ca alloy implants by dip-coating showed accelerated protein bonding capacity resulting in better osteoblast activity and a tendency to form superior trabecular bone structure in male New Zealand white rabbits compared to the uncoated samples in a shorter period of implantation [340]. Another derivate of extracellular matrix proteins promoting cell adhesion as integrin ligand [341] is RGD. By using high-affinity inorganic peptides such as TiBP that contain a Ti-binding domain, RGD can be combined with antibacterial peptides to form durable and stable coatings with both bone-promoting and antibacterial properties [342]. Similarly, Zhang et al. used TiBP to connect antimicrobial sequence from human β-defebsin-3 and RGD in a coating that was found to significantly reduce the bacterial colonization onto the Ti surface and better-supported MC3T3-E1 cell growth compared with PBS-treated Ti samples [343].
Not only proteins but also vesicles and cells can be immobilized on the surface of metallic implants to form biogenic coatings. For example, using secreted extracellular vesicles (EVs) by mesenchymal stem cells, it was found that tissue repair and regeneration can be promoted since their membranes contained signaling molecules and, additionally, EVs can carry and transfer different cargos [5]. For that reason, Chen et al. immobilized adipose-derived stem cell extracellular vesicles with physisorbed fibronectin onto the Ti surface and observed enhanced osteoblast compatibility and osteo-induction activity [344]. Another approach included the immobilization of Lactobacillus casei on the surface of heat-treated Ti to form a probiotic coating [345]. The polysaccharides in the film promoted osteogenic differentiation through immunoregulation of macrophages that secreted osteogenic factors, while the surface showed 99.98% antibacterial effectiveness against S. aureus. However, the limitations of all these organic coatings are related to low mechanical strength, difficult sterilization and rapid degradation that can be overcome by designing composites with bio-ceramics or strong materials such as synthetic polymers or metals [78].
Some benefits and shortcomings of the main types of materials used for the construction of bioactive coatings are summarized in Table 3.

6.3. Active Moiety-Containing Coatings

Cellular active substances such as growth factors, chemokines, drugs, etc., and their application on the surface of orthopedic implants are intensively examined because these substances can effectively improve surface biocompatibility and promote osseointegration. Depending on the active moieties, these coatings can be divided into (a) drug-containing; (b) osteogenic-factor-containing; (c) immunomodulatory factors-containing and (d) antibacterial films.

6.3.1. Drug-Containing Coatings

Osteoinductive drugs that accelerate bone formation and enhance implant fixation are zoledronic acid and simvastatin. They can be both loaded onto the surface of the coating or within the film. For instance, zoledronate is a long-acting bisphosphonate that was found to cause cytoskeletal alterations in osteoclasts which decreased their activity and triggered apoptosis [347]. Bilayer coating of zoledronic acid associated with CaP on Mg-Sr alloy enhanced the proliferation, osteogenic differentiation and mineralization of pre-osteoblast MC3T3-E1 cells but also inhibited osteoclast differentiation and induced apoptosis which balanced the bone remodeling process [348]. Because zoledronic acid shows numerous side effects such as osteonecrosis of the jaw, gastrointestinal irritation and impairment of renal function during systemic use [349], the administration of the drug in a controlled manner by a coating seems to have potential effectiveness.
Simvastatin is a molecular analog of HMG-CoA (3-hydroxy-3-methyl-glytaryl-coenzyme A) that was found to promote mesenchymal cell differentiation into osteoblast, downregulating osteoblast apoptosis and upregulating BMP-2 [350]. Electrohoretically deposited coatings consisting of simvastatin/gelatin nanospheres/chitosan composite on WE43 magnesium alloy were found to enhance the degradation resistance of the alloy substrate and simultaneously promoted osteogenic activity [351]. However, studies with rats proved that a high dosage of simvastatin (0.5–2.2 mg per site) may induce inflammation or even impair bone healing [352,353]. Therefore, controlled delivery and drug release in an appropriate dose are of prime importance.

6.3.2. Coatings Containing Osteogenic Factors

Hormones, cytokines and growth factors such as bone morphogenetic proteins (BMPs), transforming growth factor-beta (TGF-β), vascular endothelial growth factor (VEGF) and nerve growth factor (NGF) are known to play an essential role in bone repair and are also used to construct bioactive surfaces. BMPs are widely used cytokines to confer osteo-inductivity, but their burst release can decrease the osteogenic effect [354]. That is why porous coating are expected to be suitable for this purpose. Teng and co-authors prepared porous structured coating by 3D printing and MAO and grafted BMPs onto the surface [355]. The release of BMP-2 sustained for more than 35 days and stimulated osseointegration between the implant and bone. Kim et al. loaded BMP-2 at different concentrations in a MgO and Mg(OH)2 layer produced by micro-arc coating on AZ31B magnesium alloy and found substantial proliferation and differentiation of osteoblast cells when BMP-2 was released continuously in a concentration of 50 ng/mL after four weeks, thus stimulating stable bone growth and bone formation [356].
When incorporating BMP-9 in thermosensitive collagen and depositing it onto porous Ti, Zhu et al. found that the thermosensitive collagen degraded slowly at 37 °C thus ensuring temperature-controlled sustained release and enhanced osteogenesis around the implant [357]. Similarly, powder-processed dopamine/gelatin/rhBMP-2 coated β-TCP films on Mg-Zn alloy facilitated cell proliferation and significantly enhanced the osteogenic differentiation of Sprague-Dawley rat bone marrow-derived mesenchymal stem cells in vitro [358]. The in vivo results in New Zealand rabbits showed strong stimulation of new bone formation and matched composite degradation and bone regeneration rate. Another coating strategy accounted for the polydopamine-mediated assembly of HAp-coated alkaline treated nanoparticles and immersion of BMP-2 onto the surface of AZ31 magnesium alloy, where the coated sample showed substantial BMSCs adhesion and proliferation and stimulated osteo-inductivity and osseointegration in the New Zealand rabbit model [359]. However, some clinical and pre-clinical side effects of BMP-2 include inflammatory and wound complications, ectopic bone, osteoclast activation and osteolysis, radiculopathy and urogenital events [360]. Therefore, more research is required to understand the long-term results and bio-functionality of conjugated coatings with BMPs.

6.3.3. Immunomodulatory Factors Containing Coatings

Since the implant material is a foreign body, a series of immune responses can occur mainly from macrophage activation. Simultaneously, the formation of wear particles (debris) can also aggravate inflammatory reactions and dynamic imbalance between osteoblasts and osteoclasts which can trigger bone resorption and implant loosening [361]. Therefore, adapting the immunoreaction by incorporating immune factors to regulate immune response can have a beneficial effect on osseointegration. For example, Li et al. used spraying to deposit GO on Ti and loaded it with interleukin 4 (IL 4) [362]. During acute inflammation, IL-4 from the coating induced macrophage polarization to the Type 2 phenotype that is known to inhibit the development of inflammation. Besides weakened inflammatory response, the film also promoted osteogenesis.
Another strategy is based on immune regulation of the balance between osteoclasts and osteoblasts. Taking this into consideration, Lui et al. [363] conjugated osteogenic growth peptide (OGP) with N-acetylcysteine (NAC) and functionalized Ti substrates to examine bone metabolism balance in vitro. Their studies on RAW 264.7 cells demonstrated that the peptide-modified surfaces inhibited the cells from secreting inflammatory cytokines (IL-1β and TNF-α) and suppressed important transcription factors for osteo-clastogenesis. Simultaneously, the modified surfaces stimulated osteoblast spreading, proliferation and differentiation.

6.3.4. Antibacterial Coatings

The ideal antibacterial coating should kill pathogens during the primary contact, thus preventing biofilm formation. Since implants exist in the organism for a long time, the implant surface should also provide antibacterial properties against late infections. Except for adding different metal ions such as Ag+, Cu2+ and Zn2+ with broad antimicrobial effects as previously discussed, some non-metallic compounds and biomolecules are also prominent candidates for the production of bactericidal coatings. For instance, iodine is found to have wide antibacterial activity without developing drug resistance [364]. Kato and Shirai deposited anodic oxide film on Ti substrate and ionized iodine was electrodeposited within the pores to achieve iodine content of 0, 20, 50, 60 and 100%, where 100% corresponded to 13 μg/cm2 [365]. In vitro and in vivo experiments showed a temporal pattern of rapid initial release and subsequently slow attenuation of iodine with approximately 30% of initial iodine content remaining after 1 year. Implants with iodine contents of >20% demonstrated sufficient antibacterial activity to prevent implant-related infections even after 1 year of implantation. Similarly, chlorhexidine can be absorbed on the bacteria’s surface and destroy the membrane permeability [366]. Micro- and nano-porous Ti surface prepared by alkaline and heat treatment and covalently conjugated with amino-silane was used to graft chlorhexidine via glutaraldehyde [367]. The surface containing 1 mg/mL chlorhexidine indicated the best antibacterial results together with good osteoblast compatibility. Even in the presence of bacteria, the surface displayed great potential for osteoblast adhesion at the implant-bone interface. Another antibacterial agent—dimethyl-amino-dodecyl methacrylate (DMADDM)—introduced in HAp-modified surface via polydopamine was gradually released during the first 4 weeks after implantation and exhibited both inhibition of the pathogenic bacteria growth and osteogenic differentiation [368].
The use of natural antimicrobial peptides (AMPs) represents another strategy for imparting the bactericidal properties of coatings. For example, a hybrid antibiofilm coating of immobilized antimicrobial peptide (D-GL13K) by functional linkers—elastin-like recombinamers (ELRs—was applied on a titanium surface [369]. The presence of AMPs in the hybrid coatings provided strong antibiofilm activity against mono-species and microcosm biofilm models together with excellent cytocompatibility towards primary gingival fibroblasts. Another coating consisting of polydopamine, cationic antimicrobial peptide LL-37 and phospholipid (1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine or POPC) was deposited on MAO-modified titanium substrates [370]. The multilayered coating was found to alleviate the burst release of LL-37 in the initial phase leading to antibacterial activity against S. aureus and E. coli. LL-37 killed bacteria by blocking the expression of bacterial-related genes and stimulating immune response under controlled release of POPC. However, certain dose-dependent cytotoxicity of antimicrobial agents remains a concern.
Another recent strategy is based on photothermal therapy with near-infrared (NIR) irradiation that allows deeper tissue penetration than ultraviolet (UV) light and high selectivity. Because of the local warming effect, the biofilm or bacterial integrity can be destroyed but due to low selectivity heating may cause adverse side effects to the surrounding tissue [371]. By forming a multifunctional coating on the Ti surface of Indocyanine green (ICG) and mesoporous polydopamine Yang et al. were able to convert NIR light energy into heat to kill bacteria, while simultaneously ICG produced ROS to destroy bacteria cell walls (Figure 12) [372]. Moreover, the mesoporous polydopamine was functionalized with RGD peptide to endow the coated Ti with good cytocompatibility. After biofilm eradication, the coating still displayed osteogenetic and osteointegration potential. This strategy has the potential for remotely controlled eradication in vivo avoiding invasive treatment without side effects on the surrounding tissue. Song and co-authors modified TiO2 nanorods on titanium surfaces with dopamine and ferrocene (PDA-Fc) to obtain efficient antibacterial surfaces [373]. Because of ROS generation by PDA-Fc redox reactions and local temperature increase by photothermal transformation of ferrocene, synergetic and more efficient bactericidal activity of the coating was observed.
To realize an intelligent release of antibacterial moiety when the microenvironment changes, Sang et al. coated silk protein coating with gentamicin on the Ti surface [374]. The coating exhibited a faster gentamicin release rate in an acidic environment (characteristic for the first weeks after implantation) than in alkaline media. Another mechanism relies on controlled release based on the heating effect during infection. This possibility is shown by Li et al. who produced thermosensitive chitosan-glycerin-hydroxypropyl methylcellulose hydrogel (CGHH) that layered the top of simvastatin-loaded TiO2 nanotubes [218] and was discussed in Section 6.1.2.
Some approaches rely on the modification of the microenvironment to achieve osseointegration and bactericidal properties. For example, zeolitic imidazolate frameworks-67 (ZIF-67) coating loaded with osteogenic growth peptide prepared on TiO2 nanotubes was found to rapidly dissolve under an acidic environment, as during inflammation [375]. The hydrolysis of ZIF-67 nanoparticles released Co ions and formed an alkaline microenvironment that effectively kills E. coli, S. aureus, S. mutans and methicillin-resistant S. aureus. The coating was able to suppress the inflammatory response and simultaneously improved the mesenchymal stromal cell (MSCs) differentiation under an inflammatory environment. In vivo results also pointed to a strong antimicrobial and anti-inflammatory properties of the coated implants at the early stages of implantation and enhancement of bone-implant osteointegration at the late stage.
Table 4 reveals the typical production techniques used for the deposition of nitride, oxide, oxynitride, carbon-based, calcium phosphates and hydroxyapatite, bioactive glass, synthetic polymer and natural polymer bioactive coating materials on metallic implants.

7. Prospect and Challenges of Bioactive Coating Systems

The main issues related to the biological performance of metal implant materials are related to their poor biostability and bioactivity, low resistance to corrosion and ion release and inadequate mechanical properties. Malfunctioning of conventional metal biomaterials is frequently attributed to inflammation due to corrosion, debris formation, or microbial infections. Various deposition techniques have been used for the successful fabrication of multifunctional mechanically- and corrosion-resistant biocompatible coatings for metallic implants to enhance cell-implant interactions and reduce infections, inflammation and other post-operative complications. Chemical and physical techniques are usually used to modify metallic surfaces. Because of their numerous advantages, sol-gel deposition, plasma spraying, PVD and CVD coatings have been extensively studied while layer-by-layer and 3D printing films can achieve favorable controlled drug release profiles. However, the coating stability can vary dramatically depending on the production route and materials chosen. Lattice mismatch and high residual stresses enhance the degradation of coatings in the human body and result in poor adhesion to the substrate. Such knowledge and development of mathematical models for the calculation of degradation rate and/or adhesion are crucial for reducing the revision surgery that will lead to economic benefits.
The focus of future works should be towards an exploration of the tailored drug release properties of coatings in view of the release, controlled by different external and internal factors, of bioactive compounds. New methods to impart positive characteristics including self-healing, anti-inflammatory, antimicrobial, or controlled drug-release properties should be evaluated to improve the efficacy of biomedical coatings and metal implants. Although extensively studied recently, smart-release coatings still fail to demonstrate sufficient osteogenic capacity. Future surface modification techniques should form advanced multifunctional (anti-inflammatory, antibacterial, osteogenic) coatings integrating the merits of several coatings and offering beneficial approaches in one system that can also modify the microenvironment to modulate the immune response to promote osteogenesis. The key future development in hard tissue implants should aim at the fabrication of personalized implants where, by adapting the coating composition, including a selection of drugs, dosage and their release profiles, all individual needs of the patient can be faced. This insight will be crucial in manufacturing new implant coatings that will resolve the current issues. Therefore, more research is needed to evolve innovative surface modifications that will impart controlled antimicrobial activity, drug release, biocompatibility and wear and corrosion resistance. Further study of a combination of (bio)chemical, (bio)physical, or (bio)mechanical stimuli in one coating system can be beneficial to accelerate bone regeneration. It can be expected that the continuous development of multifunctional bioactive coatings will bring a revolutionary breakthrough in orthopedic implantable devices.

8. Conclusions

Applying coating on metallic implant materials can give competitive advantages because of their functionality, stability, durability and biocompatibility. However, those presenting the highest bioactivity usually suffer from inadequate mechanical or tribological properties or vice versa. For that reason, their properties are often improved by designing various composites. Combining the advantages of different production methods and coating materials led to better clinical results in long-term use compared to bare metal materials. Additionally, the unique functionalities of multi-layered, biomimetically structured or 3D printed coatings open a new horizon into the design of multifunctional coated implants. Therapeutic coatings with multi-beneficial effects such as controlled delivery of ions, drugs, proteins, growth factors, etc., and osteo-inductivity rely on combined or new methods for the production of highly efficient biomaterials and orthopedic implants. These methods offer precise control over the coating composition and structure, thus yielding the desired properties and providing all needs as per implant requirements. Thus, these coated implants lead to better clinical success rate and long-term use in contrast to uncoated metallic implants.

Author Contributions

Conceptualization, M.P.N. and M.D.A.; writing—original draft preparation, M.P.N.; writing—review and editing, M.P.N. and M.D.A.; visualization, M.P.N.; supervision, M.D.A.; project administration, M.P.N. and M.D.A.; funding acquisition, M.P.N. and M.D.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Bulgarian National Science Fund, grant number KP-06-H67-5 (2022) “New biofunctional Cu-containing TiO2 coatings obtained by electrophysical methods to meet modern implantology needs”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rodriguez-Gabella, T.; Voisine, P.; Puri, R.; Pibarot, P.; Rodés-Cabau, J. Aortic Bioprosthetic Valve Durability Incidence, Mechanisms, Predictors, and Management of Surgical and Transcatheter Valve Degeneration. J. Am. Coll. Cardiol. 2017, 70, 1013–1028. [Google Scholar] [CrossRef] [PubMed]
  2. Goharian, A. Fundamentals in Loosening and Osseointegration of Orthopedic Implants. In Osseointegration of Orthopaedic Implants; Academic Press: Cambridge, MA, USA, 2019; pp. 1–26. [Google Scholar] [CrossRef]
  3. Delanois, R.E.; Mistry, J.B.; Gwam, C.U.; Mohamed, N.S.; Choksi, U.S.; Mont, M.A. Current Epidemiology of Revision Total Knee Arthroplasty in the United States. J. Arthroplasty 2017, 32, 2663–2668. [Google Scholar] [CrossRef] [PubMed]
  4. Saleh, M.M.; Touny, A.; Al-Omair, M.A.; Saleh, M. Biodegradable/biocompatible coated metal implants for orthopedic applications. Bio-Med. Mater. Eng. 2016, 27, 87–99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Lu, X.; Wu, Z.; Xu, K.; Wang, X.; Wang, S.; Qiu, H.; Li, X.; Chen, J. Multifunctional Coatings of Titanium Implants Toward Promoting Osseointegration and Preventing Infection: Recent Developments. Front. Bioeng. Biotechnol. 2021, 9, 783816. [Google Scholar] [CrossRef]
  6. Gulati, K.; Maher, S.; Findlay, D.M.; Losic, D. Titania nanotubes for orchestrating osteogenesis at the bone–implant interface. Nanomedicine 2016, 11, 1847–1864. [Google Scholar] [CrossRef] [Green Version]
  7. Kazimierczak, P.; Przekora, A. Osteoconductive and Osteoinductive Surface Modifications of Biomaterials for Bone Regeneration: A Concise Review. Coatings 2020, 10, 971. [Google Scholar] [CrossRef]
  8. Burr, D.B. Bone morphology and organization. In Basic and Applied Bone Biology; Elsevier: Amsterdam, The Netherlands, 2019; pp. 3–26. ISBN 9780123914590. [Google Scholar]
  9. Parfitt, A.M. Chapter 36—Skeletal Heterogeneity and the Purposes of Bone Remodeling: Implications for the Understanding of Osteoporosis. In Osteoporosis; Academic Press: Cambridge, MA, USA, 2013; pp. 855–872. [Google Scholar] [CrossRef]
  10. Rosa, N.; Moura, M.F.S.F.; Olhero, S.; Simoes, R.; Magalhães, F.D.; Marques, A.T.; Ferreira, J.P.S.; Reis, A.R.; Carvalho, M.; Parente, M. Bone: An Outstanding Composite Material. Appl. Sci. 2022, 12, 3381. [Google Scholar] [CrossRef]
  11. Ohata, Y.; Ozono, K. [Bone and Stem Cells. The mechanism of osteogenic differentiation from mesenchymal stem cell]. Clin. Calcium 2014, 24, 501–508. [Google Scholar]
  12. Globus, R.; Doty, S.; Lull, J.; Holmuhamedov, E.; Humphries, M.; Damsky, C. Fibronectin is a survival factor for differentiated osteoblasts. J. Cell Sci. 1998, 111, 1385–1393. [Google Scholar] [CrossRef]
  13. Matsushita, K.; Nakahara, C.; Kimura, S.; Sakamoto, N.; Ii, S.; Miyoshi, H. Intranuclear mesoscale viscoelastic changes during osteoblastic differentiation of human mesenchymal stem cells. FASEB J. 2021, 35, e22071. [Google Scholar] [CrossRef]
  14. Anderson, H.C. Molecular biology of matrix vesicles. Clin. Orthop. Relat. Res. 1995, 314, 266–280. [Google Scholar] [CrossRef]
  15. Hasegawa, T.; Hongo, H.; Yamamoto, T.; Abe, M.; Yoshino, H.; Haraguchi-Kitakamae, M.; Ishizu, H.; Shimizu, T.; Iwasaki, N.; Amizuka, N. Matrix Vesicle-Mediated Mineralization and Osteocytic Regulation of Bone Mineralization. Int. J. Mol. Sci. 2022, 23, 9941. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, K.; Ren, Y.; Lin, S.; Jing, Y.; Ma, C.; Wang, J.; Yuan, X.B.; Han, X.; Zhao, H.; Wang, Z.; et al. Osteocytes but not osteoblasts directly build mineralized bone structures. Int. J. Biol. Sci. 2021, 17, 2430–2448. [Google Scholar] [CrossRef] [PubMed]
  17. Maes, C.; Kobayashi, T.; Selig, M.K.; Torrekens, S.; Roth, S.I.; Mackem, S.; Carmeliet, G.; Kronenberg, H.M. Osteoblast Precursors, but Not Mature Osteoblasts, Move into Developing and Fractured Bones along with Invading Blood Vessels. Dev. Cell 2010, 19, 329–344. [Google Scholar] [CrossRef] [Green Version]
  18. Qin, L.; Liu, W.; Cao, H.; Xiao, G. Molecular mechanosensors in osteocytes. Bone Res. 2020, 8, 1–24. [Google Scholar] [CrossRef]
  19. Klein-Nulend, J.; Bakker, A.D.; Bacabac, R.G.; Vatsa, A.; Weinbaum, S. Mechanosensation and transduction in osteocytes. Bone 2013, 54, 182–190. [Google Scholar] [CrossRef]
  20. Bolamperti, S.; Villa, I.; Rubinacci, A. Bone remodeling: An operational process ensuring survival and bone mechanical competence. Bone Res. 2022, 10, 1–19. [Google Scholar] [CrossRef]
  21. Henriksen, K.; Neutzsky-Wulff, A.V.; Bonewald, L.F.; Karsdal, M.A. Local communication on and within bone controls bone remodeling. Bone 2009, 44, 1026–1033. [Google Scholar] [CrossRef]
  22. Katagiri, T.; Takahashi, N. Regulatory mechanisms of osteoblast and osteoclast differentiation. Oral Dis. 2002, 8, 147–159. [Google Scholar] [CrossRef]
  23. Tayalia, P.; Mooney, D.J. Controlled Growth Factor Delivery for Tissue Engineering. Adv. Mater. 2009, 21, 3269–3285. [Google Scholar] [CrossRef]
  24. Gentili, C.; Cancedda, R. Cartilage and Bone Extracellular Matrix. Curr. Pharm. Des. 2009, 15, 1334–1348. [Google Scholar] [CrossRef] [PubMed]
  25. Bacakova, L.; Filova, E.; Parizek, M.; Ruml, T.; Svorcik, V. Modulation of cell adhesion, proliferation and differentiation on materials designed for body implants. Biotechnol. Adv. 2011, 29, 739–767. [Google Scholar] [CrossRef] [PubMed]
  26. Ruoslahti, E.; Pierschbacher, M.D. New Perspectives in Cell Adhesion: RGD and Integrins. Science 1987, 238, 491–497. [Google Scholar] [CrossRef] [PubMed]
  27. Ruoslahti, E. RGD and other recognition sequences for integrins. Annu. Rev. Cell Dev. Biol. 1996, 12, 697–715. [Google Scholar] [CrossRef]
  28. Aota, S.; Nomizu, M.; Yamada, K.M. The short amino acid sequence Pro-His-Ser-Arg-Asn in human fibronectin enhances cell-adhesive function. J. Biol. Chem. 1994, 269, 24756–24761. [Google Scholar] [CrossRef]
  29. Dalton, B.A.; McFarland, C.D.; Underwood, P.A.; Steele, J.G. Role of the heparin binding domain of fibronectin in attachment and spreading of human bone-derived cells. J. Cell Sci. 1995, 108, 2083–2092. [Google Scholar] [CrossRef]
  30. Dee, K.C.; Andersen, T.T.; Bizios, R. Design and function of novel osteoblast-adhesive peptides for chemical modification of biomaterials. J. Biomed. Mater. Res. 1998, 40, 371–377. [Google Scholar] [CrossRef]
  31. Staatz, W.D.; Fok, K.F.; Zutter, M.M.; Adams, S.P.; Rodriguez, B.A.; Santoro, S.A. Identification of a tetrapeptide recognition sequence for the α 2 β 1 integrin in collagen. J. Biol. Chem. 1991, 266, 7363–7367. [Google Scholar] [CrossRef]
  32. Sodek, J.; Ganss, B.; McKee, M. Osteopontin. Crit. Rev. Oral Biol. Med. 2000, 11, 279–303. [Google Scholar] [CrossRef]
  33. Hamada, Y.; Nokihara, K.; Okazaki, M.; Fujitani, W.; Matsumoto, T.; Matsuo, M.; Umakoshi, Y.; Takahashi, J.; Matsuura, N. Angiogenic activity of osteopontin-derived peptide SVVYGLR. Biochem. Biophys. Res. Commun. 2003, 310, 153–157. [Google Scholar] [CrossRef]
  34. Hamada, Y.; Egusa, H.; Kaneda, Y.; Hirata, I.; Kawaguchi, N.; Hirao, T.; Matsumoto, T.; Yao, M.; Daito, K.; Suzuki, M.; et al. Synthetic Osteopontin-derived Peptide SVVYGLR can Induce Neovascularization in Artificial Bone Marrow Scaffold Biomaterials. Dent. Mater. J. 2007, 26, 487–492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Patil, A.; Mishra, V.; Thakur, S.; Riyaz, B.; Kaur, A.; Khursheed, R.; Patil, K.; Sathe, B.; Thakur, S. Nanotechnology Derived Nanotools in Biomedical Perspectives: An Update. Curr. Nanosci. 2019, 15, 137–146. [Google Scholar] [CrossRef]
  36. McKnight, A.L.; Kugel, J.L.; Rossman, P.J.; Manduca, A.; Hartmann, L.C.; Ehman, R.L. MR Elastography of Breast Cancer: Preliminary Results. Am. J. Roentgenol. 2002, 178, 1411–1417. [Google Scholar] [CrossRef] [PubMed]
  37. Bercoff, J.; Chaffai, S.; Tanter, M.; Sandrin, L.; Catheline, S.; Fink, M.; Gennisson, J.L.; Meunier, M. In vivo breast tumor detection using transient elastography. Ultrasound Med. Biol. 2003, 29, 1387–1396. [Google Scholar] [CrossRef]
  38. Paszek, M.J.; Zahir, N.; Johnson, K.R.; Lakins, J.N.; Rozenberg, G.I.; Gefen, A.; Reinhart-King, C.A.; Margulies, S.S.; Dembo, M.; Boettiger, D.; et al. Tensional homeostasis and the malignant phenotype. Cancer Cell 2005, 8, 241–254. [Google Scholar] [CrossRef] [Green Version]
  39. Ingber, D.E. Cellular mechanotransduction: Putting all the pieces together again. FASEB J. 2006, 20, 811–827. [Google Scholar] [CrossRef]
  40. Vogel, V.; Sheetz, M. Local force and geometry sensing regulate cell functions. Nat. Rev. Mol. Cell Biol. 2006, 7, 265–275. [Google Scholar] [CrossRef]
  41. Brånemark, P.-I.; Breine, U.; Adell, R.; Hansson, B.O.; Lindström, J.; Ohlsson, Å. Intra-Osseous Anchorage of Dental Prostheses: I. Experimental Studies. Scand. J. Plast. Reconstr. Surg. 1969, 3, 81–100. [Google Scholar] [CrossRef]
  42. Miller, B.F.; Keane, C.B. Encyclopedia and Dictionary of Medicine, Nursing, and Allied Health; WB Saunders Company: Philadelphia, PA, USA, 1983; ISBN 9781455726240. [Google Scholar]
  43. Le Guéhennec, L.; Soueidan, A.; Layrolle, P.; Amouriq, Y. Surface treatments of titanium dental implants for rapid osseointegration. Dent. Mater. 2007, 23, 844–854. [Google Scholar] [CrossRef]
  44. 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]
  45. Mavrogenis, A.; Dimitriou, R.; Parvizi, J.; Babis, G.C. Biology of implant osseointegration. J. Musculoskelet. Neuronal. Interact. 2009, 9, 61–71. [Google Scholar]
  46. Tian, L.; Tang, N.; Ngai, T.; Wu, C.; Ruan, Y.C.; Huang, L.; Qin, L. Hybrid fracture fixation systems developed for orthopaedic applications: A general review. J. Orthop. Transl. 2019, 16, 1–13. [Google Scholar] [CrossRef] [PubMed]
  47. Zahir, A.; Mahmood, U.; Nazir, A.; Hussain, T.; Abid, S. 2—Biomaterials for Medical and Healthcare Products. In Medical Textiles from Natural Resources; Woodhead Publishing: Sawston, UK, 2022; pp. 43–86. [Google Scholar] [CrossRef]
  48. Basu, B.; Ghosh, S. Biomaterials for Musculoskeletal Regeneration: Applications. In Biomaterials for Musculoskeletal Regeneration: Applications; Springer Singapore: Singapore, 2017; pp. 1–262. [Google Scholar] [CrossRef]
  49. Munir, K.S.; Wen, C.; Li, Y.C. Carbon Nanotubes and Graphene as Nanoreinforcements in Metallic Biomaterials: A Review. Adv. Biosyst. 2019, 3, e1800212. [Google Scholar] [CrossRef] [PubMed]
  50. Harvey, A.G.; Hill, E.W.; Bayat, A. Designing implant surface topography for improved biocompatibility. Expert Rev. Med. Devices 2013, 10, 257–267. [Google Scholar] [CrossRef] [PubMed]
  51. Scotchford, C.A. Cell response to surface chemistry in biomaterials. In Cellular Response to Biomaterials; Woodhead Publishing: Sawston, UK, 2009; pp. 462–478. [Google Scholar] [CrossRef]
  52. Gongadze, E.; Kabaso, D.; Bauer, S.; Slivnik, T.; Schmuki, P.; van Rienen, U.; Iglic, A. Adhesion of osteoblasts to a nanorough titanium implant surface. Int. J. Nanomed. 2011, 6, 1801–1816. [Google Scholar] [CrossRef] [Green Version]
  53. Oh, S.; Tobin, E.; Yang, Y.Z.; Carnes, P.L.; Ong, J.L. In vivo evaluation of hydroxyapatite coatings of different crystallinities. Int. J. Oral Max. Impl. 2005, 20, 726–731. [Google Scholar] [CrossRef]
  54. Chen, Y.; Frith, J.E.; Dehghan-Manshadi, A.; Attar, H.; Kent, D.; Soro, N.D.M.; Bermingham, M.J.; Dargusch, M.S. Mechanical properties and biocompatibility of porous titanium scaffolds for bone tissue engineering. J. Mech. Behav. Biomed. Mater. 2017, 75, 169–174. [Google Scholar] [CrossRef]
  55. Deb, P.; Deoghare, A.B.; Borah, A.; Barua, E.; Das Lala, S. Scaffold Development Using Biomaterials: A Review. Mater. Today: Proc. 2018, 5, 12909–12919. [Google Scholar] [CrossRef]
  56. Abbasi, N.; Hamlet, S.; Love, R.M.; Nguyen, N.-T. Porous scaffolds for bone regeneration. J. Sci. Adv. Mater. Devices 2020, 5, 1–9. [Google Scholar] [CrossRef]
  57. Mosas, K.K.A.; Chandrasekar, A.R.; Dasan, A.; Pakseresht, A.; Galusek, D. Recent Advancements in Materials and Coatings for Biomedical Implants. Gels 2022, 8, 323. [Google Scholar] [CrossRef]
  58. Filip, N.; Radu, I.; Veliceasa, B.; Filip, C.; Pertea, M.; Clim, A.; Pinzariu, A.C.; Drochioi, I.C.; Hilitanu, R.L.; Serban, I.L. Biomaterials in Orthopedic Devices: Current Issues and Future Perspectives. Coatings 2022, 12, 1544. [Google Scholar] [CrossRef]
  59. Ning, C.; Zhou, L.; Tan, G. Fourth-generation biomedical materials. Mater. Today 2015, 19, 2–3. [Google Scholar] [CrossRef]
  60. Montoya, C.; Du, Y.; Gianforcaro, A.L.; Orrego, S.; Yang, M.; Lelkes, P.I. On the road to smart biomaterials for bone research: Definitions, concepts, advances, and outlook. Bone Res. 2021, 9, 12. [Google Scholar] [CrossRef] [PubMed]
  61. Barber, C.C.; Burnham, M.; Ojameruaye, O.; McKee, M.D. A systematic review of the use of titanium versus stainless steel implants for fracture fixation. OTA Int. Open Access J. Orthop. Trauma 2021, 4, e138. [Google Scholar] [CrossRef]
  62. Ghunawat, S.; Relhan, V.; Garg, V.K. Hypersensitivity Reactions to Metal Implants: Clinical, Diagnostic and Treatment Overview. Indian J. Clin. Dermatol. 2019, 2, 29–34. [Google Scholar]
  63. Yan, X.; Cao, W.; Li, H. Biomedical Alloys and Physical Surface Modifications: A Mini-Review. Materials 2022, 15, 66. [Google Scholar] [CrossRef]
  64. Li, M.; Yin, T.; Wang, Y.; Du, F.; Zou, X.; Gregersen, H.; Wang, G. Study of biocompatibility of medical grade high nitrogen nickel-free austenitic stainless steel in vitro. Mater. Sci. Eng. C Mater. 2014, 43, 641–648. [Google Scholar] [CrossRef]
  65. Garcia-Mendez, M.C.; Urrutia-Baca, V.H.; Cuao-Moreu, C.A.; Lorenzo-Bonet, E.; Alvarez-Vera, M.; Ortiz-Martinez, D.M.; de la Garza-Ramos, M.A. In Vitro Biocompatibility Evaluation of a New Co-Cr-B Alloy with Potential Biomedical Application. Metals 2021, 11, 1267. [Google Scholar] [CrossRef]
  66. Thanigaivel, S.; Priya, A.K.; Balakrishnan, D.; Dutta, K.; Rajendran, S.; Soto-Moscoso, M. Insight on recent development in metallic biomaterials: Strategies involving synthesis, types and surface modification for advanced therapeutic and biomedical applications. Biochem. Eng. J. 2022, 187, 108522. [Google Scholar] [CrossRef]
  67. Yu, W.; Li, X.; Ma, X.; Xu, X. Biomechanical analysis of inclined and cantilever design with different implant framework materials in mandibular complete-arch implant restorations. J. Prosthet. Dent. 2022, 127, 783.e1–783.e10. [Google Scholar] [CrossRef]
  68. Wilson, J. 1—Metallic biomaterials: State of the art and new challenges. In Fundamental Biomaterials: Metals; Woodhead Publishing: Sawston, UK, 2018; pp. 1–33. [Google Scholar] [CrossRef]
  69. Farooq, M.U.; Mughal, M.P.; Ahmed, N.; Mufti, N.A.; Al-Ahmari, A.M.; He, Y. On the Investigation of Surface Integrity of Ti6Al4V ELI Using Si-Mixed Electric Discharge Machining. Materials 2020, 13, 1549. [Google Scholar] [CrossRef] [Green Version]
  70. Fellah, M.; Labaïz, M.; Assala, O.; Dekhil, L.; Taleb, A.; Rezag, H.; Iost, A. Tribological behavior of Ti-6Al-4V and Ti-6Al-7Nb Alloys for Total Hip Prosthesis. Adv. Tribol. 2014, 2014, 1–13. [Google Scholar] [CrossRef] [Green Version]
  71. Anene, F.A.; Jaafar, C.N.A.; Zainol, I.; Hanim, M.A.A.; Suraya, M.T. Biomedical materials: A review of titanium based alloys. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2021, 235, 3792–3805. [Google Scholar] [CrossRef]
  72. Khadija, G.; Saleem, A.; Akhtar, Z.; Naqvi, Z.; Gull, M.; Masood, M.; Mukhtar, S.; Batool, M.; Saleem, N.; Rasheed, T.; et al. Short term exposure to titanium, aluminum and vanadium (Ti 6Al 4V) alloy powder drastically affects behavior and antioxidant metabolites in vital organs of male albino mice. Toxicol. Rep. 2018, 5, 765–770. [Google Scholar] [CrossRef] [PubMed]
  73. Ebel, T. 5.3—Titanium MIM for manufacturing of medical implants and devices. In Titanium in Medical and Dental Applications; Woodhead Publishing: Sawston, UK, 2018; pp. 531–551. [Google Scholar] [CrossRef]
  74. Darouiche, R.O. Current concepts—Treatment of Infections Associated with Surgical Implants. New Engl. J. Med. 2004, 350, 1422–1429. [Google Scholar] [CrossRef]
  75. Tan, L.; Dong, J.; Chen, J.; Yang, K. Development of magnesium alloys for biomedical applications: Structure, process to property relationship. Mater. Technol. 2018, 33, 235–243. [Google Scholar] [CrossRef]
  76. Wang, J.L.; Xu, J.K.; Hopkins, C.; Chow, D.H.K.; Qin, L. Biodegradable Magnesium-Based Implants in Orthopedics—A General Review and Perspectives. Adv. Sci. 2020, 7, 1902443. [Google Scholar] [CrossRef] [Green Version]
  77. Amukarimi, S.; Mozafari, M. Biodegradable magnesium-based biomaterials: An overview of challenges and opportunities. Medcomm 2021, 2, 123–144. [Google Scholar] [CrossRef]
  78. Sarian, M.N.; Iqbal, N.; Sotoudehbagha, P.; Razavi, M.; Ahmed, Q.U.; Sukotjo, C.; Hermawan, H. Potential bioactive coating system for high-performance absorbable magnesium bone implants. Bioact. Mater. 2022, 12, 42–63. [Google Scholar] [CrossRef]
  79. Li, Y.; Zhao, S.; Li, S.; Ge, Y.; Wang, R.; Zheng, L.; Xu, J.; Sun, M.; Jiang, Q.; Zhang, Y.; et al. Surface Engineering of Biodegradable Magnesium Alloys for Enhanced Orthopedic Implants. Small 2019, 15, e1904486. [Google Scholar] [CrossRef]
  80. Nayak, G.S.; Carradò, A.; Masson, P.; Pourroy, G.; Mouillard, F.; Migonney, V.; Falentin-Daudre, C.; Pereira, C.; Palkowski, H. Trends in Metal-Based Composite Biomaterials for Hard Tissue Applications. JOM 2021, 74, 102–125. [Google Scholar] [CrossRef]
  81. Pan, H.; Han, J.J.; Park, Y.-D.; Cho, T.H.; Hwang, S.J. Effect of sustained release of rhBMP-2 from dried and wet hyaluronic acid hydrogel carriers compared with direct dip coating of rhBMP-2 on peri-implant osteogenesis of dental implants in canine mandibles. J. Cranio-Maxill. Surg. 2016, 44, 116–125. [Google Scholar] [CrossRef] [PubMed]
  82. Neacşu, I.A.; Nicoară, A.I.; Vasile, O.R.; Vasile, B.S. Inorganic micro- and nanostructured implants for tissue engineering. Appl. Nanobiomater. 2016, 4, 271–295. [Google Scholar] [CrossRef]
  83. Metikoš-Huković, M.; Tkalčec, E.; Kwokal, A.; Piljac, J. An in vitro study of Ti and Ti-alloys coated with sol–gel derived hydroxyapatite coatings. Surf. Coat. Technol. 2003, 165, 40–50. [Google Scholar] [CrossRef]
  84. Montenero, A.; Gnappi, G.; Ferrari, F.; Cesari, M.; Salvioli, E.; Mattogno, L.; Kaciulis, S.; Fini, M. Sol-gel derived hydroxyapatite coatings on titanium substrate. J. Mater. Sci. 2000, 35, 2791–2797. [Google Scholar] [CrossRef]
  85. Barrios, E.; Fox, D.; Li Sip, Y.Y.; Catarata, R.; Calderon, J.E.; Azim, N.; Afrin, S.; Zhang, Z.Y.; Zhai, L. Nanomaterials in Advanced, High-Performance Aerogel Composites: A Review. Polymers 2019, 11, 726. [Google Scholar] [CrossRef]
  86. Horvat, G.; Xhanari, K.; Finšgar, M.; Gradišnik, L.; Maver, U.; Knez, Z.; Novak, Z. Novel ethanol-induced pectin–xanthan aerogel coatings for orthopedic applications. Carbohydr. Polym. 2017, 166, 365–376. [Google Scholar] [CrossRef]
  87. Hawkins, M.E.; Pletcher, D.L.; Thomas, B.; Zhang, K.; Brinkerhuff, H.E. Methods of Preparing Hydrogel Coatings. U.S. Patent US8110242B2, 7 February 2012. [Google Scholar]
  88. Zhang, X.; Han, W.; Fan, D.; Zheng, Y. Electroless iron plating on pure magnesium for biomedical applications. Mater. Lett. 2014, 130, 154–156. [Google Scholar] [CrossRef]
  89. González-Murguía, J.L.; Lucien, V.; Alpuche-Avilés, M. Electroless Deposits of ZnO and Hybrid ZnO/Ag Nanoparticles on Mg-Ca0.3 Alloy Surface: Multiscale Characterization. Coatings 2022, 12, 1109. [Google Scholar] [CrossRef]
  90. Buchtík, M.; Kosár, P.; Wasserbauer, J.; Tkacz, J.; Doležal, P. Characterization of Electroless Ni–P Coating Prepared on a Wrought ZE10 Magnesium Alloy. Coatings 2018, 8, 96. [Google Scholar] [CrossRef] [Green Version]
  91. Liu, Y.; Wu, G.; De Groot, K. Biomimetic coatings for bone tissue engineering of critical-sized defects. J. R. Soc. Interface 2010, 7, S631–S647. [Google Scholar] [CrossRef] [PubMed]
  92. Koju, N.; Sikder, P.; Ren, Y.; Zhou, H.; Bhaduri, S.B. Biomimetic coating technology for orthopedic implants. Curr. Opin. Chem. Eng. 2017, 15, 49–55. [Google Scholar] [CrossRef] [Green Version]
  93. Shin, K.; Acri, T.; Geary, S.; Salem, A.K. Biomimetic Mineralization of Biomaterials Using Simulated Body Fluids for Bone Tissue Engineering and Regenerative Medicine. Tissue Eng. Part A 2017, 23, 1169–1180. [Google Scholar] [CrossRef]
  94. Aryal, S.; Bhattarai, S.R.; Bahadur, S.R.K.; Khil, M.S.; Lee, D.-R.; Kim, H.Y. Carbon nanotubes assisted biomimetic synthesis of hydroxyapatite from simulated body fluid. Mater. Sci. Eng. A Struct. 2006, 426, 202–207. [Google Scholar] [CrossRef]
  95. Reggente, M.; Kriegel, S.; He, W.; Masson, P.; Pourroy, G.; Mura, F.; Faerber, J.; Passeri, D.; Rossi, M.; Palkowski, H.; et al. How alkali-activated Ti surfaces affect the growth of tethered PMMA chains: A close-up study on the PMMA thickness and surface morphology. Pure Appl. Chem. 2019, 91, 1687–1694. [Google Scholar] [CrossRef]
  96. Chouirfa, H.; Evans, M.D.M.; Castner, D.G.; Bean, P.; Mercier, D.; Galtayries, A.; Falentin-Daudré, C.; Migonney, V. Grafting of architecture controlled poly(styrene sodium sulfonate) onto titanium surfaces using bio-adhesive molecules: Surface characterization and biological properties. Biointerphases 2017, 12, 02C418. [Google Scholar] [CrossRef]
  97. Kim, H.-U.; Kulkarni, A.; Ha, S.; Shin, D.; Kim, T. Note: Electric field assisted megasonic atomization for size-controlled nanoparticles. Rev. Sci. Instrum. 2017, 88, 076106. [Google Scholar] [CrossRef]
  98. John, A.T.; Tricoli, A. Flame assisted synthesis of nanostructures for device applications. Adv. Phys. X 2022, 7, 1997153. [Google Scholar] [CrossRef]
  99. Aguilar-Frutis, M.; Kumar, S.; Falcony, C. Spray-pyrolyzed hydroxyapatite thin-film coatings. Surf. Coat. Technol. 2009, 204, 1116–1120. [Google Scholar] [CrossRef]
  100. Bargavi, P.; Chitra, S.; Durgalakshmi, D.; Radha, G.; Balakumar, S. Zirconia reinforced bio-active glass coating by spray pyrolysis: Structure, surface topography, in-vitro biological evaluation and antibacterial activities. Mater. Today Commun. 2020, 25, 101253. [Google Scholar] [CrossRef]
  101. Meierhofer, F.; Mädler, L.; Fritsching, U. Nanoparticle evolution in flame spray pyrolysis—Process design via experimental and computational analysis. AIChE J. 2020, 66, e16885. [Google Scholar] [CrossRef]
  102. Nasiri, N.; Ceramidas, A.; Mukherjee, S.; Panneerselvan, A.; Nisbet, D.R.; Tricoli, A. Ultra-Porous Nanoparticle Networks: A Biomimetic Coating Morphology for Enhanced Cellular Response and Infiltration. Sci. Rep. 2016, 6, 24305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Oliver, J.A.N.; Su, Y.C.; Lu, X.N.; Kuo, P.-H.; Du, J.C.; Zhu, D.H. Bioactive glass coatings on metallic implants for biomedical applications. Bioact. Mater. 2019, 4, 261–270. [Google Scholar] [CrossRef] [PubMed]
  104. Fasiku, V.O.; Owonubi, S.J.; Mukwevho, E.; Aderibigbe, B.; Sadiku, E.R.; Lemmer, Y.; Ibrahim, I.D.; Mochane, J.; Daramola, O.O.; Selatile, K.; et al. Polymeric Materials in Coatings for Biomedical Applications. In Advanced Coating Materials; Scrivener Publishing: Beverly, MA, USA, 2018; pp. 481–518. [Google Scholar] [CrossRef]
  105. Zhou, X.; Siman, R.; Lu, L.; Mohanty, P. Argon atmospheric plasma sprayed hydroxyapatite/Ti composite coating for biomedical applications. Surf. Coatings Technol. 2012, 207, 343–349. [Google Scholar] [CrossRef]
  106. Liu, Y.; Li, H.; Zhang, B. Nanostructured Ceramic Coating Biomaterials. In Advanced Nanomaterials and Coatings; Thermal Spray: Sun Prairie, WI, USA, 2019; pp. 291–311. [Google Scholar] [CrossRef]
  107. Liu, Y.; Dang, Z.; Wang, Y.; Huang, J.; Li, H. Hydroxyapatite/graphene-nanosheet composite coatings deposited by vacuum cold spraying for biomedical applications: Inherited nanostructures and enhanced properties. Carbon 2014, 67, 250–259. [Google Scholar] [CrossRef]
  108. Tang, J.; Zhao, Z.; Cui, X.; Wang, J.; Xiong, T. Microstructure and bioactivity of a cold sprayed rough/porous Ta coating on Ti6Al4V substrate. Sci. China Technol. Sci. 2020, 63, 731–739. [Google Scholar] [CrossRef]
  109. Onoki, T.; Hashida, T. New method for hydroxyapatite coating of titanium by the hydrothermal hot isostatic pressing technique. Surf. Coat. Technol. 2006, 200, 6801–6807. [Google Scholar] [CrossRef]
  110. Ali, A.; Iqbal, F.; Ahmad, A.; Ikram, F.; Nawaz, A.; Chaudhry, A.A.; Siddiqi, S.A.; Rehman, I. Hydrothermal deposition of high strength calcium phosphate coatings on magnesium alloy for biomedical applications. Surf. Coat. Technol. 2019, 357, 716–727. [Google Scholar] [CrossRef]
  111. Yang, C.-W. Development of Hydrothermally Synthesized Hydroxyapatite Coatings on Metallic Substrates and Weibull’s Reliability Analysis of its Shear Strength. Int. J. Appl. Ceram. Technol. 2015, 12, 282–293. [Google Scholar] [CrossRef]
  112. Kien, P.T.; Phu, H.D.; Linh, N.V.V.; Quyen, T.N.; Hoa, N.T. Recent Trends in Hydroxyapatite (HA) Synthesis and the Synthesis Report of Nanostructure HA by Hydrothermal Reaction. Adv. Exp. Med. Biol. 2018, 1077, 343–354. [Google Scholar] [CrossRef]
  113. Wenzel, K.W.C.; Thomas, J.; Hecker, M.; Brückner, W. Thin Film Preparation and Characterization Technique; Wiley-VCH: Weinheim, Germany, 2003. [Google Scholar] [CrossRef]
  114. Creighton, J.; Ho, P. Introduction to chemical vapor deposition (CVD). Chem. Vap. Depos. 2001, 2, 1–22. [Google Scholar]
  115. Smieszek, A.; Seweryn, A.; Marcinkowska, K.; Sikora, M.; Lawniczak-Jablonska, K.; Witkowski, B.S.; Kuzmiuk, P.; Godlewski, M.; Marycz, A.K. Titanium Dioxide Thin Films Obtained by Atomic Layer Deposition Promotes Osteoblasts’ Viability and Differentiation Potential While Inhibiting Osteoclast Activity—Potential Application for Osteoporotic Bone Regeneration. Materials 2020, 13, 4817. [Google Scholar] [CrossRef] [PubMed]
  116. Motola, M.; Capek, J.; Zazpe, R.; Bacova, J.; Hromadko, L.; Bruckova, L.; Ng, S.; Handl, J.; Spotz, Z.; Knotek, P.; et al. Thin TiO2 Coatings by ALD Enhance the Cell Growth on TiO2 Nanotubular and Flat Substrates. ACS Appl. Bio Mater. 2020, 3, 6447–6456. [Google Scholar] [CrossRef]
  117. Pedersen, H.; Elliott, S.D. Studying chemical vapor deposition processes with theoretical chemistry. Theor. Chim. Acta 2014, 133, 1–10. [Google Scholar] [CrossRef]
  118. Patel, K.D.; Singh, R.K.; Lee, J.-H.; Kim, H.-W. Electrophoretic coatings of hydroxyapatite with various nanocrystal shapes. Mater. Lett. 2019, 234, 148–154. [Google Scholar] [CrossRef]
  119. Wang, Y.; Deen, I.; Zhitomirsky, I. Electrophoretic deposition of polyacrylic acid and composite films containing nanotubes and oxide particles. J. Colloid Interface Sci. 2011, 362, 367–374. [Google Scholar] [CrossRef] [PubMed]
  120. Avcu, E.; Bastan, F.E.; Abdullah, H.Z.; Rehman, M.A.U.; Avcu, Y.Y.; Boccaccini, A.R. Electrophoretic deposition of chitosan-based composite coatings for biomedical applications: A review. Prog. Mater. Sci. 2019, 103, 69–108. [Google Scholar] [CrossRef]
  121. Aydemir, T.; Liverani, L.; Pastore, J.I.; Ceré, S.M.; Goldmann, W.H.; Boccaccini, A.R.; Ballarre, J. Functional behavior of chitosan/gelatin/silica-gentamicin coatings by electrophoretic deposition on surgical grade stainless steel. Mater. Sci. Eng. C Mater. 2020, 115, 111062. [Google Scholar] [CrossRef]
  122. Wan, P.; Wang, W.; Zheng, L.; Qin, L.; Yang, K. One-step electrodeposition synthesis of bisphosphonate loaded magnesium implant: A strategy to modulate drug release for osteoporotic fracture healing. J. Mater. Sci. Technol. 2021, 78, 92–99. [Google Scholar] [CrossRef]
  123. Bree, G.; Geaney, H.; Ryan, K.M. Electrophoretic Deposition of Tin Sulfide Nanocubes as High-Performance Lithium-Ion Battery Anodes. ChemElectroChem 2019, 6, 3049–3056. [Google Scholar] [CrossRef]
  124. Frank, M.J.; Walter, M.S.; Tiainen, H.; Rubert, M.; Monjo, M.; Lyngstadaas, S.P.; Haugen, H.J. Coating of metal implant materials with strontium. J. Mater. Sci. Mater. Electron. 2013, 24, 2537–2548. [Google Scholar] [CrossRef] [PubMed]
  125. Aboudzadeh, N.; Dehghanian, C.; Shokrgozar, M.A. Effect of electrodeposition parameters and substrate on morphology of Si-HA coating. Surf. Coat. Technol. 2019, 375, 341–351. [Google Scholar] [CrossRef]
  126. Benea, L.; Celis, J.P. Reactivity of porous titanium oxide film and chitosan layer electrochemically formed on Ti-6Al-4V alloy in biological solution. Surf. Coat. Technol. 2018, 354, 145–152. [Google Scholar] [CrossRef]
  127. Khazeni, D.; Saremi, M.; Soltani, R. Development of HA-CNTs composite coating on AZ31 magnesium alloy by cathodic electrodeposition. Part 1: Microstructural and mechanical characterization. Ceram. Int. 2019, 45, 11174–11185. [Google Scholar] [CrossRef]
  128. Narayanan, R.; Seshadri, S.K.; Kwon, T.Y.; Kim, K.H. Electrochemical nano-grained calcium phosphate coatings on Ti–6Al–4V for biomaterial applications. Scr. Mater. 2007, 56, 229–232. [Google Scholar] [CrossRef]
  129. Zielinski, A.; Bartmanski, M. Electrodeposited Biocoatings, Their Properties and Fabrication Technologies: A Review. Coatings 2020, 10, 782. [Google Scholar] [CrossRef]
  130. Fu, Y.; Mo, A.C. A Review on the Electrochemically Self-organized Titania Nanotube Arrays: Synthesis, Modifications, and Biomedical Applications. Nanoscale Res. Lett. 2018, 13, 187. [Google Scholar] [CrossRef]
  131. Tan, A.W.; Pingguan-Murphy, B.; Ahmad, R.; Akbar, S.A. Review of titania nanotubes: Fabrication and cellular response. Ceram. Int. 2012, 38, 4421–4435. [Google Scholar] [CrossRef]
  132. Lai, M.; Cai, K.; Zhao, L.; Chen, X.; Hou, Y.; Yang, Z. Surface Functionalization of TiO2 Nanotubes with Bone Morphogenetic Protein 2 and Its Synergistic Effect on the Differentiation of Mesenchymal Stem Cells. Biomacromolecules 2011, 12, 1097–1105. [Google Scholar] [CrossRef]
  133. Huang, P.; Wang, J.; Lai, S.; Liu, F.; Ni, N.; Cao, Q.; Liu, W.; Deng, D.Y.B.; Zhou, W. Surface modified titania nanotubes containing anti-bacterial drugs for controlled delivery nanosystems with high bioactivity. J. Mater. Chem. B 2014, 2, 8616–8625. [Google Scholar] [CrossRef]
  134. Adeleke, S.A.; Bushroa, A.R.; Sopyan, I. Recent development of calcium phosphate-based coatings on titanium alloy implants. Surf. Eng. Appl. Electrochem. 2017, 53, 419–433. [Google Scholar] [CrossRef]
  135. Yang, T.-C.; Shu, H.-Y.; Chen, H.-T.; Chung, C.-J.; He, J.-L. Interface between grown osteoblast and micro-arc oxidized bioactive layers. Surf. Coat. Technol. 2014, 259, 185–192. [Google Scholar] [CrossRef]
  136. Huang, L.; Cai, B.; Huang, Y.; Wang, J.; Zhu, C.; Shi, K.; Song, Y.; Feng, G.; Liu, L.; Zhang, L. Comparative Study on 3D Printed Ti6Al4V Scaffolds with Surface Modifications Using Hydrothermal Treatment and Microarc Oxidation to Enhance Osteogenic Activity. ACS Omega 2021, 6, 1465–1476. [Google Scholar] [CrossRef] [PubMed]
  137. Xu, C.; Chen, L.-Y.; Zheng, C.-B.; Zhang, H.-Y.; Zhao, C.-H.; Wang, Z.-X.; Lu, S.; Zhang, J.-W.; Zhang, L.-C. Improved Wear and Corrosion Resistance of Microarc Oxidation Coatings on Ti–6Al–4V Alloy with Ultrasonic Assistance for Potential Biomedical Applications. Adv. Eng. Mater. 2021, 23, 2001433. [Google Scholar] [CrossRef]
  138. Wang, Z.; Ye, F.; Chen, L.; Lv, W.; Zhang, Z.; Zang, Q.; Peng, J.; Sun, L.; Lu, S. Preparation and Degradation Characteristics of MAO/APS Composite Bio-Coating in Simulated Body Fluid. Coatings 2021, 11, 667. [Google Scholar] [CrossRef]
  139. Wang, R.; Ni, S.; Ma, L.; Li, M. Porous construction and surface modification of titanium-based materials for osteogenesis: A review. Front. Bioeng. Biotechnol. 2022, 10, 973297. [Google Scholar] [CrossRef]
  140. Li, Z.; Khor, K.A. Preparation and Properties of Coatings and Thin Films on Metal Implants. Encycl. Biomed. Eng. 2019, 1, 203–212. [Google Scholar] [CrossRef]
  141. Li, P.; Chu, P. Thin film deposition technologies and processing of biomaterials. In Thin Film Coatings for Biomaterials and Biomedical Applications; Elsevier: Amsterdam, The Netherlands, 2016; pp. 3–28. [Google Scholar] [CrossRef]
  142. Jilani, A.; Abdel-Wahab, M.S.; Hammad, A.H. Advance Deposition Techniques for Thin Film and Coating. In Modern Technologies for Creating the Thin-Film Systems and Coatings; Nikitenkov, N.N., Ed.; IntechOpen: London, UK, 2017; pp. 137–149. [Google Scholar]
  143. Duta, L.; Popescu, A.C. Current Status on Pulsed Laser Deposition of Coatings from Animal-Origin Calcium Phosphate Sources. Coatings 2019, 9, 335. [Google Scholar] [CrossRef] [Green Version]
  144. Bianchi, M.; Gambardella, A.; Berni, M.; Panseri, S.; Montesi, M.; Lopomo, N.; Tampieri, A.; Marcacci, M.; Russo, A. Surface morphology, tribological properties and in vitro biocompatibility of nanostructured zirconia thin films. J. Mater. Sci. Mater. Med. 2016, 27, 1–10. [Google Scholar] [CrossRef]
  145. Macleod, H.A. Chapter 1—Recent developments in deposition techniques for optical thin films and coatings. In Optical Thin Films and Coatings, 2nd ed.; Woodhead Publishing: Sawston, UK, 2018; pp. 3–23. [Google Scholar] [CrossRef]
  146. Pelletier, J.; Anders, A. Plasma-based ion implantation and deposition: A review of physics, technology, and applications. IEEE Trans. Plasma Sci. 2005, 33, 1944–1959. [Google Scholar] [CrossRef] [Green Version]
  147. Zhao, Y.; Wong, S.M.; Wong, H.M.; Wu, S.; Hu, T.; Yeung, K.W.K.; Chu, P.K. Effects of Carbon and Nitrogen Plasma Immersion Ion Implantation on In vitro and In vivo Biocompatibility of Titanium Alloy. ACS Appl. Mater. Interfaces 2013, 5, 1510–1516. [Google Scholar] [CrossRef] [PubMed]
  148. Zhao, Y.; Jamesh, M.I.; Li, W.K.; Wu, G.; Wang, C.; Zheng, Y.; Yeung, K.W.; Chu, P.K. Enhanced antimicrobial properties, cytocompatibility, and corrosion resistance of plasma-modified biodegradable magnesium alloys. Acta Biomater. 2014, 10, 544–556. [Google Scholar] [CrossRef] [PubMed]
  149. Wu, G.; Feng, K.; Shanaghi, A.; Zhao, Y.; Xu, R.; Yuan, G.; Chu, P.K. Effects of surface alloying on electrochemical corrosion behavior of oxygen-plasma-modified biomedical magnesium alloy. Surf. Coat. Technol. 2012, 206, 3186–3195. [Google Scholar] [CrossRef]
  150. Pham, M.T.; Matz, W.; Reuther, H.; Richter, E.; Steiner, G.; Oswald, S. Ion beam sensitizing of titanium surfaces to hydroxyapatite formation. Surf. Coat. Technol. 2000, 128–129, 313–319. [Google Scholar] [CrossRef]
  151. Lu, K.; Zhu, J.; Guo, D.; Yang, M.; Sun, H.; Wang, Z.; Hui, X.; Wu, Y. Microstructures, Corrosion Resistance and Wear Resistance of High-Entropy Alloys Coatings with Various Compositions Prepared by Laser Cladding: A Review. Coatings 2022, 12, 1023. [Google Scholar] [CrossRef]
  152. Bajda, S.; Liu, Y.; Tosi, R.; Cholewa-Kowalska, K.; Krzyzanowski, M.; Dziadek, M.; Kopyscianski, M.; Dymek, S.; Polyakov, A.V.; Semenova, I.P.; et al. Laser cladding of bioactive glass coating on pure titanium substrate with highly refined grain structure. J. Mech. Behav. Biomed. Mater. 2021, 119, 104519. [Google Scholar] [CrossRef]
  153. Liu, F.; Ji, Y.; Meng, Q.; Li, Z. Microstructure and corrosion resistance of laser cladding and friction stir processing hybrid modification Al-Si coatings on AZ31B. Vacuum 2016, 133, 31–37. [Google Scholar] [CrossRef]
  154. Zheng, Y. 3—Fabrication on bioinspired surfaces. In Bioinspired Design of Materials Surfaces; Elsevier: Amsterdam, The Netherlands, 2019; pp. 99–146. [Google Scholar] [CrossRef]
  155. Varghese, R.J.; Sakho, E.H.M.; Parani, S.; Thomas, S.; Oluwafemi, O.S.; Wu, J.H. Introduction to nanomaterials: Synthesis and applications. In Nanomaterials for Solar Cell Applications; Thomas, S., Sakho, E.H.M., Kalarikkal, N., Oluwafemi, S.O., Wu, J., Eds.; Elsevier Inc.: Amsterdam, The Netherlands, 2019; pp. 75–95. [Google Scholar] [CrossRef]
  156. Rezk, A.I.; Mousa, H.M.; Lee, J.; Park, C.H.; Kim, C.S. Composite PCL/HA/simvastatin electrospun nanofiber coating on biodegradable Mg alloy for orthopedic implant application. J. Coat. Technol. Res. 2019, 16, 477–489. [Google Scholar] [CrossRef]
  157. Kiran, A.S.K.; Kizhakeyil, A.; Ramalingam, R.; Verma, N.K.; Lakshminarayanan, R.; Kumar, T.S.; Doble, M.; Ramakrishna, S. Drug loaded electrospun polymer/ceramic composite nanofibrous coatings on titanium for implant related infections. Ceram. Int. 2019, 45, 18710–18720. [Google Scholar] [CrossRef]
  158. Khosravi, F.; Khorasani, S.N.; Khalili, S.; Neisiany, R.E.; Ghomi, E.R.; Ejeian, F.; Das, O.; Nasr-Esfahani, M.H. Development of a Highly Proliferated Bilayer Coating on 316L Stainless Steel Implants. Polymers 2020, 12, 1022. [Google Scholar] [CrossRef]
  159. Palmara, G.; Frascella, F.; Roppolo, I.; Chiappone, A.; Chiadò, A. Functional 3D printing: Approaches and bioapplications. Biosens. Bioelectron. 2021, 175, 112849. [Google Scholar] [CrossRef]
  160. Beg, S.; Almalki, W.H.; Malik, A.; Farhan, M.; Aatif, M.; Rahman, Z.; Alruwaili, N.K.; Alrobaian, M.; Tarique, M.; Rahman, M. 3D printing for drug delivery and biomedical applications. Drug Discov. Today 2020, 25, 1668–1681. [Google Scholar] [CrossRef]
  161. Zhang, J.; Yang, W.; Vo, A.Q.; Feng, X.; Ye, X.; Kim, D.W.; Repka, M.A. Hydroxypropyl methylcellulose-based controlled release dosage by melt extrusion and 3D printing: Structure and drug release correlation. Carbohydr. Polym. 2017, 177, 49–57. [Google Scholar] [CrossRef]
  162. Maver, T.; Mastnak, T.; Mihelič, M.; Maver, U.; Finšgar, M. Clindamycin-Based 3D-Printed and Electrospun Coatings for Treatment of Implant-Related Infections. Materials 2021, 14, 1464. [Google Scholar] [CrossRef]
  163. Lee, J.; Song, C.; Noh, I.; Song, S.; Rhee, Y.-S. Hot-Melt 3D Extrusion for the Fabrication of Customizable Modified-Release Solid Dosage Forms. Pharmaceutics 2020, 12, 738. [Google Scholar] [CrossRef]
  164. AlGahtani, M.S.; Mohammed, A.A.; Ahmad, J.; Saleh, E. Development of a 3D Printed Coating Shell to Control the Drug Release of Encapsulated Immediate-Release Tablets. Polymers 2020, 12, 1395. [Google Scholar] [CrossRef]
  165. Kravanja, K.A.; Finšgar, M. A review of techniques for the application of bioactive coatings on metal-based implants to achieve controlled release of active ingredients. Mater. Des. 2022, 217, 110653. [Google Scholar] [CrossRef]
  166. Maver, U.; Xhanari, K.; Žižek, M.; Gradišnik, L.; Repnik, K.; Potočnik, U.; Finšgar, M. Carboxymethyl cellulose/diclofenac bioactive coatings on AISI 316LVM for controlled drug delivery, and improved osteogenic potential. Carbohydr. Polym. 2020, 230, 115612. [Google Scholar] [CrossRef]
  167. Richardson, J.J.; Cui, J.W.; Bjornmalm, M.; Braunger, J.A.; Ejima, H.; Caruso, F. Innovation in Layer-by-Layer Assembly. Chem. Rev. 2016, 116, 14828–14867. [Google Scholar] [CrossRef] [Green Version]
  168. De Villiers, M.M.; Otto, D.P.; Strydom, S.J.; Lvov, Y.M. Introduction to nanocoatings produced by layer-by-layer (LbL) self-assembly. Adv. Drug Deliv. Rev. 2011, 63, 701–715. [Google Scholar] [CrossRef]
  169. Yavari, S.A.; Croes, M.; Akhavan, B.; Jahanmard, F.; Eigenhuis, C.C.; Dadbakhsh, S.; Vogely, H.C.; Bilek, M.M.; Fluit, A.C.; Boel, C.H.E.; et al. Layer by layer coating for bio-functionalization of additively manufactured meta-biomaterials. Addit. Manuf. 2019, 32, 100991. [Google Scholar] [CrossRef]
  170. Yang, G.; Park, S.-J. Conventional and Microwave Hydrothermal Synthesis and Application of Functional Materials: A Review. Materials 2019, 12, 1177. [Google Scholar] [CrossRef] [Green Version]
  171. Ghasemi-Mobarakeh, L.; Kolahreez, D.; Ramakrishna, S.; Williams, D. Key terminology in biomaterials and biocompatibility. Curr. Opin. Biomed. Eng. 2019, 10, 45–50. [Google Scholar] [CrossRef]
  172. Zhao, X.; Courtney, J.; Qian, H. 1—Introduction to bioactive materials in medicine. In Bioactive Materials in Medicine; Woodhead Publishing: Sawston, UK, 2011. [Google Scholar] [CrossRef]
  173. Hermawan, H. Updates on the research and development of absorbable metals for biomedical applications. Prog. Biomater. 2018, 7, 93–110. [Google Scholar] [CrossRef] [Green Version]
  174. Hübler, R.; Cozza, A.; Marcondes, T.L.; Souza, R.B.; Fiori, F.F. Wear and corrosion protection of 316-L femoral implants by deposition of thin films. Surf. Coat. Technol. 2001, 142–144, 1078–1083. [Google Scholar] [CrossRef]
  175. Magnus, F.; Sveinsson, O.B.; Olafsson, S.; Gudmundsson, J.T. Current-voltage-time characteristics of the reactive Ar/N2 high power impulse magnetron sputtering discharge. J. Appl. Phys. 2011, 110, 083306. [Google Scholar] [CrossRef]
  176. Kao, C.-T.; Ding, S.-J.; Chen, Y.-C.; Huang, T.-H. The anticorrosion ability of titanium nitride (TiN) plating on an orthodontic metal bracket and its biocompatibility. J. Biomed. Mater. Res. 2002, 63, 786–792. [Google Scholar] [CrossRef]
  177. Gobbi, S.J.; Gobbi, V.J.; Reinke, G.; Rocha, Y. Orthopedic Implants: Coating with TiN. Biomed. J. Sci. Tech. Res. 2019, 16, 11740–11742. [Google Scholar] [CrossRef]
  178. van Hove, R.P.; Sierevelt, I.N.; van Royen, B.J.; Nolte, P.A. Titanium-Nitride Coating of Orthopaedic Implants: A Review of the Literature. BioMed Res. Int. 2015, 2015, 1–9. [Google Scholar] [CrossRef]
  179. Yuan, Z.; Han, Y.; Zang, S.; Chen, J.; He, G.; Chai, Y.; Yang, Z.; Fu, Q. Damage evolution behavior of TiN/Ti multilayer coatings under high-speed impact conditions. Surf. Coat. Technol. 2021, 426, 127807. [Google Scholar] [CrossRef]
  180. Caiazzo, F.C.; Sisti, V.; Trasatti, S.P.; Trasatti, S. Electrochemical Characterization of Multilayer Cr/CrN-Based Coatings. Coatings 2014, 4, 508–526. [Google Scholar] [CrossRef] [Green Version]
  181. Cui, W.; Qin, G.; Duan, J.; Wang, H. A graded nano-TiN coating on biomedical Ti alloy: Low friction coefficient, good bonding and biocompatibility. Mater. Sci. Eng. C Mater. 2017, 71, 520–528. [Google Scholar] [CrossRef]
  182. Hussein, M.A.; Adesina, A.Y.; Kumar, A.M.; Sorour, A.A.; Ankah, N.; Al-Aqeeli, N. Mechanical, in-vitro corrosion, and tribological characteristics of TiN coating produced by cathodic arc physical vapor deposition on Ti20Nb13Zr alloy for biomedical applications. Thin Solid Films 2020, 709, 138183. [Google Scholar] [CrossRef]
  183. Uddin, G.M.; Jawad, M.; Ghufran, M.; Saleem, M.W.; Raza, M.A.; Rehman, Z.U.; Arafat, S.M.; Irfan, M.; Waseem, B. Experimental investigation of tribo-mechanical and chemical properties of TiN PVD coating on titanium substrate for biomedical implants manufacturing. Int. J. Adv. Manuf. Technol. 2019, 102, 1391–1404. [Google Scholar] [CrossRef]
  184. Carradò, A.; Schmerber, G.; Pelletier, H. Structural and mechanical investigations of magnetron sputtering TiO2/Ti/TiN multilayer films on Si(100) substrate. J. Coat. Technol. Res. 2010, 7, 821–829. [Google Scholar] [CrossRef]
  185. Blunn Gw, M.R.M. Four Station Knee Simulator Wear Testing Comparing Titanium Niobium Nitride with Cobalt Chrome. J. Bioeng. Biomed. Sci. 2013, 3, 1000125. [Google Scholar] [CrossRef] [Green Version]
  186. Liu, H.; Zhang, X.; Jin, S.; Zhao, Y.; Ren, L.; Yang, K. Effect of copper-doped titanium nitride coating on angiogenesis. Mater. Lett. 2020, 269, 127634. [Google Scholar] [CrossRef]
  187. Kumar, D.D.; Kaliaraj, G.S. Multifunctional zirconium nitride/copper multilayer coatings on medical grade 316L SS and titanium substrates for biomedical applications. J. Mech. Behav. Biomed. Mater. 2018, 77, 106–115. [Google Scholar] [CrossRef]
  188. Chen, Y.-I.; Lin, J.-H.; Chou, C.-C. Oxidation resistance and mechanical properties of Ta–Al–N coatings. Surf. Coat. Technol. 2016, 303, 41–47. [Google Scholar] [CrossRef]
  189. Levine, B.R.; Sporer, S.; Poggie, R.A.; Della Valle, C.J.; Jacobs, J.J. Experimental and clinical performance of porous tantalum in orthopedic surgery. Biomaterials 2006, 27, 4671–4681. [Google Scholar] [CrossRef]
  190. Mendizabal, L.; Bayón, R.; G-Berasategui, E.; Barriga, J.; Gonzalez, J. Effect of N2 flow rate on the microstructure and electrochemical behavior of TaNx films deposited by modulated pulsed power magnetron sputtering. Thin Solid Films 2016, 610, 1–9. [Google Scholar] [CrossRef]
  191. Zhang, Y.; Zheng, Y.; Li, Y.; Wang, L.; Bai, Y.; Zhao, Q.; Xiong, X.; Cheng, Y.; Tang, Z.; Deng, Y.; et al. Tantalum Nitride-Decorated Titanium with Enhanced Resistance to Microbiologically Induced Corrosion and Mechanical Property for Dental Application. PLoS ONE 2015, 10, e0130774. [Google Scholar] [CrossRef] [Green Version]
  192. Jin, W.; Wang, G.; Peng, X.; Li, W.; Qasim, A.M.; Chu, P.K. Tantalum nitride films for corrosion protection of biomedical Mg-Y-RE alloy. J. Alloys Compd. 2018, 764, 947–958. [Google Scholar] [CrossRef]
  193. Hsieh, J.H.; Chiu, C.H.; Li, C.; Wu, W.; Chang, S.Y. Development of anti-wear and anti-bacteria TaN-(Ag,Cu) thin films—A review. Surf. Coat. Technol. 2013, 233, 159–168. [Google Scholar] [CrossRef]
  194. Schmidt, S.; Hänninen, T.; Wissting, J.; Hultman, L.; Goebbels, N.; Santana, A.; Tobler, M.; Högberg, H. SiNxcoatings deposited by reactive high power impulse magnetron sputtering: Process parameters influencing the residual coating stress. J. Appl. Phys. 2017, 121, 171904. [Google Scholar] [CrossRef] [Green Version]
  195. Lal, S.; Caseley, E.A.; Hall, R.M.; Tipper, J.L. Biological Impact of Silicon Nitride for Orthopaedic Applications: Role of Particle Size, Surface Composition and Donor Variation. Sci. Rep. 2018, 8, 9109. [Google Scholar] [CrossRef]
  196. Skjöldebrand, C.; Hulsart-Billström, G.; Engqvist, H.; Persson, C. Si–Fe–C–N Coatings for Biomedical Applications: A Combinatorial Approach. Materials 2020, 13, 2074. [Google Scholar] [CrossRef]
  197. Correa Filho, L.; Schmidt, S.; López, A.; Cogrel, M.; Leifer, K.; Engqvist, H.; Högberg, H.; Persson, C. The Effect of Coating Density on Functional Properties of SiNx Coated Implants. Materials 2019, 12, 3370. [Google Scholar] [CrossRef] [Green Version]
  198. Pettersson, M.; Bryant, M.; Schmidt, S.; Engqvist, H.; Hall, R.M.; Neville, A.; Persson, C. Dissolution behaviour of silicon nitride coatings for joint replacements. Mater. Sci. Eng. C Mater. 2016, 62, 497–505. [Google Scholar] [CrossRef]
  199. Skjöldebrand, C.; Echeverri, E.; Hulsart-Billström, G.; Persson, C. Tailoring the dissolution rate and in vitro cell response of silicon nitride coatings through combinatorial sputtering with chromium and niobium. Biomater. Sci. 2022, 10, 3757–3769. [Google Scholar] [CrossRef]
  200. Cao, Y.-Q.; Zi, T.-Q.; Zhao, X.-R.; Liu, C.; Ren, Q.; Fang, J.-B.; Li, W.-M.; Li, A.-D. Enhanced visible light photocatalytic activity of Fe2O3 modified TiO2 prepared by atomic layer deposition. Sci. Rep. 2020, 10, 1–10. [Google Scholar] [CrossRef]
  201. He, X.; Zhang, G.; Zhang, H.; Hang, R.; Huang, X.; Yao, X.; Zhang, X. Cu and Si co-doped microporous TiO2 coating for osseointegration by the coordinated stimulus action. Appl. Surf. Sci. 2020, 503, 144072. [Google Scholar] [CrossRef]
  202. Hua, Z.; Zhang, C.; Xu, Y.; Jia, H.; Chen, X.; Bai, Y.; Wei, B. Efficiently reduced heat rise in TiO2 coating Ti-based metallic implants using anodic oxidation method. Surf. Coat. Technol. 2019, 363, 75–79. [Google Scholar] [CrossRef]
  203. Jin, W.; Wang, G.; Lin, Z.; Feng, H.; Li, W.; Peng, X.; Qasim, A.M.; Chu, P.K. Corrosion resistance and cytocompatibility of tantalum-surface-functionalized biomedical ZK60 Mg alloy. Corros. Sci. 2017, 114, 45–56. [Google Scholar] [CrossRef]
  204. Huang, H.-L.; Chang, Y.-Y.; Chen, H.-J.; Chou, Y.-K.; Lai, C.-H.; Chen, M.Y.C. Antibacterial properties and cytocompatibility of tantalum oxide coatings with different silver content. J. Vac. Sci. Technol. A 2014, 32, 02B117. [Google Scholar] [CrossRef]
  205. Xu, N.; Fu, J.; Zhao, L.; Chu, P.K.; Huo, K. Biofunctional Elements Incorporated Nano/Microstructured Coatings on Titanium Implants with Enhanced Osteogenic and Antibacterial Performance. Adv. Healthc. Mater. 2020, 9, 2000681. [Google Scholar] [CrossRef]
  206. Zhao, Q.-M.; Sun, Y.-Y.; Wu, C.-S.; Yang, J.; Bao, G.-F.; Cui, Z.-M. Enhanced osteogenic activity and antibacterial ability of manganese–titanium dioxide microporous coating on titanium surfaces. Nanotoxicology 2020, 14, 289–309. [Google Scholar] [CrossRef]
  207. Zhu, Y.; Zheng, T.; Wen, L.-M.; Li, R.; Zhang, Y.-B.; Bi, W.-J.; Feng, X.-J.; Qi, M.-C. Osteogenic capability of strontium and icariin-loaded TiO2 nanotube coatings in vitro and in osteoporotic rats. J. Biomater. Appl. 2021, 35, 1119–1131. [Google Scholar] [CrossRef]
  208. Zhang, B.; Li, B.; Gao, S.; Li, Y.; Cao, R.; Cheng, J.; Li, R.; Wang, E.; Guo, Y.; Zhang, K.; et al. Y-doped TiO2 coating with superior bioactivity and antibacterial property prepared via plasma electrolytic oxidation. Mater. Des. 2020, 192, 108758. [Google Scholar] [CrossRef]
  209. Li, K.; Liu, S.; Xue, Y.; Zhang, L.; Han, Y. A superparamagnetic Fe3O4–TiO2 composite coating on titanium by micro-arc oxidation for percutaneous implants. J. Mater. Chem. B 2019, 7, 5265–5276. [Google Scholar] [CrossRef]
  210. Gao, Q.; Feng, T.; Huang, D.; Liu, P.; Lin, P.; Wu, Y.; Ye, Z.; Ji, J.; Li, P.; Huang, W. Antibacterial and hydroxyapatite-forming coating for biomedical implants based on polypeptide-functionalized titania nanospikes. Biomater. Sci. 2019, 8, 278–289. [Google Scholar] [CrossRef] [PubMed]
  211. Nagay, B.E.; Dini, C.; Cordeiro, J.M.; Ricomini-Filho, A.P.; de Avila, E.D.; Rangel, E.C.; da Cruz, N.C.; Barao, V.A.R. Visible-Light-Induced Photocatalytic and Antibacterial Activity of TiO2 Codoped with Nitrogen and Bismuth: New Perspectives to Control Implant-Biofilm-Related Diseases. ACS Appl. Mater. Interfaces 2019, 11, 18186–18202. [Google Scholar] [CrossRef] [PubMed]
  212. van Hengel, I.A.J.; Tierolf, M.W.A.M.; Valerio, V.P.M.; Minneboo, M.; Fluit, A.C.; Fratila-Apachitei, L.E.; Apachitei, I.; Zadpoor, A.A. Self-defending additively manufactured bone implants bearing silver and copper nanoparticles. J. Mater. Chem. B 2020, 8, 1589–1602. [Google Scholar] [CrossRef] [Green Version]
  213. Yetim, T. Corrosion Behavior of Ag-doped TiO2 Coatings on Commercially Pure Titanium in Simulated Body Fluid Solution. J. Bionic Eng. 2016, 13, 397–405. [Google Scholar] [CrossRef]
  214. Ferraris, M.; Balagna, C.; Perero, S.; Miola, M.; Ferraris, S.; Baino, F.; Battiato, A.; Manfredotti, C.; Vittone, E.; Verné, E. Silver nanocluster/silica composite coatings obtained by sputtering for antibacterial applications. IOP Conf. Series: Mater. Sci. Eng. 2012, 40, 012037. [Google Scholar] [CrossRef] [Green Version]
  215. Chai, M.; An, M.; Zhang, X. Construction of a TiO2/MoSe2/CHI coating on dental implants for combating Streptococcus mutans infection. Mater. Sci. Eng. C Mater. 2021, 129, 112416. [Google Scholar] [CrossRef]
  216. Han, X.; Zhang, G.; Chai, M.; Zhang, X. Light-assisted therapy for biofilm infected micro-arc oxidation TiO2 coating on bone implants. Biomed. Mater. 2021, 16, 025018. [Google Scholar] [CrossRef]
  217. Zhang, G.; Wu, Z.; Yang, Y.; Shi, J.; Lv, J.; Fang, Y.; Shen, Z.; Lv, Z.; Li, P.; Yao, X.; et al. A multifunctional antibacterial coating on bone implants for osteosarcoma therapy and enhanced osteointegration. Chem. Eng. J. 2022, 428, 131155. [Google Scholar] [CrossRef]
  218. Li, B.; Zhang, L.; Wang, D.; Peng, F.; Zhao, X.; Liang, C.; Li, H.; Wang, H. Thermosensitive -hydrogel-coated titania nanotubes with controlled drug release and immunoregulatory characteristics for orthopedic applications. Mater. Sci. Eng. C Mater. 2021, 122, 111878. [Google Scholar] [CrossRef]
  219. Su, K.; Tan, L.; Liu, X.; Cui, Z.; Zheng, Y.; Li, B.; Han, Y.; Li, Z.; Zhu, S.; Liang, Y.; et al. Rapid Photo-Sonotherapy for Clinical Treatment of Bacterial Infected Bone Implants by Creating Oxygen Deficiency Using Sulfur Doping. ACS Nano 2020, 14, 2077–2089. [Google Scholar] [CrossRef]
  220. Leonova, L.A.; Boytsova, E.L.; Pustovalova, A.A. The study of titanium oxynitride coatings solubility deposited by reactive magnetron sputtering. IOP Conf. Series: Mater. Sci. Eng. 2016, 135, 012026. [Google Scholar] [CrossRef] [Green Version]
  221. Pana, I.; Braic, V.; Dinu, M.; Mouele, E.S.M.; Parau, A.C.; Petrik, L.F.; Braic, M. In Vitro Corrosion of Titanium Nitride and Oxynitride-Based Biocompatible Coatings Deposited on Stainless Steel. Coatings 2020, 10, 710. [Google Scholar] [CrossRef]
  222. Banakh, O.; Moussa, M.; Matthey, J.; Pontearso, A.; Cattani-Lorente, M.; Sanjines, R.; Fontana, P.; Wiskott, A.; Durual, S. Sputtered titanium oxynitride coatings for endosseous applications: Physical and chemical evaluation and first bioactivity assays. Appl. Surf. Sci. 2014, 317, 986–993. [Google Scholar] [CrossRef]
  223. Rieder, P.; Scherrer, S.; Filieri, A.; Wiskott, H.W.A.; Durual, S. TiNOx coatings increase human primary osteoblasts proliferation independently of the substrate—A short report. Bio-Med. Mater. Eng. 2012, 22, 277–281. [Google Scholar] [CrossRef] [PubMed]
  224. Durual, S.; Rieder, P.; Garavaglia, G.; Filieri, A.; Cattani-Lorente, M.; Scherrer, S.S.; Wiskott, H.W.A. TiNOx coatings on roughened titanium and CoCr alloy accelerate early osseointegration of dental implants in minipigs. Bone 2013, 52, 230–237. [Google Scholar] [CrossRef] [PubMed]
  225. Nikolova, M.; Ormanova, M.; Nikolova, V.; Apostolova, M.D. Electrochemical, Tribological and Biocompatible Performance of Electron Beam Modified and Coated Ti6Al4V Alloy. Int. J. Mol. Sci. 2021, 22, 6369. [Google Scholar] [CrossRef]
  226. Petrov, P.; Dechev, D.; Ivanov, N.; Hikov, T.; Valkov, S.; Nikolova, M.; Yankov, E.; Parshorov, S.; Bezdushnyi, R.; Andreeva, A. Study of the influence of electron beam treatment of Ti5Al4V substrate on the mechanical properties and surface topography of multilayer TiN/TiO2 coatings. Vacuum 2018, 154, 264–271. [Google Scholar] [CrossRef]
  227. Valkov, S.; Parshorov, S.; Andreeva, A.; Bezdushnyi, R.; Nikolova, M.; Dechev, D.; Ivanov, N.; Petrov, P. Influence of Electron Beam Treatment of Co–Cr Alloy on the Growing Mechanism, Surface Topography, and Mechanical Properties of Deposited TiN/TiO2 Coatings. Coatings 2019, 9, 513. [Google Scholar] [CrossRef] [Green Version]
  228. Nikolova, M.P.; Valkov, S.; Yankov, E.; Petrov, P. Composition, structure and adhesion of cathodic-arc and glow-discharge deposited TiN/TiO2 coatings: A case study. J. Phys. Conf. Ser. 2020, 1492, 012037. [Google Scholar] [CrossRef]
  229. Kaliaraj, G.S.; Kumar, N. Oxynitrides decorated 316L SS for potential bioimplant application. Mater. Res. Express 2018, 5, 036403. [Google Scholar] [CrossRef]
  230. Rtimi, S.; Baghriche, O.; Sanjines, R.; Pulgarin, C.; Bensimon, M.; Kiwi, J. TiON and TiON-Ag sputtered surfaces leading to bacterial inactivation under indoor actinic light. J. Photochem. Photobiol. A Chem. 2013, 256, 52–63. [Google Scholar] [CrossRef] [Green Version]
  231. Chang, Y.-Y.; Huang, H.-L.; Chen, Y.-C.; Weng, J.-C.; Lai, C.-H. Characterization and antibacterial performance of ZrNO–Ag coatings. Surf. Coat. Technol. 2013, 231, 224–228. [Google Scholar] [CrossRef]
  232. Castro, J.D.; Lima, M.J.; Carvalho, I.; Henriques, M.; Carvalho, S. Cu oxidation mechanism on Cu-Zr(O)N coatings: Role on functional properties. Appl. Surf. Sci. 2021, 555, 149704. [Google Scholar] [CrossRef]
  233. Chen, X.; Zhang, W. Diamond nanostructures for drug delivery, bioimaging, and biosensing. Chem. Soc. Rev. 2017, 46, 734–760. [Google Scholar] [CrossRef] [Green Version]
  234. Kulisch, W.; Popov, C. On the growth mechanisms of nanocrystalline diamond films. Phys. Status Solidi A 2006, 203, 203–219. [Google Scholar] [CrossRef]
  235. Serino, G.; Gusmini, M.; Audenino, A.L.; Bergamasco, G.; Ieropoli, O.; Bignardi, C. Multiscale Characterization of Isotropic Pyrolytic Carbon Used for Mechanical Heart Valve Production. Processes 2021, 9, 338. [Google Scholar] [CrossRef]
  236. Adkinson, J.M.; Chung, K.C. Advances in Small Joint Arthroplasty of the Hand. Plast. Reconstr. Surg. 2014, 134, 1260–1268. [Google Scholar] [CrossRef] [Green Version]
  237. Miksovsky, J.; Voss, A.; Kozarova, R.; Kocourek, T.; Pisarik, P.; Ceccone, G.; Kulisch, W.; Jelinek, M.; Apostolova, M.D.; Reithmaier, J.P.; et al. Cell adhesion and growth on ultrananocrystalline diamond and diamond-like carbon films after different surface modifications. Appl. Surf. Sci. 2014, 297, 95–102. [Google Scholar] [CrossRef]
  238. Granek, A.; Monika, M.; Ozimina, D. Diamond-like carbon films for use in medical implants. AIP Conf. Proc. 2018, 2017, 020006. [Google Scholar] [CrossRef]
  239. Hajduga, M.B.; Bobiński, R. TiN, ZrN and DLC nanocoatings—A comparison of the effects on animals, in-vivo study. Mater. Sci. Eng. C Mater. 2019, 104, 109949. [Google Scholar] [CrossRef]
  240. Hauert, R.; Thorwarth, K.; Thorwarth, G. An overview on diamond-like carbon coatings in medical applications. Surf. Coat. Technol. 2013, 233, 119–130. [Google Scholar] [CrossRef]
  241. Chen, Y.; Nie, X.; Leyland, A.; Housden, J.; Matthews, A. Substrate and bonding layer effects on performance of DLC and TiN biomedical coatings in Hank’s solution under cyclic impact–sliding loads. Surf. Coat. Technol. 2013, 237, 219–229. [Google Scholar] [CrossRef]
  242. Choudhury, D.; Lackner, J.; Fleming, R.A.; Goss, J.; Chen, J.; Zou, M. Diamond-like carbon coatings with zirconium-containing interlayers for orthopedic implants. J. Mech. Behav. Biomed. Mater. 2017, 68, 51–61. [Google Scholar] [CrossRef] [PubMed]
  243. Yonezawa, K.; Kawaguchi, M.; Kaneuji, A.; Ichiseki, T.; Iinuma, Y.; Kawamura, K.; Shintani, K.; Oda, S.; Taki, M.; Kawahara, N. Evaluation of Antibacterial and Cytotoxic Properties of a Fluorinated Diamond-Like Carbon Coating for the Development of Antibacterial Medical Implants. Antibiotics 2020, 9, 495. [Google Scholar] [CrossRef] [PubMed]
  244. Bociaga, D.; Sobczyk-Guzenda, A.; Komorowski, P.; Balcerzak, J.; Jastrzebski, K.; Przybyszewska, K.; Kaczmarek, A. Surface Characteristics and Biological Evaluation of Si-DLC Coatings Fabricated Using Magnetron Sputtering Method on Ti6Al7Nb Substrate. Nanomaterials 2019, 9, 812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Bociaga, D.; Sobczyk-Guzenda, A.; Szymanski, W.; Jedrzejczak, A.; Jastrzebska, A.; Olejnik, A.; Swiatek, L.; Jastrzebski, K. Diamond like carbon coatings doped by Si fabricated by a multi-target DC-RF magnetron sputtering method—Mechanical properties, chemical analysis and biological evaluation. Vacuum 2017, 143, 395–406. [Google Scholar] [CrossRef]
  246. Wachesk, C.C.; Seabra, S.H.; Dos Santos, T.A.T.; Trava-Airoldi, V.J.; Lobo, A.O.; Marciano, F.R. In vivo biocompatibility of diamond-like carbon films containing TiO2 nanoparticles for biomedical applications. J. Mater. Sci. Mater. Med. 2021, 32, 1–10. [Google Scholar] [CrossRef]
  247. Park, S.J.; Lee, K.-R.; Ahn, S.-H.; Kim, J.-G. Instability of diamond-like carbon (DLC) films during sliding in aqueous environment. Diam. Relat. Mater. 2008, 17, 247–251. [Google Scholar] [CrossRef]
  248. Van Der Laan, K.J.; Hasani, M.; Zheng, T.; Schirhagl, R. Nanodiamonds for In Vivo Applications. Small 2018, 14, e1703838. [Google Scholar] [CrossRef]
  249. Xue, Y.; Feng, X.; Roberts, S.C.; Chen, X. Diamond and carbon nanostructures for biomedical applications. Funct. Diam. 2021, 1, 221–242. [Google Scholar] [CrossRef]
  250. Nistor, P.A.; May, P.W. Diamond thin films: Giving biomedical applications a new shine. J. R. Soc. Interface 2017, 14, 20170382. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Ivanova, L.; Popov, C.; Kolev, I.; Shivachev, B.; Karadjov, J.; Tarassov, M.; Kulisch, W.; Reithmaier, J.P.; Apostolova, M.D. Nanocrystalline diamond containing hydrogels and coatings for acceleration of osteogenesis. Diam. Relat. Mater. 2011, 20, 165–169. [Google Scholar] [CrossRef]
  252. Pareta, R.; Yang, L.; Kothari, A.; Sirinrath, S.; Xiao, X.; Sheldon, B.W.; Webster, T.J. Tailoring nanocrystalline diamond coated on titanium for osteoblast adhesion. J. Biomed. Mater. Res. Part A 2010, 95A, 129–136. [Google Scholar] [CrossRef] [PubMed]
  253. Merker, D.; Handzhiyski, Y.; Merz, R.; Kopnarski, M.; Reithmaier, J.P.; Popov, C.; Apostolova, M.D. Influence of surface termination of ultrananocrystalline diamond films coated on titanium on response of human osteoblast cells: A proteome study. Mater. Sci. Eng. C Mater. 2021, 128, 112289. [Google Scholar] [CrossRef]
  254. Duailibi, S.E.; Duailibi, M.T.; Ferreira, L.M.; Salmazi, K.I.L.C.; Salvadori, M.C.; Teixeira, F.D.S.; Pasquarelli, A.; Vacanti, J.P.; Yelick, P.C. Tooth Tissue Engineering: The Influence of Hydrophilic Surface on Nanocrystalline Diamond Films for Human Dental Stem Cells. Tissue Eng. Part A 2013, 19, 2537–2543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Grausova, L.; Kromka, A.; Burdikova, Z.; Eckhardt, A.; Rezek, B.; Vacik, J.; Haenen, K.; Lisa, V.; Bacakova, L. Enhanced Growth and Osteogenic Differentiation of Human Osteoblast-Like Cells on Boron-Doped Nanocrystalline Diamond Thin Films. PLoS ONE 2011, 6, e20943. [Google Scholar] [CrossRef] [Green Version]
  256. Hristova, K.; Pecheva, E.; Pramatarova, L.; Altankov, G. Improved interaction of osteoblast-like cells with apatite–nanodiamond coatings depends on fibronectin. J. Mater. Sci. Mater. Med. 2011, 22, 1891–1900. [Google Scholar] [CrossRef]
  257. Medina, O.; Nocua, J.; Mendoza, F.; Gómez-Moreno, R.; Ávalos, J.; Rodríguez, C.; Morell, G. Bactericide and bacterial anti-adhesive properties of the nanocrystalline diamond surface. Diam. Relat. Mater. 2012, 22, 77–81. [Google Scholar] [CrossRef] [Green Version]
  258. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO). Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef]
  259. Khan, Z.U.; Kausar, A.; Ullah, H.; Badshah, A.; Khan, W.U. A review of graphene oxide, graphene buckypaper, and polymer/graphene composites: Properties and fabrication techniques. J. Plast. Film Sheeting 2016, 32, 336–379. [Google Scholar] [CrossRef]
  260. Yao, J.; Wang, H.; Chen, M.; Yang, M. Recent advances in graphene-based nanomaterials: Properties, toxicity and applications in chemistry, biology and medicine. Microchim. Acta 2019, 186, 395. [Google Scholar] [CrossRef] [PubMed]
  261. Malavika, C.; Thomas, P.; Jose, A.J. Graphene: Promising nanoplatform for biomedical applications. In Materials for Biomedical Engineering: Bioactive Materials, Properties, and Applications; Elsevier: Amsterdam, The Netherlands, 2019; pp. 307–322. [Google Scholar] [CrossRef]
  262. Asgar, H.; Deen, K.M.; Rahman, Z.U.; Shah, U.H.; Raza, M.A.; Haider, W. Functionalized graphene oxide coating on Ti6Al4V alloy for improved biocompatibility and corrosion resistance. Mater. Sci. Eng. C Mater. 2019, 94, 920–928. [Google Scholar] [CrossRef]
  263. Catt, K.; Li, H.; Hoang, V.; Beard, R.; Cui, X.T. Self-powered therapeutic release from conducting polymer/graphene oxide films on magnesium. Nanomedicine: Nanotechnology, Biol. Med. 2018, 14, 2495–2503. [Google Scholar] [CrossRef] [PubMed]
  264. Selim, M.S.; El-Safty, S.A.; Abbas, M.A.; Shenashen, M.A. Facile design of graphene oxide-ZnO nanorod-based ternary nanocomposite as a superhydrophobic and corrosion-barrier coating. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 611, 125793. [Google Scholar] [CrossRef]
  265. Rashad, M.; Pan, F.; Tang, A.; Asif, M. Effect of Graphene Nanoplatelets addition on mechanical properties of pure aluminum using a semi-powder method. Prog. Nat. Sci. Mater. Int. 2014, 24, 101–108. [Google Scholar] [CrossRef] [Green Version]
  266. Suo, L.; Jiang, N.; Wang, Y.; Wang, P.; Chen, J.; Pei, X.; Wang, J.; Wan, Q. The enhancement of osseointegration using a graphene oxide/chitosan/hydroxyapatite composite coating on titanium fabricated by electrophoretic deposition. J. Biomed. Mater. Res. Part B Appl. Biomater. 2019, 107, 635–645. [Google Scholar] [CrossRef]
  267. Nair, M.; Nancy, D.; Krishnan, A.G.; Anjusree, G.S.; Vadukumpully, S.; Nair, S.V. Graphene oxide nanoflakes incorporated gelatin–hydroxyapatite scaffolds enhance osteogenic differentiation of human mesenchymal stem cells. Nanotechnology 2015, 26, 161001. [Google Scholar] [CrossRef]
  268. Agarwalla, S.V.; Ellepola, K.; da Costa, M.C.F.; Fechine, G.J.M.; Morin, J.L.P.; Neto, A.H.C.; Seneviratne, C.J.; Rosa, V. Hydrophobicity of graphene as a driving force for inhibiting biofilm formation of pathogenic bacteria and fungi. Dent. Mater. 2019, 35, 403–413. [Google Scholar] [CrossRef]
  269. Shao, W.; Liu, X.; Min, H.; Dong, G.; Feng, Q.; Zuo, S. Preparation, Characterization, and Antibacterial Activity of Silver Nanoparticle-Decorated Graphene Oxide Nanocomposite. ACS Appl. Mater. Interfaces 2015, 7, 6966–6973. [Google Scholar] [CrossRef]
  270. Arbo, M.D.; Altknecht, L.F.; Cattani, S.; Braga, W.V.; Peruzzi, C.P.; Cestonaro, L.V.; Göethel, G.; Durán, N.; Garcia, S.C. In vitro cardiotoxicity evaluation of graphene oxide. Mutat. Res. Gen. Toxicol. Environ. Mutagen. 2019, 841, 8–13. [Google Scholar] [CrossRef]
  271. Liu, G.; Tang, S.; Li, D.; Hu, J. Self-adjustment of calcium phosphate coating on micro-arc oxidized magnesium and its influence on the corrosion behaviour in simulated body fluids. Corros. Sci. 2014, 79, 206–214. [Google Scholar] [CrossRef]
  272. Kaliaraj, G.S.; Siva, T.; Ramadoss, A. Surface functionalized bioceramics coated on metallic implants for biomedical and anticorrosion performance—A review. J. Mater. Chem. B 2021, 9, 9433–9460. [Google Scholar] [CrossRef] [PubMed]
  273. Al Amin, M.; Rani, A.M.A.; Aliyu, A.A.A.; Bryant, M.G.; Danish, M.; Ahmad, A. Bio-ceramic coatings adhesion and roughness of biomaterials through PM-EDM: A comprehensive review. Mater. Manuf. Proc. 2020, 35, 1157–1180. [Google Scholar] [CrossRef]
  274. Jeong, J.; Kim, J.H.; Shim, J.H.; Hwang, N.S.; Heo, C.Y. Bioactive calcium phosphate materials and applications in bone regeneration. Biomater. Res. 2019, 23, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Kim, S.R.; Lee, J.H.; Kim, Y.T.; Riu, D.H.; Jung, S.J.; Lee, Y.J.; Chung, S.C.; Kim, Y.H. Synthesis of Si,Mg substituted hydroxyapatites and their sintering behaviors. Biomaterials 2003, 24, 1389–1398. [Google Scholar] [CrossRef]
  276. Sankar, M.; Suwas, S.; Balasubramanian, S.; Manivasagam, G. Comparison of electrochemical behavior of hydroxyapatite coated onto WE43 Mg alloy by electrophoretic and pulsed laser deposition. Surf. Coat. Technol. 2017, 309, 840–848. [Google Scholar] [CrossRef]
  277. Melero, H.C.; Sakai, R.T.; Vignatti, C.A.; Benedetti, A.V.; Fernández, J.; Guilemany, J.M.; Suegama, P.H. Corrosion Resistance Evaluation of HVOF Produced Hydroxyapatite and TiO2-hydroxyapatite Coatings in Hanks’ Solution. Mater. Res. 2018, 21, e20170210. [Google Scholar] [CrossRef] [Green Version]
  278. Farrokhi-Rad, M. Electrophoretic deposition of fiber hydroxyapatite/titania nanocomposite coatings. Ceram. Int. 2018, 44, 622–630. [Google Scholar] [CrossRef]
  279. Evcin, A.; Buyukleblebici, B. Ti6Al4V Coating with B2O3 and Al2O3 containing Hydroxyapatite by HVOF Technique. Sci. Iran. 2019, 26, 1980–1989. [Google Scholar] [CrossRef] [Green Version]
  280. Yang, H.; Jia, B.; Zhang, Z.; Qu, X.; Li, G.; Lin, W.; Zhu, D.; Dai, K.; Zheng, Y. Alloying design of biodegradable zinc as promising bone implants for load-bearing applications. Nat. Commun. 2020, 11, 1–16. [Google Scholar] [CrossRef] [Green Version]
  281. Zhou, W.; Hu, Z.; Wang, T.; Yang, G.; Xi, W.; Gan, Y.; Lu, W.; Hu, J. Enhanced corrosion resistance and bioactivity of Mg alloy modified by Zn-doped nanowhisker hydroxyapatite coatings. Colloids Surfaces B Biointerfaces 2019, 186, 110710. [Google Scholar] [CrossRef] [PubMed]
  282. Qiu, X.; Wan, P.; Tan, L.; Fan, X.; Yang, K. Preliminary research on a novel bioactive silicon doped calcium phosphate coating on AZ31 magnesium alloy via electrodeposition. Mater. Sci. Eng. C Mater. 2013, 36, 65–76. [Google Scholar] [CrossRef] [PubMed]
  283. Knight, M.N.; Hankenson, K.D. Mesenchymal Stem Cells in Bone Regeneration. Adv. Wound Care 2013, 2, 306–316. [Google Scholar] [CrossRef] [Green Version]
  284. Cao, J.; Lian, R.; Jiang, X. Magnesium and fluoride doped hydroxyapatite coatings grown by pulsed laser deposition for promoting titanium implant cytocompatibility. Appl. Surf. Sci. 2020, 515, 146069. [Google Scholar] [CrossRef]
  285. Blake, G.M.; Fogelman, I. Long-Term Effect of Strontium Ranelate Treatment on BMD. J. Bone Miner. Res. 2005, 20, 1901–1904. [Google Scholar] [CrossRef] [PubMed]
  286. Shi, W.; Zhao, D.P.; Shang, P.; Nie, H.M.; Zhang, Y.; Tang, J.C. Strontium-doped Hydroxyapatite Coatings Deposited on Mg–4Zn Alloy: Physical-chemical Properties and in vitro Cell Response. Rare Metal Mater. Eng. 2018, 47, 2371–2380. [Google Scholar]
  287. Makkar, P.; Kang, H.J.; Padalhin, A.R.; Faruq, O.; Lee, B. In-vitro and in-vivo evaluation of strontium doped calcium phosphate coatings on biodegradable magnesium alloy for bone applications. Appl. Surf. Sci. 2020, 510, 145333. [Google Scholar] [CrossRef]
  288. Batebi, K.; Khazaei, B.A.; Afshar, A. Characterization of sol-gel derived silver/fluor-hydroxyapatite composite coatings on titanium substrate. Surf. Coat. Technol. 2018, 352, 522–528. [Google Scholar] [CrossRef]
  289. Zhang, L. Surface Modification of Titanium by Hydroxyapatite/CaSiO3/Chitosan Porous Bioceramic Coating. Int. J. Electrochem. Sci. 2020, 15, 3616–3626. [Google Scholar] [CrossRef]
  290. Ciobanu, G.; Harja, M. Cerium-doped hydroxyapatite/collagen coatings on titanium for bone implants. Ceram. Int. 2019, 45, 2852–2857. [Google Scholar] [CrossRef]
  291. ISO 13779-2; Implants for Surgery—Hydroxyapatite—Part 2: Thermally Sprayed Coatings of Hydroxyapatite. International Organization for Standardization: Geneva, Switzerland, 2018.
  292. Fiume, E.; Barberi, J.; Verne, 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]
  293. Cannio, M.; Bellucci, D.; Roether, J.A.; Boccaccini, D.N.; Cannillo, V. Bioactive Glass Applications: A Literature Review of Human Clinical Trials. Materials 2021, 14, 5440. [Google Scholar] [CrossRef] [PubMed]
  294. Zhang, M.; Pu, X.; Chen, X.; Yin, G. In-vivo performance of plasma-sprayed CaO–MgO–SiO2-based bioactive glass-ceramic coating on Ti6Al4V alloy for bone regeneration. Heliyon 2019, 5, e02824. [Google Scholar] [CrossRef]
  295. Rojas, O.; Prudent, M.; López, M.E.; Vargas, F.; Ageorges, H. Influence of Atmospheric Plasma Spraying Parameters on Porosity Formation in Coatings Manufactured from 45S5 Bioglass® powder. J. Therm. Spray Technol. 2020, 29, 185–198. [Google Scholar] [CrossRef]
  296. Pachaiappan, R.; Rajendran, S.; Show, P.L.; Manavalan, K.; Naushad, M. Metal/metal oxide nanocomposites for bactericidal effect: A review. Chemosphere 2021, 272, 128607. [Google Scholar] [CrossRef] [PubMed]
  297. Kaviyarasu, K.; Geetha, N.; Kanimozhi, K.; Magdalane, C.M.; Sivaranjani, S.; Ayeshamariam, A.; Kennedy, J.; Maaza, M. In vitro cytotoxicity effect and antibacterial performance of human lung epithelial cells A549 activity of Zinc oxide doped TiO2 nanocrystals: Investigation of bio-medical application by chemical method. Mater. Sci. Eng. C Mater. 2017, 74, 325–333. [Google Scholar] [CrossRef] [PubMed]
  298. Crush, J.; Hussain, A.; Seah, K.T.M.; Khan, W.S. Bioactive Glass: Methods for Assessing Angiogenesis and Osteogenesis. Front. Cell Dev. Biol. 2021, 9, 643781. [Google Scholar] [CrossRef]
  299. Raja, C.A.; Balakumar, S.; Bargavi, P.; Rajashree, P.; Anandkumar, B.; George, R.P.; Mudali, U.K. Decoration of 1-D nano bioactive glass on reduced graphene oxide sheets: Strategies and in vitro bioactivity studies. Mater. Sci. Eng. C Mater. 2018, 90, 85–94. [Google Scholar] [CrossRef]
  300. Raja, C.A.; Balakumar, S.; Anandkumar, B.; George, R.P.; Mudali, U.K. Formation of bioactive nano hybrid thin films on anodized titanium via electrophoretic deposition intended for biomedical applications. Mater. Today Commun. 2020, 25, 101666. [Google Scholar] [CrossRef]
  301. Chalisgaonkar, V.; Das, M.; Balla, V.K. Laser processing of Ti composite coatings reinforced with hydroxyapatite and bioglass. Addit. Manuf. 2018, 20, 134–143. [Google Scholar] [CrossRef]
  302. Say, Y.; Aksakal, B. Enhanced corrosion properties of biological NiTi alloy by hydroxyapatite and bioglass based biocomposite coatings. J. Mater. Res. Technol. 2020, 9, 1742–1749. [Google Scholar] [CrossRef]
  303. Heise, S.; Höhlinger, M.; Hernández, Y.T.; Palacio, J.J.P.; Ortiz, J.A.R.; Wagener, V.; Virtanen, S.; Boccaccini, A.R. Electrophoretic deposition and characterization of chitosan/bioactive glass composite coatings on Mg alloy substrates. Electrochim. Acta 2017, 232, 456–464. [Google Scholar] [CrossRef]
  304. Pozzo, L.D.Y.; da Conceição, T.F.; Spinelli, A.; Scharnagl, N.; Pires, A.T.N. The influence of the crosslinking degree on the corrosion protection properties of chitosan coatings in simulated body fluid. Prog. Org. Coat. 2019, 137, 105328. [Google Scholar] [CrossRef]
  305. Alaei, M.; Atapour, M.; Labbaf, S. Electrophoretic deposition of chitosan-bioactive glass nanocomposite coatings on AZ91 Mg alloy for biomedical applications. Prog. Org. Coatings 2020, 147, 105803. [Google Scholar] [CrossRef]
  306. Singh, S.; Singh, G.; Bala, N. Characterization, electrochemical behavior and in vitro hemocompatibility of hydroxyapatite-bioglass-iron oxide-chitosan composite coating by electrophoretic deposition. Surf. Coat. Technol. 2021, 405, 126564. [Google Scholar] [CrossRef]
  307. Hosseini, S.; Farnoush, H. Characterization and in vitro bioactivity of electrophoretically deposited Mn-modified bioglass-alginate nanostructured composite coatings. Mater. Res. Express 2019, 6, 025404. [Google Scholar] [CrossRef]
  308. Li, M.; Tian, W.; Zhang, Y.; Song, H.; Yu, Y.; Chen, X.; Yong, N.; Li, X.; Yin, Y.; Fan, Q.; et al. Enhanced Silk Fibroin/Sericin Composite Film: Preparation, Mechanical Properties and Mineralization Activity. Polymers 2022, 14, 2466. [Google Scholar] [CrossRef]
  309. Szaraniec, B.; Pielichowska, K.; Pac, E.; Menaszek, E. Multifunctional polymer coatings for titanium implants. Mater. Sci. Eng. C Mater. 2018, 93, 950–957. [Google Scholar] [CrossRef]
  310. Kolev, I.; Ivanova, L.; Markova, L.; Dimitrova, A.; Popov, C.; Apostolova, M.D. Osteoporosis: A Look at the Future. In Osteoporosis; IntechOpen: London, UK, 2012; pp. 667–694. [Google Scholar] [CrossRef]
  311. Ali, U.; Karim, K.J.B.A.; Buang, N.A. A Review of the Properties and Applications of Poly (Methyl Methacrylate) (PMMA). Polym. Rev. 2015, 55, 678–705. [Google Scholar] [CrossRef]
  312. Floroian, L.; Samoila, C.; Badea, M.; Munteanu, D.; Ristoscu, C.; Sima, F.; Negut, I.; Chifiriuc, M.C.; Mihailescu, I.N. Stainless steel surface biofunctionalization with PMMA-bioglass coatings: Compositional, electrochemical corrosion studies and microbiological assay. J. Mater. Sci. Mater. Electron. 2015, 26, 1–14. [Google Scholar] [CrossRef]
  313. Li, X.; Zhitomirsky, I. Deposition of poly(methyl methacrylate) and composites containing bioceramics and bioglass by dip coating using isopropanol-water co-solvent. Prog. Org. Coat. 2020, 148, 105883. [Google Scholar] [CrossRef]
  314. Huang, Z.-B.; Yin, G.-F.; Liao, X.-M.; Gu, J.-W. Conducting polypyrrole in tissue engineering applications. Front. Mater. Sci. 2014, 8, 39–45. [Google Scholar] [CrossRef]
  315. Balint, R.; Cassidy, N.J.; Cartmell, S.H. Conductive polymers: Towards a smart biomaterial for tissue engineering. Acta Biomater. 2014, 10, 2341–2353. [Google Scholar] [CrossRef]
  316. Guo, Y.; Jia, S.; Qiao, L.; Su, Y.; Gu, R.; Li, G.; Lian, J. A multifunctional polypyrrole/zinc oxide composite coating on biodegradable magnesium alloys for orthopedic implants. Colloids Surfaces B Biointerfaces 2020, 194, 111186. [Google Scholar] [CrossRef] [PubMed]
  317. Kumar, A.M.; Adesina, A.Y.; Hussein, M.; Umoren, S.A.; Ramakrishna, S.; Saravanan, S. Preparation and characterization of Pectin/Polypyrrole based multifunctional coatings on TiNbZr alloy for orthopaedic applications. Carbohydr. Polym. 2020, 242, 116285. [Google Scholar] [CrossRef] [PubMed]
  318. Baştan, F.E.; Rehman, M.A.U.; Avcu, Y.Y.; Avcu, E.; Üstel, F.; Boccaccini, A.R. Electrophoretic co-deposition of PEEK-hydroxyapatite composite coatings for biomedical applications. Colloids Surfaces B Biointerfaces 2018, 169, 176–182. [Google Scholar] [CrossRef]
  319. Shafiee, B.M.; Torkaman, R.; Mahmoudi, M.; Emadi, R.; Derakhshan, M.; Karamian, E.; Tavangarian, F. Surface Modification of 316L SS Implants by Applying Bioglass/Gelatin/Polycaprolactone Composite Coatings for Biomedical Applications. Coatings 2020, 10, 1220. [Google Scholar] [CrossRef]
  320. Leng, Y.; Yang, F.; Wang, Q.; Li, Z.; Yuan, B.; Peng, C.; Ren, G.; Wang, Z.; Cui, Y.; Wang, Y.; et al. Material-based therapy for bone nonunion. Mater. Des. 2019, 183, 108161. [Google Scholar] [CrossRef]
  321. Sun, Y.; Zhao, Y.-Q.; Zeng, Q.; Wu, Y.-W.; Hu, Y.; Duan, S.; Tang, Z.; Xu, F.-J. Dual-Functional Implants with Antibacterial and Osteointegration-Promoting Performances. ACS Appl. Mater. Interfaces 2019, 11, 36449–36457. [Google Scholar] [CrossRef]
  322. Gao, F.; Hu, Y.; Gong, Z.; Liu, T.; Gong, T.; Liu, S.; Zhang, C.; Quan, L.; Kaveendran, B.; Pan, C. Fabrication of chitosan/heparinized graphene oxide multilayer coating to improve corrosion resistance and biocompatibility of magnesium alloys. Mater. Sci. Eng. C Mater. 2019, 104, 109947. [Google Scholar] [CrossRef]
  323. Guo, Y.; Yu, Y.; Han, L.; Ma, S.; Zhao, J.; Chen, H.; Yang, Z.; Zhang, F.; Xia, Y.; Zhou, Y. Biocompatibility and osteogenic activity of guided bone regeneration membrane based on chitosan-coated magnesium alloy. Mater. Sci. Eng. C Mater. 2019, 100, 226–235. [Google Scholar] [CrossRef] [PubMed]
  324. Huang, L.; Yi, J.; Gao, Q.; Wang, X.; Chen, Y.; Liu, P. Carboxymethyl chitosan functionalization of CPED-treated magnesium alloy via polydopamine as intermediate layer. Surf. Coat. Technol. 2014, 258, 664–671. [Google Scholar] [CrossRef]
  325. López, A.D.L.R.; Lee, M.-R.; Ortiz, B.J.; Gastfriend, B.D.; Whitehead, R.; Lynn, D.M.; Palecek, S.P. Preventing, S. aureus biofilm formation on titanium surfaces by the release of antimicrobial β-peptides from polyelectrolyte multilayers. Acta Biomater. 2019, 93, 50–62. [Google Scholar] [CrossRef] [PubMed]
  326. Kowalski, P.; Łosiewicz, B.; Goryczka, T. Deposition of Chitosan Layers on NiTi Shape Memory Alloy. Arch. Metall. Mater. 2015, 60, 171–176. [Google Scholar] [CrossRef]
  327. Ball, V. Polydopamine Nanomaterials: Recent Advances in Synthesis Methods and Applications. Front. Bioeng. Biotechnol. 2018, 6, 109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. Kaushik, N.; Nguyen, L.N.; Kim, J.H.; Choi, E.H.; Kaushik, N.K. Strategies for Using Polydopamine to Induce Biomineralization of Hydroxyapatite on Implant Materials for Bone Tissue Engineering. Int. J. Mol. Sci. 2020, 21, 6544. [Google Scholar] [CrossRef]
  329. Peng, F.; Cheng, S.; Zhang, R.; Li, M.; Zhou, J.; Wang, D.; Zhang, Y. Zn-contained mussel-inspired film on Mg alloy for inhibiting bacterial infection and promoting bone regeneration. Regen. Biomater. 2021, 8, rbaa044. [Google Scholar] [CrossRef]
  330. Li, J.; Tan, L.; Liu, X.; Cui, Z.; Yang, X.; Yeung, K.W.K.; Chu, P.K.; Wu, S. Balancing Bacteria–Osteoblast Competition through Selective Physical Puncture and Biofunctionalization of ZnO/Polydopamine/Arginine-Glycine-Aspartic Acid-Cysteine Nanorods. ACS Nano 2017, 11, 11250–11263. [Google Scholar] [CrossRef]
  331. Long, X.; Xu, H.; Zhang, D.; Li, J. Bioinspired by both mussel foot protein and bone sialoprotein: Universal adhesive coatings for the promotion of mineralization and osteogenic differentiation. Polym. Chem. 2020, 11, 4995–5004. [Google Scholar] [CrossRef]
  332. Wang, X.; Xu, K.; Cui, W.; Yang, X.; Maitz, M.F.; Li, W.; Li, X.; Chen, J. Controlled synthesis of mussel-inspired Ag nanoparticle coatings with demonstrated in vitro and in vivo antibacterial properties. Mater. Des. 2021, 208, 109944. [Google Scholar] [CrossRef]
  333. Guo, C.; Cui, W.; Wang, X.; Lu, X.; Zhang, L.; Li, X.; Li, W.; Zhang, W.; Chen, J. Poly-l-lysine/Sodium Alginate Coating Loading Nanosilver for Improving the Antibacterial Effect and Inducing Mineralization of Dental Implants. ACS Omega 2020, 5, 10562–10571. [Google Scholar] [CrossRef] [PubMed]
  334. Xie, K.; Zhou, Z.; Guo, Y.; Wang, L.; Li, G.; Zhao, S.; Liu, X.; Li, J.; Jiang, W.; Wu, S.; et al. Long-Term Prevention of Bacterial Infection and Enhanced Osteoinductivity of a Hybrid Coating with Selective Silver Toxicity. Adv. Healthc. Mater. 2019, 8, e1801465. [Google Scholar] [CrossRef] [PubMed]
  335. Zuluaga-Vélez, A.; Quintero-Martinez, A.; Orozco, L.M.; Sepúlveda-Arias, J.C. Silk fibroin nanocomposites as tissue engineering scaffolds—A systematic review. Biomed. Pharmacother. 2021, 141, 111924. [Google Scholar] [CrossRef] [PubMed]
  336. Lin, K.; Wu, H.; Zhao, C.; Wang, X. Silk fibroin scaffolds: A promising candidate for bone regeneration. Front. Bioeng. Biotechnol. 2022, 10, 1054379. [Google Scholar] [CrossRef]
  337. Rahman, M.; Dutta, N.K.; Choudhury, N.R.; Rahman, M.; Dutta, N.K.; Choudhury, N.R. Microroughness induced biomimetic coating for biodegradation control of magnesium. Mater. Sci. Eng. C Mater. 2020, 121, 111811. [Google Scholar] [CrossRef] [PubMed]
  338. Jabbari, F.; Hesaraki, S.; Houshmand, B.; Jabari, F. The physical, mechanical, and biological properties of silk fibroin/chitosan/reduced graphene oxide composite membranes for guided bone regeneration. J. Biomater. Sci. Polym. Ed. 2019, 30, 1779–1802. [Google Scholar] [CrossRef]
  339. Dong, C.; Lv, Y. Application of Collagen Scaffold in Tissue Engineering: Recent Advances and New Perspectives. Polymers 2016, 8, 42. [Google Scholar] [CrossRef] [Green Version]
  340. Mushahary, D.; Wen, C.; Kumar, J.M.; Lin, J.; Harishankar, N.; Hodgson, P.; Pande, G.; Li, Y. Collagen type-I leads to in vivo matrix mineralization and secondary stabilization of Mg–Zr–Ca alloy implants. Colloids Surfaces B Biointerfaces 2014, 122, 719–728. [Google Scholar] [CrossRef]
  341. Atefyekta, S.; Pihl, M.; Lindsay, C.; Heilshorn, S.C.; Andersson, M. Antibiofilm elastin-like polypeptide coatings: Functionality, stability, and selectivity. Acta Biomater. 2019, 83, 245–256. [Google Scholar] [CrossRef]
  342. Wisdom, E.C.; Zhou, Y.; Chen, C.C.; Tamerler, C.; Snead, M.L. Mitigation of Peri-implantitis by Rational Design of Bifunctional Peptides with Antimicrobial Properties. ACS Biomater. Sci. Eng. 2020, 6, 2682–2695. [Google Scholar] [CrossRef]
  343. Zhang, X.; Geng, H.; Gong, L.; Zhang, Q.; Li, H.; Zhang, X.; Wang, Y.; Gao, P. Modification of the surface of titanium with multifunctional chimeric peptides to prevent biofilm formation via inhibition of initial colonizers. Int. J. Nanomed. 2018, 13, 5361–5375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  344. Chen, L.; Mou, S.; Hou, J.; Fang, H.; Zeng, Y.; Sun, J.; Wang, Z. Simple application of adipose-derived stem cell-derived extracellular vesicles coating enhances cytocompatibility and osteoinductivity of titanium implant. Regen. Biomater. 2021, 8, rbaa038. [Google Scholar] [CrossRef] [PubMed]
  345. Tan, L.; Fu, J.; Feng, F.; Liu, X.; Cui, Z.; Li, B.; Han, Y.; Zheng, Y.; Yeung, K.W.K.; Li, Z.; et al. Engineered probiotics biofilm enhances osseointegration via immunoregulation and anti-infection. Sci. Adv. 2020, 6, eaba5723. [Google Scholar] [CrossRef] [PubMed]
  346. Montazerian, M.; Hosseinzadeh, F.; Migneco, C.; Fook, M.V.L.; Baino, F. Bioceramic coatings on metallic implants: An overview. Ceram. Int. 2022, 48, 8987–9005. [Google Scholar] [CrossRef]
  347. Kates, S.L.; Ackert-Bicknell, C.L. How do bisphosphonates affect fracture healing? Injury 2016, 47, S65–S68. [Google Scholar] [CrossRef] [Green Version]
  348. Li, M.; Wan, P.; Wang, W.; Yang, K.; Zhang, Y.; Han, Y. Regulation of osteogenesis and osteoclastogenesis by zoledronic acid loaded on biodegradable magnesium-strontium alloy. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  349. Lu, Y.; Li, M.; Li, L.; Wei, S.; Hu, X.; Wang, X.; Shan, G.; Zhang, Y.; Xia, H.; Yin, Q. High-activity chitosan/nano hydroxyapatite/zoledronic acid scaffolds for simultaneous tumor inhibition, bone repair and infection eradication. Mater. Sci. Eng. C Mater. 2018, 82, 225–233. [Google Scholar] [CrossRef]
  350. Moshiri, A.; Sharifi, A.M.; Oryan, A. Role of Simvastatin on fracture healing and osteoporosis: A systematic review on in vivo investigations. Clin. Exp. Pharmacol. Physiol. 2016, 43, 659–684. [Google Scholar] [CrossRef]
  351. Qi, H.; Heise, S.; Zhou, J.; Schuhladen, K.; Yang, Y.; Cui, N.; Dong, R.; Virtanen, S.; Chen, Q.; Boccaccini, A.R.; et al. Electrophoretic Deposition of Bioadaptive Drug Delivery Coatings on Magnesium Alloy for Bone Repair. ACS Appl. Mater. Interfaces 2019, 11, 8625–8634. [Google Scholar] [CrossRef]
  352. Nyan, M.; Sato, D.; Kihara, H.; Machida, T.; Ohya, K.; Kasugai, S. Effects of the combination with α-tricalcium phosphate and simvastatin on bone regeneration. Clin. Oral Implants Res. 2009, 20, 280–287. [Google Scholar] [CrossRef]
  353. Chissas, D.; Stamatopoulos, G.; Verettas, D.; Kazakos, K.; Papalois, A.; Agrogiannis, G.; Papaeliou, A.; Agapitos, E.; Balanika, A.; Papadopoulou, E.; et al. Can low doses of simvastatin enhance fracture healing? An experimental study in rabbits. Injury 2010, 41, 687–692. [Google Scholar] [CrossRef] [PubMed]
  354. Haidar, Z.S.; Hamdy, R.C.; Tabrizian, M. Delivery of recombinant bone morphogenetic proteins for bone regeneration and repair. Part B: Delivery systems for BMPs in orthopaedic and craniofacial tissue engineering. Biotechnol. Lett. 2009, 31, 1825–1835. [Google Scholar] [CrossRef]
  355. Teng, F.-Y.; Tai, I.-C.; Ho, M.-L.; Wang, J.-W.; Weng, L.W.; Wang, Y.J.; Wang, M.-W.; Tseng, C.-C. Controlled release of BMP-2 from titanium with electrodeposition modification enhancing critical size bone formation. Mater. Sci. Eng. C Mater. 2019, 105, 109879. [Google Scholar] [CrossRef] [PubMed]
  356. Kim, S.-Y.; Kim, Y.-K.; Kim, K.-S.; Lee, K.-B.; Lee, M.-H. Enhancement of bone formation on LBL-coated Mg alloy depending on the different concentration of BMP-2. Colloids Surfaces B Biointerfaces 2019, 173, 437–446. [Google Scholar] [CrossRef] [PubMed]
  357. Zhu, Z.; Li, X.; Li, Y.; Zhu, L.; Zhu, C.; Che, Z.; Huang, L. Three-dimensionally printed porous biomimetic composite for sustained release of recombinant human bone morphogenetic protein 9 to promote osteointegration. Mater. Des. 2021, 208, 109882. [Google Scholar] [CrossRef]
  358. Liu, C.; Wang, J.; Gao, C.; Wang, Z.; Zhou, X.; Tang, M.; Yu, K.; Deng, Y. Enhanced osteoinductivity and corrosion resistance of dopamine/gelatin/rhBMP-2–coated β-TCP/Mg-Zn orthopedic implants: An in vitro and in vivo study. PLoS ONE 2020, 15, e0228247. [Google Scholar] [CrossRef] [Green Version]
  359. Jiang, Y.; Wang, B.; Jia, Z.; Lu, X.; Fang, L.; Wang, K.; Ren, F. Polydopamine mediated assembly of hydroxyapatite nanoparticles and bone morphogenetic protein-2 on magnesium alloys for enhanced corrosion resistance and bone regeneration. J. Biomed. Mater. Res. Part A 2017, 105, 2750–2761. [Google Scholar] [CrossRef]
  360. James, A.W.; Lachaud, G.; Shen, J.; Asatrian, G.; Nguyen, V.; Zhang, X.; Ting, K.; Soo, C. A Review of the Clinical Side Effects of Bone Morphogenetic Protein-2. Tissue Eng. Part B: Rev. 2016, 22, 284–297. [Google Scholar] [CrossRef]
  361. Zhou, Z.; Shi, Q.; Wang, J.; Chen, X.; Hao, Y.; Zhang, Y.; Wang, X. The unfavorable role of titanium particles released from dental implants. Nanotheranostics 2021, 5, 321–332. [Google Scholar] [CrossRef]
  362. Li, Q.; Liang, B.; Wang, F.; Wang, Z. Delivery of Interleukin 4 from a Titanium Substrate Coated with Graphene Oxide for Enhanced Osseointegration by Regulating Macrophage Polarization. ACS Biomater. Sci. Eng. 2020, 6, 5215–5229. [Google Scholar] [CrossRef]
  363. Liu, J.; Tang, Y.; Yang, W.; Tao, B.; He, Y.; Shen, X.; Shen, T.; Lin, C.; Cai, K. Functionalization of titanium substrate with multifunctional peptide OGP-NAC for the regulation of osteoimmunology. Biomater. Sci. 2019, 7, 1463–1476. [Google Scholar] [CrossRef] [PubMed]
  364. Tsuchiya, H.; Shirai, T.; Nishida, H.; Murakami, H.; Kabata, T.; Yamamoto, N.; Watanabe, K.; Nakase, J. Innovative antimicrobial coating of titanium implants with iodine. J. Orthop. Sci. 2012, 17, 595–604. [Google Scholar] [CrossRef] [PubMed]
  365. Kato, T.; Shirai, T.; Yamamoto, N.; Nishida, H.; Hayashi, K.; Takeuchi, A.; Miwa, S.; Ohtani, K.; Tsuchiya, H. Temporal Attenuation of Iodine Content and its Effect on the Antibacterial Activity of Iodine-Supported Titanium Implants. J. Microb. Biochem. Technol. 2016, 8, 285–289. [Google Scholar] [CrossRef]
  366. Kim, T.-S.; Park, S.-H.; Park, D.; Lee, J.-H.; Kang, S. Surface immobilization of chlorhexidine on a reverse osmosis membrane for in-situ biofouling control. J. Membr. Sci. 2019, 576, 17–25. [Google Scholar] [CrossRef]
  367. Wang, S.; Yang, Y.; Li, W.; Wu, Z.; Li, J.; Xu, K.; Zhang, W.; Zheng, X.; Chen, J. Study of the Relationship Between Chlorhexidine-Grafted Amount and Biological Performances of Micro/Nanoporous Titanium Surfaces. ACS Omega 2019, 4, 18370–18380. [Google Scholar] [CrossRef] [Green Version]
  368. Zhou, W.; Peng, X.; Ma, Y.; Hu, Y.; Wu, Y.; Lan, F.; Weir, M.D.; Li, M.; Ren, B.; Oates, T.W.; et al. Two-staged time-dependent materials for the prevention of implant-related infections. Acta Biomater. 2020, 101, 128–140. [Google Scholar] [CrossRef]
  369. Acosta, S.; Ibañez-Fonseca, A.; Aparicio, C.; Rodríguez-Cabello, J.C. Antibiofilm coatings based on protein-engineered polymers and antimicrobial peptides for preventing implant-associated infections. Biomater. Sci. 2020, 8, 2866–2877. [Google Scholar] [CrossRef]
  370. He, Y.; Zhang, Y.; Shen, X.; Tao, B.; Liu, J.; Yuan, Z.; Cai, K. The fabrication and in vitro properties of antibacterial polydopamine-LL-37-POPC coatings on micro-arc oxidized titanium. Colloids Surfaces B Biointerfaces 2018, 170, 54–63. [Google Scholar] [CrossRef]
  371. Tan, L.; Li, J.; Liu, X.; Cui, Z.; Yang, X.; Zhu, S.; Li, Z.; Yuan, X.; Zheng, Y.; Yeung, K.; et al. Rapid Biofilm Eradication on Bone Implants Using Red Phosphorus and Near-Infrared Light. Adv. Mater. 2018, 30, e1801808. [Google Scholar] [CrossRef]
  372. Yuan, Z.; Tao, B.; He, Y.; Mu, C.; Liu, G.; Zhang, J.; Liao, Q.; Liu, P.; Cai, K. Remote eradication of biofilm on titanium implant via near-infrared light triggered photothermal/photodynamic therapy strategy. Biomaterials 2019, 223, 119479. [Google Scholar] [CrossRef]
  373. Song, J.; Liu, H.; Lei, M.; Tan, H.; Chen, Z.; Antoshin, A.; Payne, G.F.; Qu, X.; Liu, C. Redox-Channeling Polydopamine-Ferrocene (PDA-Fc) Coating to Confer Context-Dependent and Photothermal Antimicrobial Activities. ACS Appl. Mater. Interfaces 2020, 12, 8915–8928. [Google Scholar] [CrossRef] [PubMed]
  374. Sang, S.; Guo, G.; Yu, J.; Zhang, X. Antibacterial application of gentamicin–silk protein coating with smart release function on titanium, polyethylene, and Al2O3 materials. Mater. Sci. Eng. C Mater. 2021, 124, 112069. [Google Scholar] [CrossRef] [PubMed]
  375. Tao, B.; Lin, C.; He, Y.; Yuan, Z.; Chen, M.; Xu, K.; Li, K.; Guo, A.; Cai, K.; Chen, L. Osteoimmunomodulation mediating improved osteointegration by OGP-loaded cobalt-metal organic framework on titanium implants with antibacterial property. Chem. Eng. J. 2021, 423, 130176. [Google Scholar] [CrossRef]
Figure 1. A schematic view of the influence of external mechanical forces and local (intrinsic) mechanical cues of the surrounding extracellular matrix on cell behavior of osteoblasts cultured on coated implants. The mechanical cues are transferred from the cell surface to the nucleus by the mechanisms of mechano-responsive signaling including the cytoskeleton. The mechano-responsive transduction activates downstream cell signaling cascades that determine the cell behavior, motility and fate. By a transcriptional feedback loop, the cells can change their mechano-environment by cytoskeletal remodeling and cell migration.
Figure 1. A schematic view of the influence of external mechanical forces and local (intrinsic) mechanical cues of the surrounding extracellular matrix on cell behavior of osteoblasts cultured on coated implants. The mechanical cues are transferred from the cell surface to the nucleus by the mechanisms of mechano-responsive signaling including the cytoskeleton. The mechano-responsive transduction activates downstream cell signaling cascades that determine the cell behavior, motility and fate. By a transcriptional feedback loop, the cells can change their mechano-environment by cytoskeletal remodeling and cell migration.
Materials 16 00183 g001
Figure 2. Classification of vital properties of biomaterials.
Figure 2. Classification of vital properties of biomaterials.
Materials 16 00183 g002
Figure 3. Comparison of the elastic moduli, density, yield strength and fracture toughness of different metallic biomaterials compared to those values in the cortical bone.
Figure 3. Comparison of the elastic moduli, density, yield strength and fracture toughness of different metallic biomaterials compared to those values in the cortical bone.
Materials 16 00183 g003
Figure 4. Schematic view of preparation of sol-gel preparation of coatings. Based on the type of drying method, the end product type can vary substantially.
Figure 4. Schematic view of preparation of sol-gel preparation of coatings. Based on the type of drying method, the end product type can vary substantially.
Materials 16 00183 g004
Figure 5. Principle of plasma generation and coating deposition on a metal substrate using plasma spraying.
Figure 5. Principle of plasma generation and coating deposition on a metal substrate using plasma spraying.
Materials 16 00183 g005
Figure 6. Schematic representation of the EPD process for the production of composite coatings on a metallic substrate.
Figure 6. Schematic representation of the EPD process for the production of composite coatings on a metallic substrate.
Materials 16 00183 g006
Figure 7. A scheme of MAO device (A) and growth model presenting the formation of MAO coating with complex pores.
Figure 7. A scheme of MAO device (A) and growth model presenting the formation of MAO coating with complex pores.
Materials 16 00183 g007
Figure 8. Schematic representation of magnetron sputtering PVD process for deposition of thin films.
Figure 8. Schematic representation of magnetron sputtering PVD process for deposition of thin films.
Materials 16 00183 g008
Figure 9. The therapeutic impact of some ions with angiogenic, osteogenic, anti-inflammatory and anti-bacterial activity.
Figure 9. The therapeutic impact of some ions with angiogenic, osteogenic, anti-inflammatory and anti-bacterial activity.
Materials 16 00183 g009
Figure 10. Schematic representation of the changes occurring after EBT of the metal surface and during deposition of the nitride/oxide coatings deposited by different methods.
Figure 10. Schematic representation of the changes occurring after EBT of the metal surface and during deposition of the nitride/oxide coatings deposited by different methods.
Materials 16 00183 g010
Figure 11. Cell adhesion after 6-h residence time. (A), cell viability for 24 h (B) and bone mineralization activity of MG63 osteoblasts after 31 days of incubation (C) with uncoated and TiN/TiO2 coated, polished and EBM samples of Ti64 in as-received (AR) state. Three independent studies of the cell culture samples were performed. Results are averaged and the standard deviation is indicated. * p < 0.05 compared to the polished uncoated AR sample; # p < 0.05 compared to the MS coated sample in the respective series.
Figure 11. Cell adhesion after 6-h residence time. (A), cell viability for 24 h (B) and bone mineralization activity of MG63 osteoblasts after 31 days of incubation (C) with uncoated and TiN/TiO2 coated, polished and EBM samples of Ti64 in as-received (AR) state. Three independent studies of the cell culture samples were performed. Results are averaged and the standard deviation is indicated. * p < 0.05 compared to the polished uncoated AR sample; # p < 0.05 compared to the MS coated sample in the respective series.
Materials 16 00183 g011
Figure 12. Schematic illustration revealing the coating construction and the elimination process of bacteria biofilm through NIR light triggering remote photodynamic and photothermal synergetic treatment.
Figure 12. Schematic illustration revealing the coating construction and the elimination process of bacteria biofilm through NIR light triggering remote photodynamic and photothermal synergetic treatment.
Materials 16 00183 g012
Table 1. Advantages and disadvantages of the coating methods used for deposition of bioactive coatings on metals and alloys developed for medical application.
Table 1. Advantages and disadvantages of the coating methods used for deposition of bioactive coatings on metals and alloys developed for medical application.
ClassificationMethodAdvantagesDisadvantagesRef.
ChemicalSol-gelSimplicity, cost-effectiveness, easy incorporation of active ingredients, various film thicknesses at complex shapesLong processing time, low adhesion, need for post-sintering, the occurrence of cracks[81,82,83,84,87]
Electroless depositionSimple, low-temperature process, suitable for complex geometries, excellent thickness uniformityA lengthy process, microcavities and pitting in the deposits[88,89,90]
Biomimetic depositionLow-temperature process, incorporation of osteogenic agentsNeed for initial pretreatment, long processing time, poor bonding strength[91,92,93]
Spray pyrolysisSimple, cost-effective, low setting-up costs Need for post-deposition annealing[99]
Thermo-chemicalThermal sprayingMicrorough coating surfaces, cost-effective technique, high deposition rateHigh-temperature process; porous, inhomogeneous and amorphous films, low adhesion strength and prone to coating spallation[80,93]
Hydrothermal Simplicity, cost-effectiveness, controllable morphology, high purityLong experimental time[112,170]
CVDUniform film thickness, controlled properties of the deposits, functionalizing of multidimensional surfaceslimited volatile components, generation of toxic by-products, use of toxic precursors, requires expensive equipment[114,115]
ElectrochemicalElectrophoretic depositionShort processing time, porous coatings with controllable thickness, incorporation of bioactive ingredientsLow adhesion strength, need for sintering, the occurrence of cracks, difficult scalability[120,121,122,123]
Electrochemical depositionLow-temperature process, controlled coating structure and morphology, relatively low costsinadequate bonding strength; development of stresses in the coatings; non-homogeneous and non-stoichiometric surfaces[125,127,128,129]
AnodizationVariable nanotube diameters, drug-loading capacity and controlled releaseAnnealing is required, low mechanical stability, the release of toxic ions[130,131,132]
Micro-arc oxidationHighly microporous oxide layers incorporating different compounds from the electrolyte, good adhesion, simplicity, environmentally friendly processLow crystallinity, cracks formation, need for subsequent treatment, corrosion instability[134,138,139]
Physical and electrophysicalPVDEnvironmentally friendly process, the coatings have superior mechanical biocompatible properties, the deposits can be applied to temperature-sensitive substratesHigh equipment costs, inability to incorporate biomolecules[140]
Ion implantationStrong adhesive bondingLimited sample size, need for rotation, amorphous film formation, high costs[141,150]
ElectrospinningSimple and flexible technique, production of fibers with various diameters incorporating particles or drugs, slow release of bioactive components, controlled internal porosity, scale-up potentialLow mechanical and bonding strength[157]
Laser claddingHigh deposition rate, good coating adhesion, increased surface roughnessFormation of pores and cracks because of different thermal expansion coefficients of the substrate and coating[152,153]
3D printingfused deposition modeling, semi-solid extrusionFlexibility in terms of materials used, obtaining complex geometric shapes, time efficacy, low production costsCoating swelling, low bonding strength, low long-term instability, applied heat to the drug during manufacturing[160,162,163]
Table 2. Requirements for HAp coatings for biomedical application [291].
Table 2. Requirements for HAp coatings for biomedical application [291].
PropertySpecification
Ca/P ratio1.67–1.76
Heavy metals<50 ppm
Density2.98 g/cm3
Crystallinity>62%
Thickness5–70 μm
AbrasionMass loss < 65 mg at 100 cycles
Tensile strength>50.8 MPa
Shear strength>22 MPa
Table 3. Pros and cons of various bioactive coatings applied for hard tissue applications.
Table 3. Pros and cons of various bioactive coatings applied for hard tissue applications.
Coating MaterialBenefitsShortcomingsRef.
Metal nitrides (TiN, ZrN, TaN, SiN) and oxynitrides (TiNxOy, TiN/TiO2, TiON, ZrON)Acceptable adhesion to metal substrates, high wear and corrosion resistance, biocompatibilityHigh hardness, premature coating failure, formation of flakes[175,177,179,183,196,197]
Metal oxides
(TiO2, Ta2O5)
Good mechanical properties, bioactivity, antibacterial and catalytic activity, long-term stability under the photo- and chemical corrosionBrittleness, low fracture toughness[57,200,201]
Carbon-based
(DLC, nanodiamonds, graphene, GO)
Biocompatibility, stability, good mechanical properties, low coefficient of friction, superior electrochemical properties, antimicrobial propertiesSingle-layered DLC films suffer from high internal stress, delamination in an aqueous environment, low toughness, high sensitivity to ambient conditions;
Nanodiamonds and graphene display hydrophobicity;
GO nanomaterials indicated the generation of ROS, DNA damage and mitochondrial disturbance;
[233,234,237,238,239,240,247,250,251,270]
Calcium phosphates and hydroxyapatiteExceptional biocompatibility, osteo-inductivity, osteoconductive, bioactivityVery brittle, high stiffness, low flexibility, high solubility in aqueous media[78,271,273]
Bioactive glassesClass A bioactivity, no toxic effectsSemicrystalline or amorphous structure, high brittleness, low fracture toughness, interfacial delamination, need for sintering to achieve adequate adhesion to a metallic substrate[292,297,302,346]
Synthetic polymers
(PMMA, PPy, PU, PCL, PGMA)
Inexpensive, biodegradable (PCL, PGMA), chemically stable, good tensile properties and flexural rigidityLow mechanical properties, hydrophobicity, insolubility (PMMA, PU, PPy), inflammatory reactions, inadequate degradation rate, deteriorate bone cell adhesion[57,58,309,311,315,320]
Natural polymers (chitosan, silk fibroin, dopamine)Good cell adhesion, high cell affinityLow adhesion to metals, low durability, insufficient mechanical properties[326,328,336]
ECM proteins/cell coatingsEnhance tissue regeneration of bone, tendon, ligaments and vascular and connective tissue, promoting cell adhesion, probiotic activityLow mechanical strength, difficult sterilization, rapid degradation[78,339,341,345]
Table 4. Typical production techniques used for the deposition of various bioactive coating materials on metallic implants.
Table 4. Typical production techniques used for the deposition of various bioactive coating materials on metallic implants.
Coating MaterialTypeDeposition TechniqueRef.
NitridesTiN on Ti6Al4VIon implantation[147]
TiN and “soft” Ti4N3−x on Ti6Al4V alloyDC magnetron sputtering[181]
TiN coating on Ti20Nb13Zr (TNZ) alloyCathodic arc PVD[182]
Copper-doped TiN (TiCuN) deposited on 316L SSAxial magnetic field enhanced arc ion plating[186]
OxidesZnO and ZnO/Ag on Mg-Ca alloyElectroless deposition[89]
TiO2 layers on Ti sheets and TiO2 nanotubesAtomic layer deposition[116]
Tantalum oxide on Mg alloyReactive magnetron sputtering[203]
Doped-TiO2 coatingsMAO (PEO)[206,208,209,211]
TiO2 nanotubes on pure TiAnodization[207]
TiO2 layer embedding silver (Ag) and zinc (Zn) nanoparticles3D printing[212]
Ag-doped TiO2 coatings on TiSol-gel[213]
Collagen/polydopamine/TiO2 coatings on Ti implantsMAO and hydrothermal treatment[216]
OxynitridesTiN/TiO2 coatings on Ti6Al4VCathodic arc deposition (CAD) and glow discharge oxidation[225]
TiON and ZrON films on SS 316LMagnetron sputtering[229]
Carbon-basedHybrid DLC coatings incorporating TiO2 nanoparticles on AISI 316plasma-enhanced CVD[246]
Nanocrystalline and microcrystalline diamonds on Mo substratesHot filament CVD[257]
Modified ultra-nanocrystalline diamond coatings on TiMicrowaved plasma-assisted CVD and electron-beam evaporation[253]
DLC with Zr-containing interlayersUnbalanced magnetron sputtering[242]
Mg-functionalized GO coating on Ti6Al4VElectrophoretic deposition[262]
Graphene on pure TiLiquid-free technique[268]
Apatite-nanodiamond coating on SSElectrodeposition[256]
GO loaded with interleukin 4 on TiSpraying[362]
Calcium phosphates and hydroxyapatiteUltra-porous HAp on Ti alloySpray pyrolysis[102]
HAp-graphene on Ti substratesCold spraying[107]
HAp on Ti substratesHot isostatic pressing[109]
CaP layer on Mg alloyHydrothermal crystallization[110]
Si-HAp coating on Mg-5Zn-0.3Ca alloyPulse electrodeposition[125]
HAp coatings on WE43 Mg alloyEPD and pulse laser deposition (PLD)[276]
HAp with TiO2 on Ti6Al4V alloyHigh-Velocity Oxygen Fuel (HVOF) spraying[277]
Ca-Sr-P coatings on Mg alloyChemical immersion method[287]
Ag-F-HAp coatings on Ti substrateSol-gel method[288]
Ce-doped HAp/collagen coatings on TiBiomimetic deposition[290]
Bioactive glassesZirconia-incorporated bioactive glass films on pure Ti substratesSpray pyrolysis[100]
Bioactive glass onto ultra-fine-grained Ti substrates Laser cladding[152]
CaO-MgO-SiO2-based bioactive glass-ceramic on Ti6-Al-4V alloyAtmospheric plasma spraying[294]
Bioactive glass nanorods of 45S5 integrated with rGO sheetsSol-gel deposition[299]
Bioactive glass-rGO hybrid thin on TiO2 nanotubesElectrophoretic deposition[300]
Bioglass coatings reinforced with Ti on Ti substrateLaser Engineered Net Shaping (LENS)[301]
Synthetic polymersPoly(sodium styrene sulfonate) on TiWet-chemical method[96]
PU coating with graphene and β-TCP on Ti implantsDip coating[309]
PPy/ZnO composite coating on Mg alloyElectrochemical synthesis[316]
PCL/gelatin coatings on 316 SSElectrospinning[320]
PEEK-HAp composite coating on 316 SSElectrophoretic deposition[318]
PGMA coupled with quaternized polyethyleneimine and alendronate on Ti implantsImmersion treatment[321]
Polymeric gels (natural polymers)Methoxyl pectin and xanthan incorporating indomethacin coatings on SSSol-gel[86]
Chitosan over porous oxide layer on Ti6Al4VElectrodeposition[126]
Chitosan-Mg composite coating on Mg-Gd alloyDip coating[323]
Carboxymethyl chitosan coating on AZ31D alloyImmersion treatment[324]
Chitosan/hyaluronic acid coating on Ti substrateLayer-by-layer synthesis[325]
Silk fibroin layer on HAp/Mg(OH)2 coating on WE43 magnesium alloySpin coating[337]
Simvastatin/gelatin nanospheres/chitosan composite on WE43 magnesium alloyElectrophoretic deposition[351]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nikolova, M.P.; Apostolova, M.D. Advances in Multifunctional Bioactive Coatings for Metallic Bone Implants. Materials 2023, 16, 183. https://doi.org/10.3390/ma16010183

AMA Style

Nikolova MP, Apostolova MD. Advances in Multifunctional Bioactive Coatings for Metallic Bone Implants. Materials. 2023; 16(1):183. https://doi.org/10.3390/ma16010183

Chicago/Turabian Style

Nikolova, Maria P., and Margarita D. Apostolova. 2023. "Advances in Multifunctional Bioactive Coatings for Metallic Bone Implants" Materials 16, no. 1: 183. https://doi.org/10.3390/ma16010183

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