Skip to main content
ACS Omega logoLink to ACS Omega
. 2022 Mar 11;7(11):9088–9107. doi: 10.1021/acsomega.2c00440

Progress in Niobium Oxide-Containing Coatings for Biomedical Applications: A Critical Review

Mir Saman Safavi †,§, F C Walsh , Livia Visai §,∥,*, Jafar Khalil-Allafi †,*
PMCID: PMC8944537  PMID: 35356687

Abstract

graphic file with name ao2c00440_0004.jpg

Typically, pure niobium oxide coatings are deposited on metallic substrates, such as commercially pure Ti, Ti6Al4 V alloys, stainless steels, niobium, TiNb alloy, and Mg alloys using techniques such as sputter deposition, sol–gel deposition, anodizing, and wet plasma electrolytic oxidation. The relative advantages and limitations of these coating techniques are considered, with particular emphasis on biomedical applications. The properties of a wide range of pure and modified niobium oxide coatings are illustrated, including their thickness, morphology, microstructure, elemental composition, phase composition, surface roughness and hardness. The corrosion resistance, tribological characteristics and cell viability/proliferation of the coatings are illustrated using data from electrochemical, wear resistance and biological cell culture measurements. Critical R&D needs for the development of improved future niobium oxide coatings, in the laboratory and in practice, are highlighted.

1. Introduction

Aging, illness, and accidents provide threats to human health. It is essential to seek appropriate materials and devices to aid repair or replacement of damaged body parts over extended periods.1 According to a definition presented by American National Institute of Health, a biomaterial is “a synthetic or natural material or combination of the materials excluding drugs, which is deployed to partially or entirely replace any tissue/organ or maintain or enhance the life quality of the individuals.” This definition excludes gypsum, waxes, finishing materials, etc.2,3 The biomaterials should encourage favorable biological responses in the host tissue.4 Depending on the type of constituent material, biomaterials can be broadly categorized as follows: (i) metallic, e.g., Ti and Ti-based alloys, Mg alloys, Co–Cr, stainless steel, NiTi; (ii) ceramic, e.g., calcium phosphates (CaP); (iii) polymer, e.g., polyurethane (PU); (iv) composite, e.g., CaP-PU; and (v) natural, e.g., skin.514 It is important to appreciate the contributions made by biomaterials to modern medicine. They are used in a variety of clinical applications, including orthopedic, cardiovascular, facial, and dentistry treatments. The total market for bioimplants in 2020 is about 89 million USD, forecasted to grow with a compound annual growth rate (CAGR) of 6.5% during the time period of 2020–2026. Orthopedic implants hold the largest fraction of the global implant market. Over 22 million orthopedic surgeries take place annually worldwide, which is expected to reach about 28 million in 2022.1518

Metallic materials are by far the most commonly employed synthetic bulk implants since they took advantage of high mechano-corrosion properties together with acceptable biological performance. The appropriate strength of these materials offers the opportunity to exploit them as load-bearing implants in dentistry and orthopedics.1925 The first scientific reports on the use of metals as synthetic implantable material traced back to 1895. For over 100 years, the development progress of metal implants has not been stopped so that the focus on controlling the overall properties of these implants to attain better performance has grown rapidly in recent years,26 but several problems limit the successful long-term in vivo use of metallic implants:

  • The presence of a harsh environment in the human body, wear between joints, and stress load cycles that may be repeated millions of times in a lifetime crucially challenge the implantation process. While corrosion of orthopedic implants may be caused by the presence of corrosive species, such as Cl ions, the low pH associated with the presence of F ions around dental implants can dominate during corrosion.27 Although research in this field has placed a high premium on assessing the influence of the corrosive physiological medium on releasing the toxic ions, tribology plays a central role in realizing the concentration of the released ions especially in the knee, hip, joints, etc.28 For instance, it is reported that 6.7 × 1013 to 2.5 × 1015 Co particles per year can be released from implanted Co–Cr material. When evaluating the Co level in the periprosthetic fluid of patients suffered from failed implants, it has been recognized that the Co level surpassed 22 mg/L, at least 4 times higher than in those without complaints. Furthermore, every step walk leads to the release of 106 debris nanoparticles from metalware.29 NiTi, stainless steel, and Ti6Al4 V implants are the other materials that the ion release from their surface greatly threaten the implantation success and patient’s health, by inducing allergic reactions.30 It is important to adopt effective, empirical strategies to increase the corrosion resistance of biomaterials. This not only prolongs the service lifetime of the implant but also avoids inflammation.

  • The mismatch between the elastic modulus of the synthetic implant material and natural bone, which can lead to stress shielding. While the elastic modulus of the natural bone is in the range of 10–40 GPa, Co–Cr alloys and stainless steels show an elastic modulus above 200 GPa. This can substantially decrease the bone density (osteopenia) since the implanted material with higher elastic modulus bears the applied loads.3133

  • Insufficient osseointegration, which refers to the inefficiency of the implanted material in enabling an appropriate platform for apatite to nucleate and grow at the implant/host tissue interface. When the material is implanted in vivo, the present proteins adsorb on the surface of the implant. Hence, the surface condition of the biomaterial plays a major role in osseointegration. Under in vivo conditions, different cells, such as osteoblasts, fibroblasts, and endothelials come into contact with the biomaterial for tissue remodeling. As a brief summary, the more the bioactive surface, the higher the integration between the implant and host tissue occurs. Among the common metallic biomaterials, stainless steels can suffer from a lack of bioactivity.3440

  • Bacterial infection: when the biomaterial is implanted in vivo, in particular for knee joints and dental applications, a competition between the present microorganisms and host cells starts to occupy the largest fraction of the surface of the biomaterial, so-called “race for the surface”. If the microorganisms win this race, bacterial colonies can be formed over the surface, leading to the preprosthetic infection that can damage the local tissues. To avoid the mentioned challenge at an early stage of implantation, and to get higher implantation efficiency, it is necessary to adopt cost-effective, flexible, and feasible strategies.4146

The surface of the implant is the place where the first interactions between implant and host cell/tissue occur. Surface-related features, encompassing topography, energy, and chemical composition, determine the protein adsorption, osseointegration quality, and implant fixation. It is possible to change the final characteristics of the bulk implant by modifying its surface so that the surface, i.e., the initial interaction site, offers quite different functions compared to the bulk implant.4753 To date, a variety of surface finishing techniques, including electrochemical deposition,54 sol–gel,55 sputter deposition,56 spray deposition,57 biomimetic deposition,58 as well as coating materials, such as CaP,59 TiO2,60 ZrO2,61 ZnO,62 Nb2O5,63 etc., have been proposed to induce favorable properties to the surface of the biomaterials. For instance, it is possible to induce bioactivity for superior osseointegration by the application of CaP family coatings or prevent bacterial inflammations through deposition of silver or zinc-containing layers. These strategies can open new windows to develop the new generation of implants that bypass the challenges faced by metal implants and can avoid putting an additional burden on the patient.64 The nanostructured topography of the surface of the biomaterials brings forth promising perspectives since a vast majority of the biological interactions, such as osteoblast adhesion and biomineralization, occur on the nanoscale, and the nanometric surface roughness can appreciably affect the adhesion process.65Figure 1 illustrates a timeline for the development of biomaterials, biomedical application of Nb, and biocompatible Nb2O5 coatings.

Figure 1.

Figure 1

Timeline for the development of biomaterials, biomedical application of Nb, and biocompatible Nb2O5 coatings.6674

The present review paper attempts to discuss the properties, application, and market of niobium and its oxides, putting emphasis on their biomedical properties and applications. The overall characteristics of the pure NbxOy and NbxOy-containing coatings deposited by various surface finishing techniques will also be treated in detail, putting stress on the importance of surface finishing with pure NbxOy and NbxOy-containing layers along with a relationship between the operating parameters and final properties of the resultant films.

2. Important Properties of Niobium and Niobium Oxide

Niobium (Nb) is a refractory metal, possessing high thermal stability, erosion-resistant properties, low vapor pressure, good chemical stability in a variety of hostile environments, excellent corrosion resistance, suitable tribomechanical properties, e.g., high wear resistance, excellent ductility at room temperature, favorable biocompatibility, and low thermal expansion. Nb falls under the category of transition metals, crystallizing in a body-centered cubic (BCC) crystal lattice.7579Table 1 lists some of the physical properties of the Nb.

Table 1. Some of the Physical Properties of the Nb80,81.

physical properties amount
melting point 2447 °C
boiling point 4741 °C
density 8.57 g cm–3
electrical resistivity 15.2 μΩ cm at 273 K
relative atomic mass 92.906

Regarding biological characteristics of Nb, it is a nontoxic and allergy-free metal, indicting acceptable biocompatibility and osteoconductivity.8284 As in the case of titanium (Ti), Nb is known as an “essential metal” since it enables significant chemical stability in the physiological medium. Therefore, Nb is considered as a potential metallic element in fabrication of synthetic implantable materials and devices for soft and hard tissues.85 Thanks to its promising properties, the main focus of research has been to realize the in vitro and in vivo performance of Nb and in particular Nb-containing alloys. Results of a comparative work on the biological performance of Nb, Ti, and stainless steel, as three important and useful synthetic implant materials, showed that there are much more mitochondrial activity and cell proliferation on Nb compared to the others. The highest rate of the in vivo osseointegration of tested implants is also registered for Nb.86,87 To fabricate bulk synthetic implants, it is encouraged to alloy the Nb with other biocompatible materials, such as Ti, rather than using it in its pure form. When alloyed with Ti, Nb can stabilize the β phase, which contributes to decreasing the elastic modulus of Ti, thereby diminishing the risk of the “stress shielding”. Moreover, the positive role of this element in improving corrosion and biological features of the composite implants should not be neglected.72,88,89

Besides technical aspects, cost and availability during materials selection for the fabrication of a synthetic implant can be critical.90 Brazil has the largest Nb reserves, equivalent to approximately 98.53% of total Nb in the world. Nb ore reserves exceed 212 billion tonnes in Brazil, which makes this country the biggest producer of Nb, about 96% of world production. Brazil is followed by Canada and Australia, which contain 1.01% and 0.46% of world’s Nb ores reserves, respectively.9193 Interestingly, Europe is the main consumer of Nb, where 30% of total Nb is consumed. Worldwide, Nb is supplied in various forms, including niobium pentoxide (Nb2O5), niobium carbide (NbC), metal niobium, ferroniobium alloy (FeNb), and lithium niobate (LiNbO3). Furthermore, it is possible to purchase niobium pentachloride (NbCl5) only from Russia. FeNb dominates the global production of Nb-containing materials. FeNb is used for fabrication of high strength steels, occupying about 89% of Nb world production. It is to be noted that the largest application of Nb is corresponded to structure steels, more than 34% of total Nb market. Overall, the total Nb market value is about 1389 USD in 2019, which is anticipated to stand at 1748 USD by the end of 2027.94

The tremendous osteoconductivity, biocompatibility, and corrosion performance of Nb are dominated by the oxide layer formed over its surface. The corrosion properties of the formed oxide layer determine the overall corrosion resistance of the Nb in acidic or neutral environments. Among various oxides, Nb2O5 is the most common type of niobium oxide and can be spontaneously generated over the surface when Nb is exposed to an oxygen-containing environment. It is tightly adherent to the Nb and can fully cover its surface.86,9599 Nb2O5 illustrates high resistance against both corrosion and wear, excellent thermodynamic, thermal, and chemical stabilities, high reflective index, low density, outstanding mechanical strength, and fracture toughness, nontoxicity, hypoallergenic, great biological characteristics, enabling cellular attachment, growth, spreading, proliferation, metabolic activity, collagen synthesis, seriously influencing endothelial cell responses, alkaline phosphatase activity (ALP), and supporting HAp formation in the various physiological media due to the small lattice mismatch (1.1%) between the oxide and mineral HAp phase.73,100110 Due to the above properties, Nb2O5 coatings show significant potential for applications in a variety of industrial applications, such as electrochromic devices, optical windows, solar cells, sensing devices, catalysts, supercapacitors, and biomedical.111,112 Although the initial interests in Nb2O5 and its polymorphs, e.g., NbO6, are traced back to the 1940s, the potential benefits and applications of Nb2O5 have not been fully exploited so far. In the early years, the main industrial applications of this transition metal were as sensors and catalysts, in which Nb2O5 was used in the bulk or film form.113,114 Recently, the biomedical applications of Nb2O5 in bulk and coating form have drastically increased owing to their exceptional biological features. Synthetic orthopedic implants, biosensors, and radiopacifying agent coupled with Portland cement in dentistry and filler for dental adhesive resin are important medical applications of Nb2O5.72,112,115,116

3. Pure NbxOy Coatings

Background

Typically, pure niobium oxide coatings are deposited over a variety of metallic substrates, including commercially pure Ti, Ti6Al4 V alloys, stainless steels (particularly AISI 316L grade), Nb, TiNb alloy, and Mg alloys using techniques such as sputter deposition, sol–gel deposition, anodizing, and wet plasma electrolytic oxidation (PEO).117121

Magnetron-sputtered pure Nb2O5 coatings can be obtained using a Nb metallic target in either pure oxygen117,122 or mixed oxygen + argon atmospheres.123,124 To deposit sol–gel derived Nb2O5 coatings, three types of sols can be prepared using various precursors with specific molar ratios as below: (i) butanol, acetylacetone, and niobium alkoxide;121,125 (ii) niobium ethoxide, iso-propanol, acetylacetone, and polyethylene glycol (PEG);126 and (iii) niobium butoxide, triethanolamine, ethanol, and PEG.127

The anodizing of Nb-containing substrates can be carried out from three types of electrolyte: (i) NaF or Na2HPO4-containing HF(aq);118 (ii) a mixture of Ca(H2PO4)2, Ca(OOCCH3)2, and C10H14N2Na2O8;88 and (iii) a mixture of C2H6O2 and NH4F.128

PEO coating of Nb anode material can be carried out from phosphoric acid and hydrogen peroxide-containing and calcium acetate monohydrate electrolytes. It is also possible to employ a bath composed of a mixture of phosphoric acid and calcium acetate monohydrate.120Table 2 provides the advantages, limitations, and schematic illustration of common techniques employed for achieving pure niobium oxide coatings.

Table 2. Advantages, Limitations, and Schematic Illustration of the Common Techniques Used to Apply Pure Niobium Oxide Coatings129137.

graphic file with name ao2c00440_0003.jpg

The published literature on the evaluation of final characteristics of pure NbxOy coatings has emphasized the influence of the coatings on the overall performance of the underlying implants, drawing relationships between the role of operating parameters and properties of the resultant coating and providing an insight into the biological performance of metal oxide coatings, such as Nb2O5, Ta2O5, and TiO2.105,124,138 The following text will provide a systematic overview of the properties of pure NbxOy coatings.

Morphology and Topography

The surface morphology and topography of the NbxOy deposits are highly dependent on the type of the employed coating technique, applied operating factors, and post-treatments. Among the various niobium oxides, Nb2O5 coatings have attracted increasing attention. Regardless of the coating method, four different types of surface morphologies: conelike,118 nodular,138 granular,91 and spheroidal120 have been observed for Nb2O5 coatings.

Overall, the surface of the pure coatings produced via the radio frequency (RF) magnetron sputtering and reactive sputtering methods are homogeneous, tightly adherent to the substrate, and without visible defects, such as cracks and discontinuities. The constituent particles are uniformly arranged over the surface of the coating.91,117,139,140 The surface topography of the films confirmed the formation of a densely packed surface without surface-related defects.91,117,139,140 It has been claimed that the application of magnetron sputtering can ensure formation of a dense and uniform coating.124 The insertion of nitrogen gas during RF magnetron sputtering can lead to the formation of niobium-oxynitride coating, in which a higher degree of the crystallinity is seen in deposits from a high gas flow rate. This can also give rise to surface roughness. While amorphous coatings are grown in an equiaxial fashion, crystalline deposits tend to be columnar.138

Published literature on the surface morphology of the sol–gel derived films is very contradictory. While some reports demonstrated the formation of the coatings containing a network of pores with different diameters from nanometer to micrometers by sol–gel dip coating,105,127 one study confirmed the formation of a crack-free surface with a low porosity content through sol–gel spin coating even after calcination at 500 °C.126 The aim behind the application of the calcination is to eliminate solvent and residual organics, convert oxides, and achieve a crystalline deposit.126 However, the evaporated solvent can lead to pore generation in the so-called “evaporation induced self-assembly process”, which can significantly decrease the density of the obtained film. It should be noted that the pores remain at the top surface of the film and may have little effect on the mechano-corrosion properties of the coatings.105,127,141 Moreover, the sol–gel derived NbxOy films are susceptible to blistering originated from the difference between the coefficient of thermal expansion (CTE) of the underlying metallic implant and ceramic top layer.126,142 The constituent particles of the sol–gel derived films include both spherical125 and wavy126 morphologies. The formation of some aggregated particles over the surface of the films have also been reported.126 Although the deposited oxide coating increases the surface roughness of the underlying substrate, the degree of increment greatly depends on the calcination temperature. Eisenbrath et al.121 have indicated that the surface roughness of the Nb2O5 coatings increased from 7 to 40 nm when the calcination temperature was raised from 450 to 700 °C. The nanoporous nature of the sol–gel derived coatings is the reason why the deposited film enhances the surface roughness of the substrate. The higher the surface roughness, the superior bioactivity and biocompatibility are obtained,105,143 which will be comprehensively discussed in the following text.

The operating parameters involved in the anodizing process, e.g., time, cell voltage applied, and acid concentration, noticeably affect the surface morphology of the deposited films. The prolonged anodization time from 20 to 90 min results in a change in the morphology of the coating from amorphous to the self-assembled crystalline one, consisting of microcones. Such a microstructure can be emerged when the oxygen ions diffuse through the newly formed oxide layer and interact with the present Nb ions that are present over the underlying substrate. The presence of salts in the acidic electrolyte determines the morphology and porosity content of the formed microstructure.118,144,145 Furthermore, at longer anodizing times, a higher rate of oxide growth is seen.146 It is reported that the deposition of Nb2O5 layer over the Nb foil changes its morphology through formation of some small pores. The content, distribution, and size of the pores crucially depends on the applied potential. While the surface of oxide films fabricated at the potential of 200 V contains 2 μm-sized pores with quite uniform dispersion, those deposited under 250 V show heterogeneous surfaces with three-time larger pores. The applied potential not only varies the surface characteristics but also changes the thickness and surface roughness, where a thicker layer with higher surface roughness can be obtained when a higher potential is applied.88 The type and concentration of acidic electrolyte have significant influence on the surface properties. The lower the HF(aq) concentration leads to the higher oxide layer growth. The as-reached F ions to the surface of Nb metal interact with the interstitial Nb ions to form Nb2O5.146

Contradictory results have been reported on the surface morphology of Nb2O5 coatings obtained by PEO deposition. While Pereira et al.120 showed the formation of the compact surface with spheroidal particles after PEO in phosphoric acid solution, Lokeshkumar et al.147 demonstrated that the deposited layer was porous, consisting of interconnected open pores. It is stated that the chemical composition of the used solution and operating parameters determine the homogeneity of the Nb2O5 deposit. For instance, the film produced from Na3SiO4·12H2O + KOH electrolyte under a final cell voltage of 80 V had a heterogeneous surface containing microcracks, while that deposited from Na3PO4·9H2O + KOH at a final cell voltage of 75 V produced a uniform surface morphology. In general, the use of a silicate bath leads to thicker deposits.147 It is also possible to couple the PEO route with other electrochemical deposition technique, such as electrophoretic deposition (EPD), to incorporate desired second phase to the growing oxide layer147 and apply two-step PEO with various electrolyte composition.120 For example, biocompatible hydroxyapatite (HAp) particles can be incorporated to the PEO-fabricated Nb2O5 film via EPD. The included particles markedly affect the surface properties of the oxide layer with closing the pores and diminishing their size.147Table 3 summarizes the thickness and surface roughness of the NbxOy coatings deposited via different methods on a variety of substrates.

Table 3. Thickness and Surface Roughness of the NbxOy Coatings Deposited via Different Methods on Various Substratesa.

coating method substrate type thickness/nm surface roughness (Ra)/nm refs
magnetron sputtering Ti6Al4 V 210 2.4 ± 0.5 (117)
RF magnetron sputtering cp Ti 121.63   (148)
magnetron sputtering Ti6Al4 V 210 2.4 ± 0.5 (139)
magnetron sputtering Ti6Al4 V 180 16.5 (122)
magnetron sputtering Ti6Al4 V 210   (140)
RF magnetron sputtering 316L stainless steel 688.5–825.9 2.05–3.41 (138)
reactive magnetron sputtering Ti6Al4 V 300–450   (91)
reactive and nonreactive magnetron sputtering AISI 316LVM steel   18.5 (124)
DC magnetron sputtering 316L stainless steel 300 2.94 (123)
sol–gel cp Ti 100 7–40 (121)
sol–gel 316L stainless steel 2–3 × 103 18.6 (126)
sol–gel β-type Ti alloy 160 4–50 (125)
sol–gel AZ31 Mg alloy 2 × 103 26.75 (105)
anodizing Nb 2–13 × 103 130–900 (88)
a

A change in the deposition technique or processing parameters can significantly alter the surface characteristics.

Microstructure

The published results regarding the atomic arrangement of the sputter deposited Nb2O5 coatings show differences. While an enormous majority of the reports indicated that the magnetron sputtered films are amorphous,117,122,139,149 there is a study confirming the generation of a crystalline Nb2O5 coating via magnetron sputtering at 60 °C.148 Deposition at low temperature and a lack of substrate heating during the deposition process are the main reasons explaining the generation of an amorphous Nb2O5 layer.117,122,139 The postheat treatment of the crystalline Nb2O5 coatings at 700 °C for 60 min leads to the emergence of new Nb2O5 and βNb2O5 peaks as well as changing the height and width of existing peaks.148 The oxygen/nitrogen gas flow ratio realizes the atomic arrangement of the sputtered niobium oxynitride film. The amorphous coating changes to the polycrystalline one when the oxygen content exceeds beyond a threshold. In such a condition, film growth is dependent on the oxygen and niobium reaction even if nitrogen flows into the vacuum chamber. The presence of nitrogen led to a niobium oxynitride coating without changing the crystal lattice.138,150

Sol–gel derived coatings can be either amorphous or crystalline depending on the temperature of the must be carefully established. While Pradhan et al.119 showed that the films calcined at 450 °C are amorphous and crystalline transformation occurred above 500 °C, Amaravathy et al.105 approved the deposition of the sol–gel derived crystalline Nb2O5 layer through calcination at 380 °C. The films calcined at about 500 °C are composed of a hybrid amorphous + crystalline lattice. A rise in calcination temperature not only leads to the appearance of some new Nb2O5 peaks but also changes the unit cell geometry of the lattice. For instance, a change in unit cell geometry from hexagonal to orthorhombic is obtained at the calcination temperature above 575 °C.119Table 4 outlines the crystallite size of the sol–gel derived Nb2O5 films as a function of calcination temperature. The crystallite size of the coatings is measured using the Scherrer equation as follows:151153

graphic file with name ao2c00440_m001.jpg 1

where k, λ, β, and θ are the Scherrer constant, wavelength of X-ray, fwhm, and Bragg angle, respectively. The Scherrer constant depends on the crystallite shape. For instance, it is taken to be 0.89 for spherical crystallites.154

Table 4. Crystallite Size of the Sol–Gel Derived Nb2O5 Films as a Function of Calcination Temperature.

type of sol–gel film deposition calcination temperature/°C atomic arrangement/unit cell crystallite size/nm refs
not mentioned 500 mixture of amorphous + crystalline 23 ± 4 (119)
550 crystalline/hexagonal 37 ± 6
575 crystalline/hexagonal 43 ± 3
650 crystalline/orthorhombic 74 ± 3
dip-coating 500 crystalline/hexagonal 30–40 (127)
dip-coating 380 crystalline/monoclinic 48 (105)
spin-coating 500 crystalline/hexagonal 41.49 (126)
spin-coating 550 not mentioned 24 (155)
600 38
650 39
700 58

As seen, all of the sol–gel derived coatings are nanostructured, which provide a larger surface area to interact with the cells improving the cell adhesion and adsorption of proteins.105

The phase structure of the Nb2O5 produced via anodization of Nb substrate is only composed of peaks assigned to the orthorhombic Nb2O5.118

PEO-coated Nb2O5 are crystalline and can be present in both hexagonal and orthorhombic unit cell geometry.120,147 Although the composition of the acidic electrolyte used in the PEO process has no influence on the emerged phases, it can only alter the height and width of the emerged peaks.147

Tribomechanical Behavior

The implanted material should benefit from appropriate tribomechanical properties since it endures normal and shear forces as well as undergoing friction during the service time. This issue is particularly highlighted for dental and orthopedic implants. Thus, the method and composition of the coating should be controlled in such a way that meet the required tribomechanical properties. Also, the bonding strength and durability of the coating should be taken into consideration since the film may detach as a result of the forces endured.156158 Meanwhile, there should be a balance between the elastic modulus of the implant material and natural bone, especially for orthopedic implants. A big mismatch can result in “stress-shielding”, which eventually causes a failure in implantation. For instance, the elastic modulus values of the stainless steel and cortical bone are 180 and 17 GPa, respectively. Such a mismatch can lead to stress-shielding. Surface modification of the implants is needed to satisfy mechanical and chemical requirements.159,160 Nanoindentation, nanoscratching, pull-off tensile testing, wear testing, and corrosion studies can be employed to analyze the tribomechanical characteristics of biocoatings.161164

The application of the Nb2O5 coating by the sputter deposition route can greatly raise the nanohardness and elastic modulus of the metallic implants, in particular the Ti-based ones. It is reported that the nanohardness and elastic modulus of the Ti6Al4 V implant increases from ≈5.5 to ≈8.6 GPa and from 112 to 142 GPa, respectively.117,122,139,140 The sputter deposited Nb2O5 films possess desirable tribological performance as there are no scratches detected on the surface of the coatings after a steel wool test under a load of 0.5 N.117,122,139 Kalisz et al.140 have demonstrated that the nanohardness of the coatings negligibly decreases, about 5%, after a corrosion test in a corrosive medium containing NaCl and KF. The surface morphology of the Nb2O5 coating after a tribocorrosion test, which provides a perspective on the synergistic action of corrosion and wear, in Ringer’s solution under the normal load of 5 N at rotating speed of 5 rpm that applies to a bone pin indicating the formation of some pits together with a sign of an adhesive wear track. In general, it is claimed that the Nb2O5 layer offers tremendous wear resistance; however, it could be degraded when it underwent wear in a corrosive, in vivo environment.124 An SEM micrograph of a Nb2O5 coating after tribocorrosion testing is shown in Figure 2. The sign of adhesive wear is observable for Nb2O5 coating after the tribocorrosion assay.

Figure 2.

Figure 2

SEM micrograph of Nb2O5 coating after tribocorrosion testing.124 Reprinted with permission from ref (124). Copyright 2021 by the authors.

The results of an in vivo assay can enable a more realistic view of the mechanical performance of the coatings. The in vivo removal of torque is a suitable characterization method to reveal whether the osseointegration is sufficiently firm. The sputter deposited Nb2O5 layer shows an acceptable torque value. The value was enhanced with time, confirming the generation of a favorable implant/bone bonding, which can promote bone growth and facilitate mechanical interlocking between the coating and underlying implant.148

The application of Nb2O5 coating via a sol–gel route gives rise to the hardness of the metallic implant. Moreover, the elastic modulus of the layer is much lower than that of underlying metallic substrate due to its porous microstructure, decreasing the risk of stress-shielding effect. The sol–gel derived ceramic layer possesses high bonding strength, qualifying its use in orthopedic applications.126

The adhesion strength between the PEO-fabricated Nb2O5 coating and Nb substrate is reported to be sufficient for biomedical applications, irrespective of the composition of the used electrolyte.120Table 5 summarizes the reported mechanical properties of NbxOy biocoatings.

Table 5. Reported Mechanical Properties of NbxOy Biocoatings.

coating type deposition method characterization method hardness elastic modulus/GPa refs
Nb2O5 sputter deposition nanoindentation 8.64 GPa 142 (117)
Nb2O5 sputter deposition nanoindentation 8.64 GPa   (139)
Nb2O5 sputter deposition nanoindentation 5.58 ± 0.3 GPa   (122)
Nb2O5 sputter deposition nanoindentation 8.64 GPa (before corrosion)   (140)
8.16 GPa (after corrosion in NaCl and KF-containing media)
Nb2O5 sol–gel Vickers microhardness 444 ± 3.2 HV   (126)

Information on tribomechanical properties of the NbxOy biocoatings is scarce in the open literature. For future works, establishing the mechanisms involved in a change in the tribomechanical properties with the application of NbxOy coating, identifying relationships between the processing parameters and tribomechanical performance of the resultant deposit, studying all of the tribomechanical aspects, encompassing hardness, elastic modulus, interfacial strength, and tribocorrosion response of the coatings, should be addressed in detail.

Corrosion Performance

In general, the corrosion resistance of the metallic implants, including Ti6Al4 V and 316L stainless steel, outstandingly promotes when they are protected by sputter deposited Nb2O5 coatings. However, the factors governing such an enhancement in the corrosion behavior have not been clearly considered in the literature.117,122,139,140 In the case of an uncoated metallic implant, e.g., Ti6Al4 V, a passive layer can be formed over its surface when it is immersed in the corrosive medium. Since the Ti has a negative reversible potential, it is susceptible to oxidation upon exposing to aqueous electrolytes. The passive layer contributes to the corrosion resistance through providing an initial barrier against the present corrosive species. It is to be emphasized that the protective capability of the layer directly depends on the pH of the medium so that the layer cannot resist against aggressive anions.165,166 There are practical works that provide a comparative insight into the corrosion protection performance of the ceramic coatings, e.g., Nb2O5, NbN, and TiO2, in the biological media. For instance, Ramirez et al.149 have reported that sputter deposited Nb2O5 provided stronger corrosion protection than that of NbN produced by the same technique. On the other hand, Orozco-Hernández et al.124 have revealed the better corrosion inhibition of TiO2 compared to Nb2O5 without correlating the results with their morphological and microstructural characteristics. Unlike TiO2 films, the sputter deposited Nb2O5 ones are prone to fall into pitting corrosion.

The sol–gel produced Nb2O5 coatings improve the corrosion resistance of the underlying metallic substrates since they can act as a potential barrier against corrosive species.105,125127 Unlike bare substrate, a passive film forms over the coated one when it is immersed into the corrosive electrolyte.105 Although the breakdown potential obtained from the potentiodynamic polarization assay is dependent on the oxygen evolution at substrate/coating interface, by which the corrosive species reach to the substrate via cracks and/or pores passage, the potential values of bare and coated-substrate reveal that the porous nature of the sol–gel derived films did not adversely affect their corrosion resistance. The simulated body fluid (SBF)-soaked Nb2O5 coatings show much higher corrosion resistance than that of as-deposited Nb2O5 due to the fact that the immersion in SBF leads to an apatite layer, which provides an extra blocking barrier and prevents the dissolution of the metal ions.126,127,167 It is well established that the volume of the evolved hydrogen when an implant is immersed in SBF alters the biocompatibility of the surrounding cells. Thus, the corrosion resistant material, which produces lower hydrogen volume in SBF, is more biocompatible. There are much lower hydrogen bubbles emerged over the surface of the Nb2O5-coated AZ31 Mg alloy than bare alloy, contributing to its better biological characteristics.105 The results of a comparative study of the corrosion performance of the several sol–gel derived metal oxide coatings on Ti implants revealed that the Nb2O5 coating has the lowest corrosion resistance without addressing the parameters lead to difference in resistance of the coatings.125

Anodized Nb substrates show better corrosion behavior compared with the bare Nb.88,168 The improvement degree deeply depends on the cell voltage during anodizing. Canepa et al.88 have reported that the corrosion resistance of the anodized Nb linearly increases with the increase in cell voltage from 0 to 200 V, followed by a slight drop with a further rise up to 250 V. While the increased thickness and better morphological properties of the oxide layer leads to enhanced corrosion resistance, a higher porosity content degrades the resistance of the samples produced at 250 V. Results of a comparative work revealed that the formed oxide layer during the anodization of Ta and Nb possesses superior corrosion protection than that of Ti While Nb2O5 coatings deposited by PEO on the Nb substrate promote its corrosion behavior, the enhancement is not marked due to the high porosity of the film. In addition, the thickness and phase composition of the deposit govern the corrosion protection efficiency. The PEO process that performed in silicate-based solution has superior corrosion resistance to that produced from phosphate-based electrolyte may be due to the lower porosity content.147,169Table 6 lists the corrosion parameters of the NbxOy biocoatings deposited by various methods.

Table 6. Corrosion Parameters of the NbxOy Biocoatings Deposited by Various Methodsa.

coating type deposition method test medium highest Ecorr/mV lowest jcorr/μA cm–2 highest Rp/kΩ highest Rct/kΩ ref
Nb2O5 sputter deposition 0.5 M/L NaCl, 2 g/L KF; pH 2 at room temperature –401 0.24     (117)
Nb2O5 sputter deposition 0.5 M/L NaCl, 2 g/L KF; pH 2 at room temperature –401 0.24     (139)
Nb2O5 sputter deposition 0.5 mol/L NaCl, 2 g/L KF; pH 2 at room temperature –251 0.09     (122)
Nb2O5 sputter deposition 0.5 mol/L NaCl, 2 g/L KF; pH 2 at room temperature –401 0.24     (140)
Nb2O5 sputter deposition 8.9 g/L NaCl; pH 7.4 at room temperature   0.12 3200   (149)
Nb2O5 sputter deposition Ringer’s solution, at 37 ± 0.2 °C –170 22.1   52.8 (124)
Nb2O5 sol–gel SBF, at room temperature –202     141.41 (126)
Nb2O5 Sol–gel 0.9% NaCl, pH 7.4 at 37 °C   83 4820   (125)
Nb2O5 sol–gel SBF, pH 7.4 –1586 ± 49 53.2 ± 4     (105)
Nb2O5 anodizing Ringer’s solution, at 37 °C –200   32400 29 (88)
Nb2O5 PEO SBF, pH 7.4 –472.97 0.0218     (147)
a

Ecorr, jcorr, and Rp obtained from potentiodynamic polarization are corrosion potential, corrosion current density, and polarization resistance, respectively. Rct values are charge transfer resistance obtained from electrochemical impedance spectroscopy (EIS).

Biological Characteristics

The biological characteristics of the pure NbxOy coatings, including cell viability, adhesion, proliferation, differentiation, and growth along with the ability of the oxide coatings to provide a favorable substrate for nucleation and growth of stoichiometric HAp upon immersion in the SBF, and antibacterial activity, are of the important aspects that have been addressed in the literature.119,120,149,155

Overall, the published results on the influence of sputter deposited NbxOy coatings on the biological characteristics of underlying implants show pronounced differences. While a vast majority of the results confirmed the improvement of the coating characteristics is related to the type of the underlying substrate. Unlike Nb2O5-coated 316L SS,123 the application of Nb2O5 layer over the Ti6Al4 V degrades the cell adhesion and proliferation. The reason addressing such a different behavior is assigned to the higher hydrophobicity of the Nb2O5 layer than that of Ti6Al4 V.149

The enhanced performance of the coated-implants is attributed to the synergistic role of higher hydrophilic nature, surface roughness, and more polarized structure of the NbxOy. The level of improvement in the biological properties is a function of surface topography, wettability, and atomic order, which can be controlled by operating parameters involved in the sputtering technique. For example, the Nb2O5 coatings sputter deposited under a higher N2/O2 gas ratio offer superior human fibroblast cell adhesion due to their amorphous atomic order.138

As an important factor realizing the biocompatibility of a synthetic implanted material, besides cell compatibility, the material should not lead to a harmful inflammation when it comes in contact with living cells and tissues. Corrosion and abrasion occurrence in the physiological medium can result in the inflammatory due to the released toxic ions from the surface of the material, thereby failing the implantation.170172 IL-10 and IL-17 are of anti-inflammatory cytokines, which play a key role in impeding inflammatory, e.g., chronic inflammatory and autoimmune pathologies. Moreover, TNF-α deals with inflammation prevention and osteoblast activation.173,174 The Nb2O5-coated metallic substrates induce higher anti-inflammatory cytokines, revealing their excellent capability in diminishing the toxicity.91,123

The antibacterial activity of the metallic implant is highly improved with protecting its surface by sputter deposited NbOxNy coating. The antibacterial activity of the coatings against S. aureus is promoted when the atomic order is changed from amorphous to crystalline with tuning the N2/O2 gas ratio.138

The sol–gel derived Nb2O5 deposits benefit from the appropriate biological properties; however, the processing parameters in particular calcination time and temperature markedly alter the properties. Besides, the immersion time can also govern the features of the formed apatite layer.74,119,126,127 The in vitro bioactivity assay by immersing the sample into the SBF solution containing an analogous ionic composition to human blood plasma is a preliminary route to realize whether the sample is bioactive. If the incubated material supports the formation of apatite, it is bioactive. The more compact the apatite layer is formed and chemical composition is closer to that of stoichiometric HAp, i.e., Ca/P = 1.67, the better bioactivity is obtained. The results of SBF soaking assay give a reliable view of in vivo bone bonding, where the sample that cannot stimulate the apatite formation is not qualified for further in vivo assays.175178 Pradhan et al.119 have observed that the platelike apatite crystals are generated over the surface of the sol–gel derived Nb2O5 coating calcined at 525 °C after immersion in SBF for 30 days, while the calcination at lower and higher temperatures, e.g., 450 and 650 °C, as well as immersion for 1–7 days do not stimulate apatite formation. The difference in the bioactivity of the coatings has been ascribed to their various atomic arrangement and unit cell geometry, where the hexagonal-Nb2O5 coating possesses bioactivity but amorphous and orthorhombic-Nb2O5 ones do not. Oxygen depletion and oxide vacancies may seriously affect the bioactivity of oxide-containing coatings. It is essential to perform in-depth studies to address their role in the bioactivity of NbxOy coatings.

The implant/cell interaction is greatly affected by the surface characteristics of the implant, such as chemical composition, surface roughness, surface energy, and surface wettability.179181 During the initial stage of implantation, when the surrounding cells are connected to the implant, the material adsorbs the existing proteins in the body fluid, such as vitronectin and fibronectin, to provide a desirable platform for cell adhesion and osseointegration. At higher values of hydrophilicity, i.e., at lower contact angles, the number of adsorbed proteins over the surface of the implant is increased, leading to superior cell adhesion and improved implant/cell interaction. The higher surface roughness results in a higher surface wettability.182184 The sol–gel Nb2O5-coated implants have more hydrophilic surface than bare ones irrespective of the chemical composition of the implant and the medium used for contact angle measurement.105,126

Depending on the operating factors involved in the sol–gel technique, which alter the surface roughness and phase structure of the Nb2O5 deposit, the deposit can variously affect the cell viability.119,121,155 For instance, while hexagonal-Nb2O5 coating showed a slight decrease in cell viability after 30 days of incubation, orthorhombic-Nb2O5 did not influence the number of living cells. The various unit cell geometry is obtained due to the difference in the calcination temperature. Although Nb2O5 coatings show no toxicity to the cells, in-depth studies are required to address how the crystalline structure changes the biological behavior of the coatings.119 The different surface roughness originating from various calcination times and temperatures determines the cell viability, where the highest cell number is attained for the Nb2O5 film with the smoothest surface. It is emphasized that the influence of surface roughness on the cell adherence rate can be neglected at longer incubation times. Whether the surface roughness alters the osteoblast adhesion greatly depends on the atomic arrangement of the coating. While the surface roughness has no influence on the cell spreading in the crystalline coating, the amorphous deposit with low surface roughness facilitates the spreading. Moreover, the increased surface roughness of the Nb2O5 film, e.g., at a surface roughness of 15 and 40 nm, promotes the collagen-I synthesis, where there is no dense collagen-I layer formed over the film with a roughness of 7 nm.121,155

Vascular stent grafts are largely employed in aortic aneurysms and coronary bypasses. Development of synthetic vascular grafts, which possess both high performance and durability are of prime significance as there is still a serious fail in stenting, about 10–25% out of total cases. To address this issue and develop a hybrid ceramic coated-metallic stent, Chai et al.185 have made a comparative insight into the performance of cp Ti that coated by various biocompatible ceramic coatings. They have reported both TiO2 and Nb2O5, coatings deposited by the sol–gel technique show favorable homocompatibility, while SiO2 is less homocompatibile. They have also inferred that the Nb2O5-coated Ti exhibits the highest endothelial cell proliferation after various days of incubation.

As mentioned earlier, the bioactivity of the implant can be examined via immersion in some standard media, containing various salts and proteins that simulate in vivo conditions, including SBF, simulated salvia, phosphate buffer saline (PBS), Dulbecco’s Modified Eagle’s Medium (DMEM), etc.118,186,187 The results of in vitro bioactivity assay carried out in human saliva medium, which is different with simulated saliva due to encompassing more than 43 proteins and enzymes, such as statherin and mucins, etc., indicated that the microcone shaped Nb2O5 formed during the anodization of Nb foil in the electrolyte composed of 2.5 wt % HF(aq) and 100 mg of NaF can support mineral formation. The phase structure of the minerals is consisted of Hap.118,188,189 Elemental analysis showed a higher phosphorus content than calcium in the structure of the film.118 This indicates the presence of an intermediate amorphous phase, covering the underlying HAp crystals.190

In general, the PEO fabricated Nb2O5 coatings are more hydrophilic than the underlying Nb substrate. However, the chemical composition of the electrolyte bath can slightly alter the wettability.120,147 For instance, the coatings produced from silicate-based solutions are more hydrophilic than those deposited from phosphate-based electrolytes. Evaluating the biocompatibility of the PEO produced Nb2O5 coatings with a variety of assays, such as 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl tetrazolium bromide (MTT), apoptosis test, direct contact assay, confirms that they are nontoxic and the cultured cells over their surface can stay alive and grow.147Table 7 outlines the biological properties of pure NbxOy coatings.

Table 7. Biological Properties of Pure NbxOy Coatingsa.

coating type deposition method type of cultured cell assay type % highest cell viability/% highest reported absorbance apatite generation ability refs
NbOxNy sputter deposition healthy human lung fibroblast cell line (MRC5) MTT 125 ± 5     (138)
Nb2O5 sputter deposition cells derived from a human cementoblastoma MTT   0.49   (149)
Nb2O5 sol–gel mouse fibroblast cell line L-929 MTT 105 ± 45     (119)
Nb2O5 sol–gel MC3T3-E1 osteoblast cell direct contact 102 ± 10     (121)
Nb2O5 sol–gel HPMEC* endothelialcells triplicate assay   1.2 ± 0.1   (185)
Nb2O5 PEO fibroblast cell line L-929 MTT 83     (147)
a

HPMEC: Primary human pulmonary microvascular endothelial cells.

4. NbxOy-Containing Coatings

Background

Irrespective of the applied surface finish technique, the deposited pure NbxOy coatings may face several drawbacks. One practical way to address these drawbacks and induce new beneficial properties to the pure NbxOy coatings is incorporating elements or compounds into the pure coating. The concept of “composite coating” is a potential approach that leads to fruitful results in this case. The included reinforcing agent can be either organic or inorganic chemicals. PEG,148 CuO,122 graphene,122 cerium (Ce),191 zinc (Zn),192 copper (Cu),140 and erythromycin43 are examples of reinforcing phases that have been successfully included to enhance the final properties of the pure NbxOy coatings. There are also some studies that have exploited Nb2O5 as a secondary phase to develop more efficient biocoatings, which are encompassed in this section.73,110 Apart from the well-established deposition techniques, e.g., sputter deposition and sol–gel, NbxOy-containing layers can also be obtained through annealing the CaP-coated Nb-containing substrates193 and/or air plasma spraying of Nb coatings.194 The following text presents the characteristics of the NbxOy-containing coatings based on the type of the added material(s).

Metallic Additives

Cu, Ce, and Zn are the metallic elements that have been incorporated into the Nb2O5 coatings to provide special properties and functions to them. The metallic element-reinforced Nb2O5 coatings can be deposited through sputter deposition and sol–gel routes.117,140,191,192 In the former case, an extra target composed of the desired element is coupled with the Nb target, leading to the metallic element-reinforced Nb2O5 coating. The amount of the reinforcing phase in the resultant film can be controlled either via changing the operating parameters or the amount of element in the target.117 To obtain sol–gel derived metallic element-reinforced Nb2O5 deposits, a sol corresponding to the additive composition is prepared, followed by adding the as-prepared sol to the Nb2O5 sol. The content of additives in the deposited film can be tailored by varying the concentration of the precursors in the sol.191,192 For instance, the concentration of zinc nitrate hexahydrate in the sol governs the content of Zn in the sol–gel Zn–Nb2O5 composite deposits.192

Cu is an essential trace element, known as the most active biological active metal, playing a central role in catalyzing metabolic processes. It also contributes to the angiogenesis and bone formation. Excellent activity of the Cu, in particular Cu2+ cation as the most active form of Cu, against microorganisms along with negligible cytotoxicity makes it a propitious candidate for antibacterial applications.195197 Cu–Nb2O5 composite coatings are obtained using the magnetron sputtering technique. The addition of Cu to the Nb2O5 deposit markedly affects the surface-related characteristics, where a more smooth and homogeneous surface is obtained. The rounded particles formed in microstructure of the pure coatings disappear with the inclusion of Cu.117,140 The surface roughness of the Nb2O5 decreases with incorporation of the Cu, followed by a slight increase in roughness with further increase in the content of Cu in the coating. However, the surface roughness of the composite layers is still much less than that of pure film.117 Tribomechanical properties, including nanohardness, elastic modulus, and scratch resistance of the Nb2O5 films, degrade when the Cu is included in the film. A 10% and 26% decrease in nanohardness and elastic modulus is reported for Nb2O5-25 at % Cu composite coatings compared to Nb2O5 coating, respectively. The introduction of a noble metal to the ceramic coating promotes its corrosion resistance.117 The more Cu that is included, the higher the corrosion protection obtained.117,140 The postcorrosion surface morphology reveals that, unlike the pure Nb2O5 layer, there is no change in morphology of the composite films due to the slow generation of the passive layer over the composite film.140 Unlike the Nb2O5 layer, the nanohardness of the composite coating did not change after corrosion assay. The diminished nanohardness of the pure film is attributed to the formation of the passive layer during the assay.140,198,199 The contact angle of composite coating is slightly higher than that of pure layer, irrespective of the type of employed droplet.117

Recently, tremendous attention has been paid to Ce in biomedical applications due to its exponential corrosion resistance. In its oxide form, Ce gives rise to the biological responses of the cultured cells, such as proliferation and differentiation.200,201 Katta et al.191 have merely studied the influence of Ce content in the structure of sol–gel derived Ce-included Nb2O5 coating on final characteristics of the produced layer without addressing how embedded metallic additives can affect the properties of pure Nb2O5 coating. They have reported that an increase in Ce content from 1 to 2 wt % has no noticeable effect on the morphology and phase structure of the films. However, the higher Ce amount in the film leads to the enhanced thickness and surface roughness. On the other hand, the hydrophilicity and corrosion protection efficiency of the sol–gel layers enhance with a higher included Ce phase. The Ce-reinforced Nb2O5 coatings exhibit self-healing behavior within the anodic region of the potentiodynamic polarization curve. The formed cerium oxides and/or hydroxides can prevent corrosion progress through healing the present defects over the surface of the sol–gel film. The following reactions may occur within the anodic region:202,203

graphic file with name ao2c00440_m002.jpg 2
graphic file with name ao2c00440_m003.jpg 3

Regardless of the content of included Ce, Ce-included Nb2O5 films illustrate appropriate biocompatibility where osteoblast cells can attach and spread over the surface of the films due to the presence of porous and rough surfaces. Moreover, an increase in Ce content has constructive influence on the biocompatibility.191

Similar to Cu, Zn is an essential element, playing a major role in mineralization, metabolic functions, cell signaling, and immune system performance. High antibacterial performance of the Zn offers the prospect of using Zn as a strong antibacterial agent.192,204 The present microcracks all over the microstructure of the sol–gel derived Nb2O5 films that are generated due to difference between the CTE of the top layer and biomaterial can be eliminated by addition of Zn to the pure coating. This can contribute to a higher hardness and corrosion resistance.67,192 Furthermore, the increased Zn content not only changes the surface morphology but also raises the surface roughness. Unlike morphological characteristics, the phase composition of the films do not change with Zn inclusion. Compared to pure Nb2O5, Zn-reinforced ones illustrate higher in vitro bioactivity through facilitating the precipitation of higher amounts of mineral stoichiometric HAp. There is an inverse relation between the content of Zn in the Zn-incorporated Nb2O5 layer and the measured contact angle value. It is reported that there is an optimum content for Zn at the highest corrosion performance. In such an optimum condition, the highest compactness, crystallinity, and sufficient thickness are obtained. Simply put, increasing the Zn amount for any percent may degrade this property. The included Zn considerably improves the antibacterial activity of the pure Nb2O5 layer owing to the formation of ZnO in the structure of the layer.192 The formed ZnO can inhibit bacteria adhesion and suppress their growth via enabling a desirable platform for generation of reactive oxygen species (ROS), e.g., O2–, OH, H2O2, that severely harms bacteria.205 The higher surface roughness and presence of functional active groups over the surface of Zn-reinforced Nb2O5 films facilitate the osteoblast cell adhesion and proliferation.192

Ceramic Additives

Recently, a rising interest has been seen in using CuO for biomedical applications owing to its excellent physicochemical properties and high antibacterial and antimicrobial performance against a broad spectrum of bacteria, such as S. aureus, E. coli, P. aeruginosa, S. typhimurium, K. pneumoniae, Enterococcus faecalis, etc. In general, it can be stated that the oxide form of the metals that possess antibacterial activity, e.g., Zn, Cu, and Zn, can exhibit similar performance.206208 However, a controlled dosage of CuO should be employed for biomedical applications since a higher concentration of this metal oxide can cause toxicity to mammalians and vertebrates. This is due to fabrication of ROS, which can highly damage mitochondria and deoxyribonucleic acid (DNA).209211 CuO-containing Nb2O5 films are deposited via magnetron sputtering method, where the CuO reinforcing phase can be included in the microstructure of growing Nb2O5 though a cosputtering route. The composite coating deposition process accomplished by sputtering Nb and Cu targets under a pure oxygen atmosphere. A smoother surface can be obtained when the CuO reinforcement is incorporated into the microstructure of pure Nb2O5 film. The included CuO phase can also improve the nanohardness of the pure layer from 4.78 to 7.19 GPa. Moreover, a slight increase in maximum scratch depth of coatings with CuO addition is observed. Interestingly, both Jcorr and Ecorr values are escalated with inclusion of CuO to a pure Nb2O5 layer. This means while the embedded reinforcing agent has a beneficial influence on the thermodynamic aspect of corrosion, it degrades the kinetic aspect. The mechanism(s) governing such a corrosion behavior remains to be elucidated in the published literature.122

Polymeric Additives

The polymeric additives, e.g., PEG and erythromycin, can be added to the pure Nb2O5 layers to tackle specific deficiencies or evolve new functions. The polymer-reinforced Nb2O5 composite layers can be deposited via magnetron sputtering and sol–gel techniques. While sputter deposition of composite coatings is carried out using composite targets,45 to obtain sol–gel derived polymer-added Nb2O5 composite films it is necessary to prepare a polymer-containing sol, followed by adding the sol to the as-prepared Nb2O5 sol with desired ratios.43

PEG is a class of antifouling polymers, adopting its high antifouling performance from surface hydration and steric hindrance effects. The hydrophilic nature, low toxicity, solubility in water, and bioactivity of PEG appreciably extend its clinical applications, such as drug delivery and wound healing.212 In drug delivery applications, PEG serves as a carrier for hydrophobic drug molecules. This can highly promote the dissolution properties and aqueous stability of the loaded drugs.45,148 Besides, PEG can serve as a linker, which covalently binds peptides to antibacterial agents, such as gold nanoparticles.213 While the included PEG has no noticeable effect on the phase structure of the pure Nb2O5 layers, an increase in surface roughness with the PEG embedment has been reported.45 The composite coating illustrates higher bonding strength to the underlying implant than that of the pure film. It is attributed to the higher content of formed apatite throughout the coated-implant. Moreover, the composite layer that implanted in vivo for 2 weeks has higher adhesion strength than that of the pure layer implanted for 6 weeks, indicating the faster healing period for dental implants coated by composite film. Such an achievement brings forth promising outcomes for both patients and dentists.148 For antibacterial applications, the Nb2O5-PEG composite deposit can be immersed in an antimicrobial peptide-containing medium, e.g. glycopeptide, so that the PEG acts as the linker for loading the antimicrobial peptide on the coating. The included PEG gives rise to drug delivery efficiency, thereby promoting the antibacterial activity of the coated implant against both Gram-negative and Gram-positive bacteria. However, the concentration of the incorporated antimicrobial peptide that induced acceptable antibacterial performance is different for various types of the bacteria.45

Erythromycin is usually employed to treat bacterial infection and can be synthesized using a strain of Streptomyces erythreus, which is present in soil. It shows high antibacterial efficiency against some of the Gram-positive bacteria, including listeria, streptococci, etc. Antibacterial activity of erythromycin is related to its ability to prevent protein synthesis of bacteria through binding to the 50S ribosomal element.214216 The antibacterial performance of sol–gel derived erythromycin-loaded Nb2O5 films directly depends on the dosage of the loaded antibiotic and immersion time. It may be correlated with the higher concentration of the released erythromycin at higher antibiotic dosage and/or prolonged immersion time.43

Summary of Properties

As it is perceived, section 4 of this review encompasses descriptions about general aspects of Nb2O5 coatings reinforced by various additives and attempts to provide a meaningful relationship between the processing variables, such as content of included reinforcements and parameters involved in the deposition method, and overall characteristics of the resultant composite layers. To give a better view of the influence of studied variables on the mechano-corrosion properties and wettability of the composite films, comparative tables are presented. This section not only contributes to scholars in selection of the most appropriate fabrication method, material type, and operating factors depending on necessities of the targeted application but also reveals the untreated points in this field that remain to be addressed. Table 8 and Table 9 summarize the mechanical and corrosion properties of the particle-strengthened Nb2O5 composite layers.

Table 8. Mechanical Properties of the Particle-Strengthened Nb2O5 Composite Layers.

type of reinforcing phase studied variable (amount) deposition technique characterization method highest reported hardness highest reported elastic modulus/GPa refs
Cu content of reinforcing phase; 17–25 at % magnetron sputtering nanoindentation 7.79 GPa 105 (117)
Cu composite film containing constant content of 25 at % magnetron sputtering nanoindentation 7.79 GPa   (140)
Zn molar ratios of Nb to Zn: ranging from 0.5:0 to 0.5:0.75 sol–gel Vickers 363 HV   (192)
CuO composite film containing constant content of CuO magnetron sputtering nanoindentation 7.19 GPa   (122)

Table 9. Mechanical Properties of the Particle-Strengthened Nb2O5 Composite Layers.

type of reinforcing phase studied variable (amount) deposition technique corrosive mediuma lowest reported jcorr/μA cm–2 highest reported Ecorr/mV highest reported Rp/kΩ cm2 refs
Cu content of reinforcing phase; 17–25 at % magnetron sputtering 0.5 M NaCl, 2 g/L KF; pH 2 at room temperature 0.07 9   (117)
Cu composite film containing constant content of 25 at % magnetron sputtering 0.5 M NaCl, 2 g/L KF; pH 2 at room temperature 0.07 10   (140)
Ce weight percent of Ce in the prepared sols; 1–2 wt % sol–gel SBF 0.012 –234 297930 (191)
Zn molar ratios of Nb to Zn: ranging from 0.5:0 to 0.5:0.75 sol–gel SBF 0.07 –172 548.30 (192)
CuO composite film containing constant content of CuO magnetron sputtering 0.5 M NaCl, 2 g/L KF; pH 2 at room temperature 9.93 262   (122)
a

The test temperature is a key factor affecting the results of corrosion assay. The majority of the published papers can be criticized for not reporting the accurate temperature recorded during the corrosion test since room temperature may be varied in a considerable range depending on the place and time at which the test was performed.

Reported results suggested that the highest value of nanohardness of the composite films is in the range of 7–8 GPa. To obtain different nanohardness and elastic modulus values, the content of included reinforcement and/or processing parameters can be changed.

The results reveal that the particle-reinforced Nb2O5 composite layers are generally corrosion resistant and show high corrosion performance in a variety of corrosive media. However, the importance of the test medium cannot be neglected. The contact angle values of the particle-reinforced Nb2O5 composite coatings are outlined in Table 10.

Table 10. Contact Angle Values of the Particle-Reinforced Nb2O5 Composite Coatingsa.

type of reinforcing phase studied variable (amount) deposition technique used droplet type in wettability assay the lowest reported contact angle/deg refs
Cu 17–25 at % magnetron sputtering water, ethanol, ethylene glycol 7.2 ± 2.3 (117)
Ce 1–2 wt % sol–gel water 26 (191)
Zn molar ratios of Nb:Zn from 0.5:0 to 0.5:0.75 sol–gel water 26 (192)
a

In accordance with reported data, the particle-reinforced Nb2O5 composite coatings are hydrophilic, irrespective of the fabrication method.

Nb2O5 as Reinforcing Phase

Nb2O5 carries the promise of improved morphological characteristics, hardness, corrosion resistance, bioactivity, and biocompatibility if it is added as a reinforcing phase to a matrix at the proper concentration. Usually, Nb2O5 is employed to strengthen the polymer layers, such as polypyrrole (PPY) and polydimethylsiloxane (PDMS).73,110

The surface uniformity of the electrochemically deposited PPY coatings is enhanced with the introduction of Nb2O5 particles. Furthermore, the size and number of present microcracks, surface roughness, and grain size decrease when Nb2O5 is codeposited.110 The microstructure of sol–gel derived PDMS-Nb2O5 film becomes more polymorphic with an increase in the concentration of Nb2O5 dopant. The content of the included dopant has no positive contribution to elimination of the present microcracks.73 The composite PPY-Nb2O5 layer shows higher hardness than that of pure PPY film due to the densification of the polymer matrix. Moreover, the corrosion protection efficiency of the PPY films rises with codeposition of the Nb2O5 phase due to increasing the localization of charge in the PPY chain and improving the surface homogeneity. The higher the included reinforcing phase, the superior the corrosion resistance. The hydrophilicity of PPY coating enhances when Nb2O5 is included in the deposit. This is ascribed to the hydrophilic nature of Nb2O5 in the polymer matrix.110 On the other hand, there is an interesting relationship between the wettability of the PDMS-Nb2O5 composite layer and Nb2O5 concentration. The hydrophobicity of the PDMS coating is enhanced with the codeposition of Nb2O5 up to a certain concentration, i.e., 80%, followed by a slight decrease in contact angle value with a further rise in the concentration of Nb2O5 in the range of 90–100%. The enhanced hydrophobicity in the mentioned range is attributed to the formation of hybrid spheres in the structure of the deposit that enhances the surface roughness. The interesting point is that the pure Nb2O5 coating (100% Nb2O5) exhibits a hydrophilic nature. The contact angle value of pure Nb2O5 is less than a half and a third of pure PDMS and PDMS-80% Nb2O5, respectively.73 The reported results on the effect of Nb2O5 concentration on the biocompatibility of polymer matrix composite deposits are different. While Kumar et al.110 have reported a direct correlation between the amounts of loaded Nb2O5 and biocompatibility of the electrochemically deposited PPY-Nb2O5 coatings, Young et al.73 have illustrated that the biocompatibility of PDMS-Nb2O5 increases with a certain amount of codeposited Nb2O5, i.e., 40%, followed by a decrease in the property by a further rise in Nb2O5 concentration up to 70%. Again, an increase in biocompatibility has been observed when Nb2O5 concentration is increased in the range of 70–100%. The dependence of final characteristics of a ceramic-reinforced polymer matrix composite coating to the content of the included ceramic reinforcing phase, herein Nb2O5, hold the promise to meet the various expectations in fabrication and applications of implants. The composite films containing low ceramic phase are suitable for soft tissue applications, while for hard tissue applications it is recommended to increase the content of Nb2O5 in the coating. Table 11 outlines the reported results for contact angle and cell proliferation of Nb2O5-reinforced polymer matrix composite layers.

Table 11. Reported Results for Contact Angle and Cell Proliferation of Nb2O5-Reinforced Polymer Matrix Composite Layers.

coating type deposition method concentration of Nb2O5 droplet liquid used in wettability assay lowest reported contact angle/deg type of cultured cells highest reported absorbance refs
PPY-Nb2O5 electrochemical deposition 1–20 mg L–1 in the electrolyte SBF 12.9 MG-63 osteoblast cell ≈1.1 (110)
PDMS-Nb2O5 sol–gel 20–100% in the coating distilled water 43.4 ± 6.7 fibroblast 2.5 (73)

5. Conclusions and Future Horizons

Surface finishing techniques and the type of materials used to modify the surface of the biomaterial have undergone unprecedented development over the past 3 decades. Niobium oxide-containing coatings are promising candidates to tackle the challenges that threaten the successful long-term use of the metallic implants. The present review paper attempts to discuss the properties, application, and market of niobium and its oxides, putting emphasis on their biomedical properties and applications. The overall characteristics of the pure NbxOy and NbxOy-containing coatings deposited by various surface finishing techniques will also be treated in detail, putting stress on the importance of surface finishing with pure NbxOy and NbxOy-containing layers along with drawing the relationship between the operating parameters and final properties of the resultant films.

In summary, pure NbxOy and NbxOy-containing coatings can be deposited using a variety of dry and wet techniques. The coatings greatly affect the overall characteristics of the underlying biomaterials, where the deposited layers not only address the deficiencies of the biomaterials, such as corrosion resistance but also induce some new properties, e.g., antibacterial activity. The NbxOy-containing composite coatings in which the NbxOy matrix is reinforced by metallic, ceramic, and polymeric additives, are developed to provide much more beneficial properties. The selection of the included reinforcing phase is highly dependent on the targeted application. However, the potential of NbxOy-containing coatings is still not really being put to work for developing high-quality biomaterials used in a variety of clinical applications. Therefore, there is still much to be investigated and understood in this field. In the following text, a list of challenges and upcoming development prospects is presented in light of the status quo of NbxOy-coated biomaterials and the roles that NbxOy-containing coatings can play in biomedical implants.

Having reviewed the benefits and limitations of surface finishing methods, it can be readily appreciated that several approaches that can be employed to fabricate efficient NbxOy-containing coatings for clinical applications. Electrodeposition, EPD, and spraying are potential techniques that may lead to uniform layers with fruitful properties. Regarding pure NbxOy films, in-depth studies are needed to fully understand the effects of operating parameters for each coating technique on the final properties of the resultant layer to explicitly state how a given parameter can change the final properties. The published literature fails to address this important challenge. Moreover, studying the role of pre/post-treatments may lead to some profitable outcomes. Detailed assessments are needed to correlate the changes in mechano-corrosion and biological performance of the NbxOy coated-implants with the physicochemical characteristics of the coatings caused by alteration of the reaction environment during coating deposition.

With respect to a variety of reinforcing phases, biocompatible and bioactive additives, e.g., graphene oxide, carbon nanotubes, and metal oxides, come to mind, which can potentially promote the final performance of the pure NbxOy layers and/or induce new functions to them. The concept of duplex particles-reinforced NbxOy coatings should attract considerable future R&D attention. Having a glance at the number of the published papers in the field of NbxOy-containing composite coatings, it can be inferred that much work remains to establish reliable relationships between the type of reinforcing phases and functions of the resultant composite films. The authors believe that a more concerted focus on NbxOy-containing composite coatings produced by different wet and dry techniques, in which NbxOy acts as the matrix or reinforcing phase, can offer the prospect of using these coatings as the next generation of biocompatible layers in biomedical applications. The role of processing parameters in obtaining efficient NbxOy-containing composite coatings should not be overlooked. The deposition of duplex particles-reinforced NbxOy-containing composite coatings in which the NbxOy matrix is strengthened by two reinforcing agents, and the application of interlayer(s) are two other feasible suggestions which are believed to open numerous possibilities to realize high quality NbxOy-containing composite coatings.

Biographies

Mir Saman Safavi received a B.Sc. (2015) and M.Sc. (2018) from the Department of Materials Science and Engineering at University of Tabriz. During his B.Sc. and M.Sc. studies, he worked on “studying on mechanical properties of Cu-based composites” and “electrodeposition of Ni-based nanocomposite coatings”, respectively. He is a Ph.D. candidate at Sahand University of Technology, working on the surface modification of biomaterials using the electrochemical deposition technique. Mir Saman Safavi has been the author and coauthor of over 40 ISI and conference papers, which have been cited more than 440 times, giving him an H-index of 14.

F. C. Walsh is an Emeritus Professor in Electrochemical Engineering in the Department of Mechanical Engineering at the University of Southampton, U.K., where he founded the Electrochemical Engineering Laboratory He is a member of the National Centre for Advanced Tribology at Southampton and the Materials Engineering Research Group. Previous positions have included Director of the Research Institute for Industry at Southampton, Head of Chemical Engineering at the University of Bath, Head of Chemistry, Physics & Radiography then Pharmacy & Biomedical Sciences at the University of Portsmouth. He has been a Visiting Professor at the University of Wollongong and the University of Strathclyde. Frank has spent his life crossing the boundaries of chemistry, biology, materials science, and chemical engineering. He holds the qualifications of B.Sc. (Applied Chemistry), M.Sc. (Materials Protection), and Ph.D. (Electrodeposition and Electrochemical Engineering) following studies at Portsmouth, Manchester, and Loughborough Universities. He is a Chartered-Chemist, Environmentalist, Scientist, and Engineer and a Fellow of the Royal Society of Chemistry, the Institute of Mining, Materials and Metallurgy, the Institute of Materials Finishing, and the Institute of Corrosion. His R&D output, which spans the areas of energy conversion, electroactive nanomaterials, coating technology, electrochemical monitoring and sensors, corrosion, surface finishing, and electrochemical process engineering, includes 4 text books, 200 conference presentations, and over 500 research papers. Frank was awarded the Westinghouse Prize in 1999 and 2009 (best paper on metallic coatings) and the Johnson Matthey Silver Medal of the Institute of Metal Finishing in 2007 (precious metal nanoparticles inside nanotube arrays), the Breyer Medal of the Royal Australian Chemical Institute in 2000 (international contributions to electrochemical engineering and energy), the NACE Fellow honour in 2011, the Inman Medal of the UK Centre for CO2 reduction, London (international contributions to environment and energy) in 2012, the 2017 Castner Medal of the (achievements in industrial electrochemistry), and the Gold Medal, the highest award of the Institute of Materials Finishing (lifelong contributions to electrochemical surface finishing) in 2018. In collaboration with industry, Frank has supervised over 50 Ph.D. degrees and developed/improved over 60 industrial processes.

Livia Visai is an Associate Professor in Biochemistry at the Department of Molecular Medicine (DMM), Faculty of Medicine, University of Pavia, Pavia, Italy. She is Referent of the Academic Strategic project of University of Pavia titled “Center for Health Technologies” (CHT) and Coordinator of the Nanomedicine Pillar, Pavia, Italy. She is also operating as an Associate Professor within the National Health Service of the Scientific Clinical Institutes (ICS) Maugeri, Società Benefit Spa, IRCCS, Pavia, Italy. She is coordinating a laboratory interested in studying the effect exerted by the application of nanotechnology in tissue regeneration and in bone remodeling (in particular in microgravity), in the reduction of bacterial infection, and in the diagnosis and therapy of cancer. She has more than 184 papers published in PubMed with a WoS H-index of 46.

Jafar Khalil-Allafi received a B.Sc. (1988; Isfahan University of Technology-Iran), M.Sc. (1992; Iran University of Science and Technology-Iran), and Ph.D. (2002; Ruhr University Bochum-Germany) with the highest rank. Prof. Jafar Khalil-Allafi has received international and national awards and fellowships such as the Thyssen-Krupp innovation price 2002 in Germany. He has been also the supervisor and advisor of more than 100 postgraduate theses. Prof. Jafar Khalil-Allafi has been the author and coauthor of 3 patents, 3 books, and over 100 papers which have been cited more than 3500 times, giving him an H-index of 29.

The authors declare no competing financial interest.

References

  1. Zimmermann E. A.; Busse B.; Ritchie R. O. The fracture mechanics of human bone: influence of disease and treatment. Bonekey Rep. 2015, 4, 743. 10.1038/bonekey.2015.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Tabatabaei F. S.; Torres R.; Tayebi L. Biomedical materials in dentistry. Applications of Biomedical Engineering in Dentistry 2020, 3–20. 10.1007/978-3-030-21583-5_2. [DOI] [Google Scholar]
  3. National Institutes of Health Clinical Applications of Biomaterials. NIH Consens. Statement 1982, 4, 1–19. [Google Scholar]
  4. Hench L. L.An Introduction to Bioceramics, 2nd ed.; World Scientific, 2013. [Google Scholar]
  5. Safavi M. S.; Surmeneva M. A.; Surmenev R. A.; Khalil-Allafi J. RF-magnetron sputter deposited hydroxyapatite-based composite & multilayer coatings: A systematic review from mechanical, corrosion, and biological points of view. Ceram. Int. 2021, 47, 3031–3053. 10.1016/j.ceramint.2020.09.274. [DOI] [Google Scholar]
  6. Safavi M. S.; Walsh F. C.; Surmeneva M. A.; Surmenev R. A.; Khalil-Allafi J. Electrodeposited Hydroxyapatite-Based Biocoatings: Recent Progress and Future Challenges. Coatings 2021, 11, 110. 10.3390/coatings11010110. [DOI] [Google Scholar]
  7. Shokri N.; Safavi M. S.; Etminanfar M.; Walsh F. C.; Khalil-Allafi J. Enhanced corrosion protection of NiTi orthopedic implants by highly crystalline hydroxyapatite deposited by spin coating: The importance of pre-treatment. Mater. Chem. Phys. 2021, 259, 124041. 10.1016/j.matchemphys.2020.124041. [DOI] [Google Scholar]
  8. Mohammad Salahi Tohidi P.; Safavi M. S.; Etminanfar M.; Khalil-Allafi J. Pulsed electrodeposition of compact, corrosion resistant, and bioactive HAp coatings by application of optimized magnetic field. Mater. Chem. Phys. 2020, 254, 123511. 10.1016/j.matchemphys.2020.123511. [DOI] [Google Scholar]
  9. Sheykholeslami S. O.; Fathyunes L.; Etminanfar M.; Khalil-Allafi J. In-vitro biological behavior of calcium phosphate coating applied on nanostructure surface of anodized Nitinol alloy. Mater. Res. Express. 2019, 6, 095407. 10.1088/2053-1591/ab2f39. [DOI] [Google Scholar]
  10. Fathyunes L.; Khalil-Allafi J.; Sheykholeslami S. O.; Moosavifar M. Biocompatibility assessment of graphene oxide-hydroxyapatite coating applied on TiO2 nanotubes by ultrasound-assisted pulse electrodeposition. Mater. Sci. Eng., C 2018, 87, 10–21. 10.1016/j.msec.2018.02.012. [DOI] [PubMed] [Google Scholar]
  11. Fathyunes L.; Khalil-Allafi J. Effect of employing ultrasonic waves during pulse electrochemical deposition on the characteristics and biocompatibility of calcium phosphate coatings. Ultrason. Sonochem. 2018, 42, 293–302. 10.1016/j.ultsonch.2017.11.041. [DOI] [PubMed] [Google Scholar]
  12. Etminanfar M. R.; Khalil-Allafi J.; Sheykholeslami S. O. The effect of hydroxyapatite coatings on the passivation behavior of oxidized and unoxidized superelastic nitinol alloys. J. Mater. Eng. Perform. 2018, 27, 501–509. 10.1007/s11665-018-3200-6. [DOI] [Google Scholar]
  13. Mohandesnezhad S.; Etminanfar M.; Mahdavi S.; Safavi M. S. Enhanced bioactivity of 316L stainless steel with deposition of polypyrrole/hydroxyapatite layered hybrid coating: Orthopedic applications. Surf. Interfaces 2022, 28, 101604. 10.1016/j.surfin.2021.101604. [DOI] [Google Scholar]
  14. Etminanfar M.; Khalil-Allafi J. On the electrodeposition of Ca-P coatings on nitinol alloy: a comparison between different surface modification methods. J. Mater. Eng. Perform. 2016, 25, 466–473. 10.1007/s11665-015-1876-4. [DOI] [Google Scholar]
  15. Hu C.; Ashok D.; Nisbet D. R.; Gautam V. Bioinspired surface modification of orthopedic implants for bone tissue engineering. Biomaterials 2019, 219, 119366. 10.1016/j.biomaterials.2019.119366. [DOI] [PubMed] [Google Scholar]
  16. Holman H.; Kavarana M. N.; Rajab T. K. Smart materials in cardiovascular implants: Shape memory alloys and shape memory polymers. Artif. Organs 2021, 45, 454–463. 10.1111/aor.13851. [DOI] [PubMed] [Google Scholar]
  17. Anabtawi M.; Thomas M.; Lee N. J. The use of interlocking polyetheretherketone (PEEK) patient-specific facial implants in the treatment of facial deformities. A retrospective review of ten patients. J. Oral Maxillofac. Surg. 2021, 79, 1145.e1–1145.e9. 10.1016/j.joms.2020.12.009. [DOI] [PubMed] [Google Scholar]
  18. Hämmerle C. H.; Tarnow D. The etiology of hard-and soft-tissue deficiencies at dental implants: A narrative review. J. Clin. Periodontol. 2018, 45, S267–S277. 10.1111/jcpe.12955. [DOI] [PubMed] [Google Scholar]
  19. Bundy K. Biomaterials and the Chemical Environment of the Body. In Joint Replacement Technology; Woodhead Publishing Limited, 2008; pp 56–80. [Google Scholar]
  20. Pandey A.; Awasthi A.; Saxena K. K. Metallic implants with properties and latest production techniques: a review. Adv. Mater. Process. Technol. 2020, 6, 405–440. 10.1080/2374068X.2020.1731236. [DOI] [Google Scholar]
  21. Wong K. C.; Scheinemann P. Additive manufactured metallic implants for orthopaedic applications. Sci. China Mater. 2018, 61, 440–454. 10.1007/s40843-017-9243-9. [DOI] [Google Scholar]
  22. Silva-Bermudez P.; Rodil S. An overview of protein adsorption on metal oxide coatings for biomedical implants. Surf. Coat. Technol. 2013, 233, 147–158. 10.1016/j.surfcoat.2013.04.028. [DOI] [Google Scholar]
  23. Etminanfar M. R.; Sheykholeslami S. O.; Khalili V.; Mahdavi S. Biocompatibility and drug delivery efficiency of PEG-b-PCL/hydroxyapatite bilayer coatings on Nitinol superelastic alloy. Ceram. Int. 2020, 46, 12711–12717. 10.1016/j.ceramint.2020.02.038. [DOI] [Google Scholar]
  24. Khalili V.; Khalil-Allafi J.; Xia W.; Parsa A. B.; Frenzel J.; Somsen C.; Eggeler G. Preparing hydroxyapatite-silicon composite suspensions with homogeneous distribution of multi-walled carbon nano-tubes for electrophoretic coating of NiTi bone implant and their effect on the surface morphology. Appl. Surf. Sci. 2016, 366, 158–165. 10.1016/j.apsusc.2016.01.053. [DOI] [Google Scholar]
  25. Maleki-Ghaleh H.; Khalil-Allafi J. Characterization, mechanical and in vitro biological behavior of hydroxyapatite-titanium-carbon nanotube composite coatings deposited on NiTi alloy by electrophoretic deposition. Surf. Coat. Technol. 2019, 363, 179–190. 10.1016/j.surfcoat.2019.02.029. [DOI] [Google Scholar]
  26. Bazaka O.; Bazaka K.; Kingshott P.; Crawford R. J.; Ivanova E. P.. Metallic Implants for Biomedical Applications. Royal Society of Chemistry, 2021; pp 1–98. [Google Scholar]
  27. Kinani L.; Najih R.; Chtaini A. Corrosion inhibition of titanium in artificial saliva containing fluoride. Leonardo J. Sci. 2008, 12, 243–250. [Google Scholar]
  28. Choudhury N.; Kannan A.; Dutta N. Novel Nanocomposites and Hybrids for High-Temperature Lubricating Coating Applications. In Tribology of Polymeric Nanocomposites: Friction and Wear of Bulk Materials and Coatings, 2nd ed.; Elsevier B.V., 2013; pp 717–778. [Google Scholar]
  29. Lison D. Cobalt. In Handbook on the Toxicology of Metals; Elsevier, 2015; pp 743–763. [Google Scholar]
  30. Asghari R.; Safavi M. S.; Khalil-Allafi J. A facile and cost-effective practical approach to develop clinical applications of NiTi: Fenton oxidation process. Trans. IMF 2020, 98, 250–257. 10.1080/00202967.2020.1802104. [DOI] [Google Scholar]
  31. Nagels J.; Stokdijk M.; Rozing P. M. Stress shielding and bone resorption in shoulder arthroplasty. J. Shoulder Elb. Surg. 2003, 12, 35–39. 10.1067/mse.2003.22. [DOI] [PubMed] [Google Scholar]
  32. Al-Tamimi A. A.; Fernandes P. R.; Peach C.; Cooper G.; Diver C.; Bartolo P. J. Metallic bone fixation implants: A novel design approach for reducing the stress shielding phenomenon. Virtual Phys. Prototyp. 2017, 12, 141–151. 10.1080/17452759.2017.1307769. [DOI] [Google Scholar]
  33. Zhang E.; Xu L.; Yu G.; Pan F.; Yang K. In vivo evaluation of biodegradable magnesium alloy bone implant in the first 6 months implantation. J. Biomed. Mater. Res. 2009, 90A, 882–893. 10.1002/jbm.a.32132. [DOI] [PubMed] [Google Scholar]
  34. Hayashi K.; Matsuguchi N.; Uenoyama K.; Sugioka Y. Re-evaluation of the biocompatibility of bioinert ceramics in vivo. Biomaterials 1992, 13, 195–200. 10.1016/0142-9612(92)90184-P. [DOI] [PubMed] [Google Scholar]
  35. Kumar S.; Simpson D.; Smart R. S. C. Plasma processing for inducing bioactivity in stainless steel orthopaedic screws. Surf. Coat. Technol. 2007, 202, 1242–1246. 10.1016/j.surfcoat.2007.07.075. [DOI] [Google Scholar]
  36. MacDonald D. E.; Deo N.; Markovic B.; Stranick M.; Somasundaran P. Adsorption and dissolution behavior of human plasma fibronectin on thermally and chemically modified titanium dioxide particles. Biomaterials 2002, 23, 1269–1279. 10.1016/S0142-9612(01)00317-9. [DOI] [PubMed] [Google Scholar]
  37. Tambasco de Oliveira P.; Nanci A. Nanotexturing of titanium-based surfaces upregulates expression of bone sialoprotein and osteopontin by cultured osteogenic cells. Biomaterials 2004, 25, 403–413. 10.1016/S0142-9612(03)00539-8. [DOI] [PubMed] [Google Scholar]
  38. Lossdörfer S.; Schwartz Z.; Wang L.; Lohmann C. H.; Turner J. D.; Wieland M.; Cochran D. L.; Boyan B. D. Microrough implant surface topographies increase osteogenesis by reducing osteoclast formation and activity. J. Biomed. Mater. Res. 2004, 70A, 361–369. 10.1002/jbm.a.30025. [DOI] [PubMed] [Google Scholar]
  39. Shahryari A.; Azari F.; Vali H.; Omanovic S. The response of fibrinogen, platelets, endothelial and smooth muscle cells to an electrochemically modified SS316LS surface: Towards the enhanced biocompatibility of coronary stents. Acta Biomater. 2010, 6, 695–701. 10.1016/j.actbio.2009.07.007. [DOI] [PubMed] [Google Scholar]
  40. Horandghadim N.; Khalil-Allafi J.; Urgen M. Effect of Ta2O5 content on the osseointegration and cytotoxicity behaviors in hydroxyapatite-Ta2O5 coatings applied by EPD on superelastic NiTi alloys. Mater. Sci. Eng., C 2019, 102, 683–695. 10.1016/j.msec.2019.05.005. [DOI] [PubMed] [Google Scholar]
  41. Pham V. T. H.; Truong V. K.; Orlowska A.; Ghanaati S.; Barbeck M.; Booms P.; Fulcher A. J.; Bhadra C. M.; Buividas R.; Baulin V.; Kirkpatrick C. J.; Doran P.; Mainwaring D. E.; Juodkazis S.; Crawford R. J.; Ivanova E. P. Race for the surface”: eukaryotic cells can win. ACS Appl. Mater. Interfaces 2016, 8, 22025–22031. 10.1021/acsami.6b06415. [DOI] [PubMed] [Google Scholar]
  42. Subbiahdoss G.; Kuijer R.; Grijpma D. W.; van der Mei H. C.; Busscher H. J. Microbial biofilm growth vs. tissue integration:“the race for the surface” experimentally studied. Acta Biomater. 2009, 5, 1399–1404. 10.1016/j.actbio.2008.12.011. [DOI] [PubMed] [Google Scholar]
  43. Pradhan D.; Wren A.; Mellott N. A preliminary study into the efficacy of antibiotic doped niobium oxide coatings on 316L prepared by incipient wetness impregnation and sol-gel synthesis. Mater. Lett. 2017, 190, 150–153. 10.1016/j.matlet.2017.01.001. [DOI] [Google Scholar]
  44. Antoci V. Jr; Adams C. S.; Parvizi J.; Davidson H. M.; Composto R. J.; Freeman T. A.; Wickstrom E.; Ducheyne P.; Jungkind D.; Shapiro I. M.; Hickok N. J. The inhibition of Staphylococcus epidermidis biofilm formation by vancomycin-modified titanium alloy and implications for the treatment of periprosthetic infection. Biomaterials 2008, 29, 4684–4690. 10.1016/j.biomaterials.2008.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Khalaf H.; Hussien B.; Alzubaidi T. Anti-bacterial evaluation of biocomposite coated dental implant after immersion in glycopeptide. Indian J. Nat. Sci. 2018, 8, 49. [Google Scholar]
  46. Ahmadiyan S.; Khalil-Allafi J.; Etminanfar M. R.; Safavi M. S.; Hosseini M. Antibacterial activity and biocompatibility of Ag-coated Ti implants: importance of surface modification parameters. Trans. IMF 2022, 1–10. 10.1080/00202967.2021.2022842. [DOI] [Google Scholar]
  47. Safavi M. S.; Etminanfar M. A review on the prevalent fabrication methods, microstructural, mechanical properties, and corrosion resistance of nanostructured hydroxyapatite containing bilayer and multilayer coatings used in biomedical applications. J. Ultrafine Grained Nanostruct. Mater. 2019, 52, 1–17. 10.22059/JUFGNSM.2019.01.01. [DOI] [Google Scholar]
  48. Abdal-hay A.; Barakat N. A.; Lim J. K. Hydroxyapatite-doped poly (lactic acid) porous film coating for enhanced bioactivity and corrosion behavior of AZ31 Mg alloy for orthopedic applications. Ceram. Int. 2013, 39, 183–195. 10.1016/j.ceramint.2012.06.008. [DOI] [Google Scholar]
  49. Moskalewicz T.; Seuss S.; Boccaccini A. R. Microstructure and properties of composite polyetheretherketone/Bioglass coatings deposited on Ti-6Al-7Nb alloy for medical applications. Appl. Surf. Sci. 2013, 273, 62–67. 10.1016/j.apsusc.2013.01.174. [DOI] [Google Scholar]
  50. Surmenev R. A.; Surmeneva M. A. A critical review of decades of research on calcium phosphate-based coatings: How far are we from their widespread clinical application?. Curr. Opin. Biomed. Eng. 2019, 10, 35–44. 10.1016/j.cobme.2019.02.003. [DOI] [Google Scholar]
  51. Surmenev R. A.; Surmeneva M. A.; Ivanova A. A. Significance of calcium phosphate coatings for the enhancement of new bone osteogenesis-a review. Acta Biomater. 2014, 10, 557–579. 10.1016/j.actbio.2013.10.036. [DOI] [PubMed] [Google Scholar]
  52. Surmenev R. A.; Ivanova A. A.; Epple M.; Pichugin V. F.; Surmeneva M. A. Physical principles of radio-frequency magnetron sputter deposition of calcium-phosphate-based coating with tailored properties. Surf. Coat. Technol. 2021, 413, 127098. 10.1016/j.surfcoat.2021.127098. [DOI] [Google Scholar]
  53. Surmenev R. A.; Grubova I. Y.; Neyts E.; Teresov A. D.; Koval N. N.; Epple M.; Tyurin A. I.; Pichugin V. F.; Chaikina M. V.; Surmeneva M. A. Ab initio calculations and a scratch test study of RF-magnetron sputter deposited hydroxyapatite and silicon-containing hydroxyapatite coatings. Surf. Interfaces 2020, 21, 100727. 10.1016/j.surfin.2020.100727. [DOI] [Google Scholar]
  54. Hu R.; Lin C.; Shi H.; Wang H. Electrochemical deposition mechanism of calcium phosphate coating in dilute Ca-P electrolyte system. Mater. Chem. Phys. 2009, 115, 718–723. 10.1016/j.matchemphys.2009.02.022. [DOI] [Google Scholar]
  55. Robertson S. F.; Bandyopadhyay A.; Bose S. Titania nanotube interface to increase adhesion strength of hydroxyapatite sol-gel coatings on Ti-6Al-4V for orthopedic applications. Surf. Coat. Technol. 2019, 372, 140–147. 10.1016/j.surfcoat.2019.04.071. [DOI] [Google Scholar]
  56. Qi J.; Yang Y.; Zhou M.; Chen Z.; Chen K. Effect of transition layer on the performance of hydroxyapatite/titanium nitride coating developed on Ti-6Al-4V alloy by magnetron sputtering. Ceram. Int. 2019, 45, 4863–4869. 10.1016/j.ceramint.2018.11.183. [DOI] [Google Scholar]
  57. Ke D.; Vu A. A.; Bandyopadhyay A.; Bose S. Compositionally graded doped hydroxyapatite coating on titanium using laser and plasma spray deposition for bone implants. Acta Biomater. 2019, 84, 414–423. 10.1016/j.actbio.2018.11.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Zhang Q.; Leng Y.; Xin R. A comparative study of electrochemical deposition and biomimetic deposition of calcium phosphate on porous titanium. Biomaterials 2005, 26, 2857–2865. 10.1016/j.biomaterials.2004.08.016. [DOI] [PubMed] [Google Scholar]
  59. Drevet R.; Velard F.; Potiron S.; Laurent-Maquin D.; Benhayoune H. In vitro dissolution and corrosion study of calcium phosphate coatings elaborated by pulsed electrodeposition current on Ti6Al4V substrate. J. Mater. Sci.: Mater. Med. 2011, 22, 753–761. 10.1007/s10856-011-4251-5. [DOI] [PubMed] [Google Scholar]
  60. Jiang J. Y.; Xu J. L.; Liu Z. H.; Deng L.; Sun B.; Liu S. D.; Wang L.; Liu H. Y. Preparation, corrosion resistance and hemocompatibility of the superhydrophobic TiO2 coatings on biomedical Ti-6Al-4V alloys. Appl. Surf. Sci. 2015, 347, 591–595. 10.1016/j.apsusc.2015.04.075. [DOI] [Google Scholar]
  61. Kaliaraj G. S.; Vishwakarma V.; Alagarsamy K.; Kamalan Kirubaharan A.M. Biological and corrosion behavior of m-ZrO2 and t-ZrO2 coated 316L SS for potential biomedical applications. Ceram. Int. 2018, 44, 14940–14946. 10.1016/j.ceramint.2018.05.103. [DOI] [Google Scholar]
  62. Singh A.; Singh S. ZnO nanowire-coated hydrophobic surfaces for various biomedical applications. Bull. Mater. Sci. 2018, 41, 94. 10.1007/s12034-018-1611-5. [DOI] [Google Scholar]
  63. Dinu M.; Braic L.; Padmanabhan S. C.; Morris M. A.; Titorencu I.; Pruna V.; Parau A.; Romanchikova N.; Petrik L. F.; Vladescu A. Characterization of electron beam deposited Nb2O5 coatings for biomedical applications. J. Mech. Behav. Biomed. Mater. 2020, 103, 103582. 10.1016/j.jmbbm.2019.103582. [DOI] [PubMed] [Google Scholar]
  64. Milošev I.; Levašič V.; Kovač S.; Sillat T.; Virtanen S.; Tiainen V. M.; Trebše R. Metals for Joint Replacement. In Joint Replacement Technology; Elsevier, 2021; pp 65–122. [Google Scholar]
  65. Catledge S. A.; Vohra Y. K.; Bellis S. L.; Sawyer A. A. Mesenchymal stem cell adhesion and spreading on nanostructured biomaterials. J. Nanosci. Nanotechnol. 2004, 4, 986–989. 10.1166/jnn.2004.137. [DOI] [PubMed] [Google Scholar]
  66. Ratner B. D.; Zhang G. A History of Biomaterials. In Biomaterials Science; Elsevier, 2020; pp 21–34. [Google Scholar]
  67. Allo B. A.; Costa D. O.; Dixon S. J.; Mequanint K.; Rizkalla A. S. Bioactive and biodegradable nanocomposites and hybrid biomaterials for bone regeneration. J. Funct. Biomater. 2012, 3, 432–463. 10.3390/jfb3020432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Tennese W.; Cahoon J. Sensitization” still a problem in the intergranular corrosion of stainless steel surgical implants. Trends Biomater. Artif. Organs 1973, 1, 635–645. 10.3109/10731197309118566. [DOI] [PubMed] [Google Scholar]
  69. Sittig C.; Hahner G.; Marti A.; Textor M.; Spencer N. D.; Hauert R. The implant material, Ti6Al7Nb: surface microstructure, composition and properties. J. Mater. Sci. Mater. Med. 1999, 10, 191–198. 10.1023/A:1008997726370. [DOI] [PubMed] [Google Scholar]
  70. Shilder S. Tantalum and niobium as potential prosthetic materials. Biomaterials 1982, 13-10, 1982. [Google Scholar]
  71. Ward L.; Datta P. K. Physical Properties of Nb Coatings for Biomedical Applications. Key Eng. Mater. 1992, 72, 505–522. 10.4028/www.scientific.net/KEM.72-74.505. [DOI] [Google Scholar]
  72. Leitune V. C.; Collares F. M.; Takimi A.; de Lima G. B.; Petzhold C. L.; Bergmann C. P.; Samuel S. M. Niobium pentoxide as a novel filler for dental adhesive resin. J. Dent. 2013, 41, 106–113. 10.1016/j.jdent.2012.04.022. [DOI] [PubMed] [Google Scholar]
  73. Young M. D.; Tran N.; Tran P. A.; Jarrell J. D.; Hayda R. A.; Born C. T. Niobium oxide-polydimethylsiloxane hybrid composite coatings for tuning primary fibroblast functions. J. Biomed. Mater. Res. 2014, 102, 1478–1485. 10.1002/jbm.a.34832. [DOI] [PubMed] [Google Scholar]
  74. Velten D.; Eisenbarth E.; Schanne N.; Breme J. Biocompatible Nb2O5 thin films prepared by means of the sol-gel process. J. Mater. Sci. Mater. Med. 2004, 15, 457–461. 10.1023/B:JMSM.0000021120.86985.f7. [DOI] [PubMed] [Google Scholar]
  75. Wirth J.; Tayebi L. Engineering of Dental Titanium Implants and Their Coating Techniques. In Applications of Biomedical Engineering in Dentistry; Springer, 2020; pp 149–160. [Google Scholar]
  76. Pauline S. A.; Rajendran N. Corrosion behaviour and biocompatibility of nanoporous niobium incorporated titanium oxide coating for orthopaedic applications. Ceram. Int. 2017, 43, 1731–1739. 10.1016/j.ceramint.2016.08.207. [DOI] [Google Scholar]
  77. Zhang S.; Cheng X.; Yao Y.; Wei Y.; Han C.; Shi Y.; Wei Q.; Zhang Z. Porous niobium coatings fabricated with selective laser melting on titanium substrates: preparation, characterization, and cell behavior. Mater. Sci. Eng., C 2015, 53, 50–59. 10.1016/j.msec.2015.04.005. [DOI] [PubMed] [Google Scholar]
  78. Saranya K.; Thirupathi Kumara Raja S.; Subhasree R.S.; Gnanamani A.; Das S. K.; Rajendran N. Fabrication of nanoporous sodium niobate coating on 316L SS for orthopaedics. Ceram. Int. 2017, 43, 11569–11579. 10.1016/j.ceramint.2017.05.104. [DOI] [Google Scholar]
  79. Shi K.; Zhang Y.; Zhang J.; Xie Z. Electrochemical properties of niobium coating for biomedical application. Coatings 2019, 9, 546. 10.3390/coatings9090546. [DOI] [Google Scholar]
  80. Babaei K.; Fattah-alhosseini A.; Chaharmahali R. A review on plasma electrolytic oxidation (PEO) of niobium: Mechanism, properties and applications. Surf. Interfaces 2020, 21, 100719. 10.1016/j.surfin.2020.100719. [DOI] [Google Scholar]
  81. Stojadinović S.; Tadić N.; Radić N.; Stefanov P.; Grbić B.; Vasilić R. Anodic luminescence, structural, photoluminescent, and photocatalytic properties of anodic oxide films grown on niobium in phosphoric acid. Appl. Surf. Sci. 2015, 355, 912–920. 10.1016/j.apsusc.2015.07.174. [DOI] [Google Scholar]
  82. Chi-Hsiun Kuo P.; Chou H.-H.; Lin Y.-H.; Peng P.-W.; Ou K.-L.; Lee W.-R. Effects of surface functionalization on the nanostructure and biomechanical properties of binary titanium-niobium alloys. J. Electrochem. Soc. 2012, 159, E103 10.1149/2.094204jes. [DOI] [Google Scholar]
  83. Matsuno H.; Yokoyama A.; Watari F.; Uo M.; Kawasaki T. Biocompatibility and osteogenesis of refractory metal implants, titanium, hafnium, niobium, tantalum and rhenium. Biomaterials 2001, 22, 1253–1262. 10.1016/S0142-9612(00)00275-1. [DOI] [PubMed] [Google Scholar]
  84. Hryniewicz T.; Rokosz K.; Sandim H. Z. SEM/EDX and XPS studies of niobium after electropolishing. Appl. Surf. Sci. 2012, 263, 357–361. 10.1016/j.apsusc.2012.09.060. [DOI] [Google Scholar]
  85. Niinomi M. Recent metallic materials for biomedical applications. Metall. Mater. Trans. A 2002, 33, 477. 10.1007/s11661-002-0109-2. [DOI] [Google Scholar]
  86. Ge J.; Wang F.; Xu Z.; Shen X.; Gao C.; Wang D.; Hu G.; Gu J.; Tang T.; Wei J. Influences of niobium pentoxide on roughness, hydrophilicity, surface energy and protein absorption, and cellular responses to PEEK based composites for orthopedic applications. J. Mater. Chem. B 2020, 8, 2618–2626. 10.1039/C9TB02456E. [DOI] [PubMed] [Google Scholar]
  87. Samudrala R.; Abdul Azeem P.; Penugurti V.; Manavathi B. In vitro evaluation of niobia added soda lime borosilicate bioactive glasses. J. Alloys Compd. 2018, 764, 1072–1078. 10.1016/j.jallcom.2018.06.069. [DOI] [Google Scholar]
  88. Canepa P.; Ghiara G.; Spotorno R.; Canepa M.; Cavalleri O. Structural vs. electrochemical investigation of niobium oxide layers anodically grown in a Ca and P containing electrolyte. J. Alloys Compd. 2021, 851, 156937. 10.1016/j.jallcom.2020.156937. [DOI] [Google Scholar]
  89. Johansson C.; Albrektsson T. A removal torque and histomorphometric study of commercially pure niobium and titanium implants in rabbit bone. Clin. Oral Implants Res. 1991, 2, 24–29. 10.1034/j.1600-0501.1991.020103.x. [DOI] [PubMed] [Google Scholar]
  90. Nascimento W. J.; Bonadio T. G.; Freitas V. F.; Weinand W. R.; Baesso M. L.; Lima W. M. Nanostructured Nb2O5-natural hydroxyapatite formed by the mechanical alloying method: A bulk composite. Mater. Chem. Phys. 2011, 130 (1–2), 84–89. 10.1016/j.matchemphys.2011.05.069. [DOI] [Google Scholar]
  91. de Almeida Bino M. C.; Eurídice W. A.; Gelamo R. V.; Leite N. B.; da Silva M. V.; de Siervo A.; Pinto M. R.; de Almeida Buranello P. A.; Moreto J. A. Structural and morphological characterization of Ti6Al4V alloy surface functionalization based on Nb2O5 thin film for biomedical applications. Appl. Surf. Sci. 2021, 557, 149739. 10.1016/j.apsusc.2021.149739. [DOI] [Google Scholar]
  92. Alves A. R.; Coutinho A.d.R. The evolution of the niobium production in Brazil. Mater. Res. 2015, 18, 106–112. 10.1590/1516-1439.276414. [DOI] [Google Scholar]
  93. Schulz K. J., DeYoung J. H., Seal R. R.; Bradley D. C., Eds. Critical Mineral Resources of the United States: Economic and Environmental Geology and Prospects for Future Supply, Geological Survey, 2018.
  94. Nikishina E.; Drobot D.; Lebedeva E. Niobium and tantalum: State of the world market, fields of application, and raw sources. Part I. Russ. J. Non-Ferr. Met. 2013, 54, 446–452. 10.3103/S1067821213060187. [DOI] [Google Scholar]
  95. Vandrovcova M.; Jirka I.; Novotna K.; Lisa V.; Frank O.; Kolska Z.; Stary V.; Bacakova L. Interaction of human osteoblast-like Saos-2 and MG-63 cells with thermally oxidized surfaces of a titanium-niobium alloy. PloS one 2014, 9, e100475 10.1371/journal.pone.0100475. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Jirka I.; Vandrovcová M.; Frank O.; Tolde Z.; Plšek J.; Luxbacher T.; Bačáková L.; Starý V. On the role of Nb-related sites of an oxidized β-TiNb alloy surface in its interaction with osteoblast-like MG-63 cells. Mater. Sci. Eng., C 2013, 33, 1636–1645. 10.1016/j.msec.2012.12.073. [DOI] [PubMed] [Google Scholar]
  97. Robin A.; Rosa J. L. Corrosion behavior of niobium, tantalum and their alloys in hot hydrochloric and phosphoric acid solutions. Int. J. Refract. Hard Met. 2000, 18, 13–21. 10.1016/S0263-4368(99)00034-7. [DOI] [Google Scholar]
  98. Covino B. S. Jr.; Carter J. P.; Cramer S. D. The corrosion behavior of niobium in hydrochloric acid solutions. Corrosion 1980, 36, 554–558. 10.5006/0010-9312-36.10.554. [DOI] [Google Scholar]
  99. Manam N. S.; Harun W. S.; Shri D. N.; Ghani S. A.; Kurniawan T.; Ismail M. H.; Ibrahim M. H. Study of corrosion in biocompatible metals for implants: A review. J. Alloys Compd. 2017, 701, 698–715. 10.1016/j.jallcom.2017.01.196. [DOI] [Google Scholar]
  100. Masse J. P.; Szymanowski H.; Zabeida O.; Amassian A.; Klemberg-Sapieha J. E.; Martinu L. Stability and effect of annealing on the optical properties of plasma-deposited Ta2O5 and Nb2O5 films. Thin Solid Films 2006, 515, 1674–1682. 10.1016/j.tsf.2006.05.047. [DOI] [Google Scholar]
  101. Ramírez G.; Rodil S. E.; Muhl S.; Turcio-Ortega D.; Olaya J. J.; Rivera M.; Camps E.; Escobar-Alarcón L. Amorphous niobium oxide thin films. J. Non Cryst. Solids 2010, 356, 2714–2721. 10.1016/j.jnoncrysol.2010.09.073. [DOI] [Google Scholar]
  102. Usha N.; Sivakumar R.; Sanjeeviraja C.; Arivanandhan M. Niobium pentoxide (Nb2O5) thin films: rf Power and substrate temperature induced changes in physical properties. Optik (Stuttg.) 2015, 126, 1945–1950. 10.1016/j.ijleo.2015.05.036. [DOI] [Google Scholar]
  103. Lee S. H.; Kwon J. D.; Ahn J. H.; Park J. S. Compositional and electrical modulation of niobium oxide thin films deposited by plasma-enhanced atomic layer deposition. Ceram. Int. 2017, 43, 6580–6584. 10.1016/j.ceramint.2017.02.089. [DOI] [Google Scholar]
  104. Marques C.; Louro L. H. L.; Prado da Silva M. H. Bioactive ceramics based on Nb2O5 and Ta2O5. Key Eng. Mater. 2008, 396-398, 641–644. 10.4028/www.scientific.net/KEM.396-398.641. [DOI] [Google Scholar]
  105. Amaravathy P.; Sowndarya S.; Sathyanarayanan S.; Rajendran N. Novel sol gel coating of Nb2O5 on magnesium alloy for biomedical applications. Surf. Coat. Technol. 2014, 244, 131–141. 10.1016/j.surfcoat.2014.01.050. [DOI] [Google Scholar]
  106. Marins N. H.; Lee B. E.J.; e Silva R. M.; Raghavan A.; Villarreal Carreno N. L.; Grandfield K. Niobium pentoxide and hydroxyapatite particle loaded electrospun polycaprolactone/gelatin membranes for bone tissue engineering. Colloids Surf., B 2019, 182, 110386. 10.1016/j.colsurfb.2019.110386. [DOI] [PubMed] [Google Scholar]
  107. Demirkol N.; Oktar F. N.; Kayali E. S. Influence of niobium oxide on the mechanical properties of hydroxyapatite. Key Eng. Mater. 2012, 529-530, 29–33. 10.4028/www.scientific.net/KEM.529-530.29. [DOI] [Google Scholar]
  108. Turan M.; Demirkol N. Comparison of mechanical and bioactivity properties of nano hydroxyapatite-magnesium oxide and nano hydroxyapatite-niobium (V) oxide biocomposites. Adv. Nano-Bio. Mater. Dev. 2013, 3, 521–527. [Google Scholar]
  109. Li Y.; Munir K. S.; Lin J.; Wen C. Titanium-niobium pentoxide composites for biomedical applications. Bioact. Mater. 2016, 1, 127–131. 10.1016/j.bioactmat.2016.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Madhan Kumar A.; Nagarajan S.; Ramakrishna S.; Sudhagar P.; Kang Y. S.; Kim H.; Gasem Z. M.; Rajendran N. Electrochemical and in vitro bioactivity of polypyrrole/ceramic nanocomposite coatings on 316L SS bio-implants. Mater. Sci. Eng., C 2014, 43, 76–85. 10.1016/j.msec.2014.07.005. [DOI] [PubMed] [Google Scholar]
  111. Rani R. A.; Zoolfakar A. S.; O’Mullane A. P.; Austin M. W.; Kalantar-Zadeh K. Thin films and nanostructures of niobium pentoxide: fundamental properties, synthesis methods and applications. J. Mater. Chem. A 2014, 2, 15683–15703. 10.1039/C4TA02561J. [DOI] [Google Scholar]
  112. Singh C.; Pandey M. K.; Biradar A. M.; Srivastava A. K.; Sumana G. A bienzyme-immobilized highly efficient niobium oxide nanorod platform for biomedical application. RSC Adv. 2014, 4, 15458–15465. 10.1039/C4RA00215F. [DOI] [Google Scholar]
  113. Holtzberg F.; Reisman A.; Berry M.; Berkenblit M. Chemistry of the group VB pentoxides. VI. The polymorphism of Nb2O5. J. Am. Chem. Soc. 1957, 79, 2039–2043. 10.1021/ja01566a004. [DOI] [Google Scholar]
  114. Schäfer H.; Gruehn R.; Schulte F. The modifications of niobium pentoxide. Angew. Chem., Int. Ed. 1966, 5, 40–52. 10.1002/anie.196600401. [DOI] [Google Scholar]
  115. Guerreiro Tanomaru J. M.; Storto I.; Da Silva G. F.; Bosso R.; Costa B. C.; Bernardi M. I.; Tanomaru-Filho M. Radiopacity, pH and antimicrobial activity of Portland cement associated with micro-and nanoparticles of zirconium oxide and niobium oxide. Dent. Mater. J. 2014, 33, 466–470. 10.4012/dmj.2013-328. [DOI] [PubMed] [Google Scholar]
  116. Viapiana R.; Flumignan D. L.; Guerreiro-Tanomaru J. M.; Camilleri J.; Tanomaru-Filho M. Physicochemical and mechanical properties of zirconium oxide and niobium oxide modified P ortland cement-based experimental endodontic sealers. Int. Endod. J. 2014, 47, 437–448. 10.1111/iej.12167. [DOI] [PubMed] [Google Scholar]
  117. Mazur M.; Kalisz M.; Wojcieszak D.; Grobelny M.; Mazur P.; Kaczmarek D.; Domaradzki J. Determination of structural, mechanical and corrosion properties of Nb2O5 and (NbyCu1- y) Ox thin films deposited on Ti6Al4V alloy substrates for dental implant applications. Mater. Sci. Eng., C 2015, 47, 211–221. 10.1016/j.msec.2014.11.047. [DOI] [PubMed] [Google Scholar]
  118. Karlinsey R. L.; Hara A. T.; Yi K.; Duhn C. W. Bioactivity of novel self-assembled crystalline Nb2O5 microstructures in simulated and human salivas. Biomed. Mater. 2006, 1, 16. 10.1088/1748-6041/1/1/003. [DOI] [PubMed] [Google Scholar]
  119. Pradhan D.; Wren A. W.; Misture S. T.; Mellott N. P. Investigating the structure and biocompatibility of niobium and titanium oxides as coatings for orthopedic metallic implants. Mater. Sci. Eng., C 2016, 58, 918–926. 10.1016/j.msec.2015.09.059. [DOI] [PubMed] [Google Scholar]
  120. Pereira B. L.; Lepienski C. M.; Mazzaro I.; Kuromoto N. K. Apatite grown in niobium by two-step plasma electrolytic oxidation. Mater. Sci. Eng., C 2017, 77, 1235–1241. 10.1016/j.msec.2016.10.073. [DOI] [PubMed] [Google Scholar]
  121. Eisenbarth E.; Velten D.; Müller M.; Thull R.; Breme J. Nanostructured niobium oxide coatings influence osteoblast adhesion. J. Biomed. Mater. Res. 2006, 79A, 166–175. 10.1002/jbm.a.30823. [DOI] [PubMed] [Google Scholar]
  122. Grobelny M.; Kalisz M.; Mazur M.; Wojcieszak D.; Kaczmarek D. A.; Domaradzki J. A.; Świniarski M.; Mazur P. Functional Nb2O5 film and Nb2O5+ CuO, Nb2O5+ Graphene, Nb2O5+ CuO+ Graphene composite films to modify the properties of Ti6Al4V titanium alloy. Thin Solid Films 2016, 616, 64–72. 10.1016/j.tsf.2016.07.049. [DOI] [Google Scholar]
  123. Moreto J. A.; Gelamo R. V.; da Silva M. V.; Steffen T. T.; de Oliveira C. J.; de Almeida Buranello P. A.; Pinto M. R. New insights of Nb2O5-based coatings on the 316L SS surfaces: enhanced biological responses. J. Mater. Sci.: Mater. Med. 2021, 32, 25. 10.1007/s10856-021-06498-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Orozco-Hernández G.; Durán P. G.; Aperador W. Tribocorrosion Evaluation of Nb2O5, TiO2, and Nb2O5+ TiO2 Coatings for Medical Applications. Lubricants 2021, 9, 49. 10.3390/lubricants9050049. [DOI] [Google Scholar]
  125. Velten D.; Eisenbarth E.; Winter S.; Schanne N.; Breme J. Biomedical Oxide Coatings on Ti-alloys prepared by means of the Sol-Gel Process. BioNanoMaterials 2005, 6 (1), 22–28. 10.1515/BIOMAT.2005.6.1.22. [DOI] [Google Scholar]
  126. Pauline S. A.; Rajendran N. Biomimetic novel nanoporous niobium oxide coating for orthopaedic applications. Appl. Surf. Sci. 2014, 290, 448–457. 10.1016/j.apsusc.2013.11.112. [DOI] [PubMed] [Google Scholar]
  127. Nagarajan S.; Raman V.; Rajendran N. Synthesis and electrochemical characterization of porous niobium oxide coated 316L SS for orthopedic applications. Mater. Chem. Phys. 2010, 119, 363–366. 10.1016/j.matchemphys.2009.10.033. [DOI] [Google Scholar]
  128. Rafieerad A. R.; Bushroa A. R.; Nasiri-Tabrizi B.; Vadivelu J.; Yusof F.; Baradaran S. Graphene oxide modified anodic ternary nanobioceramics on Ti6Al7Nb alloy for orthopedic and dental applications. Procedia Eng. 2017, 184, 409–417. 10.1016/j.proeng.2017.04.111. [DOI] [Google Scholar]
  129. Mohseni E.; Zalnezhad E.; Bushroa A. R. Comparative investigation on the adhesion of hydroxyapatite coating on Ti-6Al-4V implant: A review paper. Int. J. Adhes. Adhes. 2014, 48, 238–257. 10.1016/j.ijadhadh.2013.09.030. [DOI] [Google Scholar]
  130. Priyadarshini B.; Rama M.; Chetan S.; Vijayalakshmi U. Bioactive coating as a surface modification technique for biocompatible metallic implants: a review. J. Asian Ceram. Soc. 2019, 7, 397–406. 10.1080/21870764.2019.1669861. [DOI] [Google Scholar]
  131. Aliofkhazraei M.; Macdonald D.D.; Matykina E.; Parfenov E.V.; Egorkin V.S.; Curran J.A.; Troughton S.C.; Sinebryukhov S.L.; Gnedenkov S.V.; Lampke T.; Simchen F.; Nabavi H.F. Review of plasma electrolytic oxidation of titanium substrates: Mechanism, properties, applications and limitations. Appl. Surf. Sci. Adv. 2021, 5, 100121. 10.1016/j.apsadv.2021.100121. [DOI] [Google Scholar]
  132. Raj V.; Rajaram M. P.; Balasubramanian G.; Vincent S.; Kanagaraj D. Pulse anodizing—an overview. Trans. IMF 2003, 81, 114–121. 10.1080/00202967.2003.11871515. [DOI] [Google Scholar]
  133. Safavi M. S.; Fathi M.; Charkhesht V.; Jafarpour M.; Ahadzadeh I. Electrodeposition of Co-P coatings reinforced by MoS2+Y2O3 hybrid ceramic nanoparticles for corrosion-resistant applications: influences of operational parameters. Metall. Mater. Trans. A 2020, 51, 6740–6758. 10.1007/s11661-020-05987-8. [DOI] [Google Scholar]
  134. Fathi M.; Safavi M. S.; Mahdavi S.; Mirzazadeh S.; Charkhesht V.; Mardanifar A.; Mehdipour M. Co-P alloy matrix composite deposits reinforced by nano-MoS2 solid lubricant: An alternative tribological coating to hard chromium coatings. Tribol. Int. 2021, 159, 106956. 10.1016/j.triboint.2021.106956. [DOI] [Google Scholar]
  135. Khalil-Allafi J.; Daneshvar H.; Safavi M. S.; Khalili V. A survey on crystallization kinetic behavior of direct current magnetron sputter deposited NiTi thin films. Phys. B: Condens. Matter. 2021, 615, 413086. 10.1016/j.physb.2021.413086. [DOI] [Google Scholar]
  136. Safavi M. S.; Walsh F. C. Electrodeposited Co-P alloy and composite coatings: A review of progress towards replacement of conventional hard chromium deposits. Surf. Coat. Technol. 2021, 422, 127564. 10.1016/j.surfcoat.2021.127564. [DOI] [Google Scholar]
  137. Safavi M. S.; Fathi M.; Ahadzadeh I. Feasible strategies for promoting the mechano-corrosion performance of Ni-Co based coatings: Which one is better?. Surf. Coat. Technol. 2021, 420, 127337. 10.1016/j.surfcoat.2021.127337. [DOI] [Google Scholar]
  138. Senocak T. C.; Ezirmik K. V.; Aysin F.; Simsek Ozek N.; Cengiz S. Niobium-oxynitride coatings for biomedical applications: Its antibacterial effects and in-vitro cytotoxicity. Mater. Sci. Eng., C 2021, 120, 111662. 10.1016/j.msec.2020.111662. [DOI] [PubMed] [Google Scholar]
  139. Kalisz M.; Grobelny M.; Mazur M.; Zdrojek M.; Wojcieszak D.; Świniarski M.; Judek J.; Kaczmarek D. Comparison of mechanical and corrosion properties of graphene monolayer on Ti-Al-V and nanometric Nb2O5 layer on Ti-Al-V alloy for dental implants applications. Thin Solid Films 2015, 589, 356–363. 10.1016/j.tsf.2015.05.059. [DOI] [Google Scholar]
  140. Kalisz M.; Grobelny M.; Mazur M.; Wojcieszak D.; Świniarski M.; Zdrojek M.; Domaradzki J.; Kaczmarek D. Mechanical and electrochemical properties of Nb2O5, Nb2O5: Cu and graphene layers deposited on titanium alloy (Ti6Al4V). Surf. Coat. Technol. 2015, 271, 92–99. 10.1016/j.surfcoat.2015.01.002. [DOI] [Google Scholar]
  141. Brinker C. J.; Lu Y.; Sellinger A.; Fan H. Evaporation-induced self-assembly: nanostructures made easy. Adv. Mater. 1999, 11, 579–585. . [DOI] [Google Scholar]
  142. Torre G. C.-d. l.; Espinosa-Medina M. A.; Martinez-Villafane A.; Gonzalez-Rodriguez J. G.; Castano V. M. Study of ceramic and hybrid coatings produced by the sol-gel method for corrosion protection. Open Corros. J. 2009, 2, 197–203. 10.2174/1876503300902010197. [DOI] [Google Scholar]
  143. Nijhawan S.; Bali P.; Gupta V. The international J. of oral implantology and clinical research. Int. J. Oral Implant. Clin. Res. 2010, 1, 77–82. 10.5005/jp-journals-10012-1012. [DOI] [Google Scholar]
  144. Karlinsey R. L. Preparation of self-organized niobium oxide microstructures via potentiostatic anodization. Electrochem. Commun. 2005, 7, 1190–1194. 10.1016/j.elecom.2005.08.027. [DOI] [Google Scholar]
  145. Karlinsey R. L. Self-assembled Nb2O5 microcones with tailored crystallinity. J. Mater. Sci. 2006, 41, 5017–5020. 10.1007/s10853-006-0135-3. [DOI] [Google Scholar]
  146. Mackey A. C.; Karlinsey R. L.; Chu T. M.; MacPherson M.; Alge D. L. Development of niobium oxide coatings on sand-blasted titanium alloy dental implants. Mater. Sci. Appl. 2012, 3 (5), 301–305. 10.4236/msa.2012.35044. [DOI] [Google Scholar]
  147. Lokeshkumar E.; Manojkumar P.; Saikiran A.; Premchand C.; Hariprasad S.; Rameshbabu N. Fabrication of Ca and P containing niobium oxide ceramic coatings on niobium by PEO coupled EPD process. Surf. Coat. Technol. 2021, 416, 127161. 10.1016/j.surfcoat.2021.127161. [DOI] [Google Scholar]
  148. Khalaf H. A.; Hussien B. M.; Al Zubaidi T. L. Influence of Antimicrobial Bio-Composite Coating on Osseointegration of Dental Implant. Health Sci. 2018, 7 (7), 110–118. [Google Scholar]
  149. Ramírez G.; Rodil S. E.; Arzate H.; Muhl S.; Olaya J. J. Niobium based coatings for dental implants. Appl. Surf. Sci. 2011, 257, 2555–2559. 10.1016/j.apsusc.2010.10.021. [DOI] [Google Scholar]
  150. Ye Y.; Lim R.; White J. M. High mobility amorphous zinc oxynitride semiconductor material for thin film transistors. J. Appl. Phys. 2009, 106, 074512. 10.1063/1.3236663. [DOI] [Google Scholar]
  151. Safavi M. S.; Rasooli A.; Sorkhabi F. Electrodeposition of Ni-P/Ni-Co-Al2O3 duplex nanocomposite coatings: towards improved mechanical and corrosion properties. Trans. IMF 2020, 98, 320–327. 10.1080/00202967.2020.1802106. [DOI] [Google Scholar]
  152. Fathi M.; Safavi M. S.; Mirzazadeh S.; Ansariyan A.; Ahadzadeh I. A promising horizon in mechanical and corrosion properties improvement of Ni-Mo coatings through incorporation of Y2O3 nanoparticles. Metall. Mater. Trans. A 2020, 51, 897–908. 10.1007/s11661-019-05559-5. [DOI] [Google Scholar]
  153. Rasooli A.; Safavi M. S.; Ahmadiyeh S.; Jalali A. Evaluation of TiO2 Nanoparticles Concentration and Applied Current Density Role in Determination of Microstructural, Mechanical, and Corrosion Properties of Ni-Co Alloy Coatings. Prot. Met. Phys. Chem. Surf. 2020, 56, 320–327. 10.1134/S2070205120020215. [DOI] [Google Scholar]
  154. Ross J. R.Contemporary Catalysis: Fundamentals and Current Applications; Elsevier, 2018. [Google Scholar]
  155. Eisenbarth E.; Velten D.; Breme J. Biomimetic implant coatings. Biomol. Eng. 2007, 24, 27–32. 10.1016/j.bioeng.2006.05.016. [DOI] [PubMed] [Google Scholar]
  156. Li J.; Jansen J. A.; Walboomers X. F.; van den Beucken J. J. Mechanical aspects of dental implants and osseointegration: A narrative review. J. Mech. Behav. Biomed. Mater. 2020, 103, 103574. 10.1016/j.jmbbm.2019.103574. [DOI] [PubMed] [Google Scholar]
  157. Prakasam M.; Locs J.; Salma-Ancane K.; Loca D.; Largeteau A.; Berzina-Cimdina L. Biodegradable materials and metallic implants—a review. J. Funct. Biomater. 2017, 8, 44. 10.3390/jfb8040044. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Cosson M.; Debodinance P.; Boukerrou M.; Chauvet M. P.; Lobry P.; Crépin G.; Ego A. Mechanical properties of synthetic implants used in the repair of prolapse and urinary incontinence in women: which is the ideal material?. Int. Urogynecol. J. 2003, 14, 169–178. 10.1007/s00192-003-1066-z. [DOI] [PubMed] [Google Scholar]
  159. Sumner D. Long-term implant fixation and stress-shielding in total hip replacement. J. Biomech. 2015, 48, 797–800. 10.1016/j.jbiomech.2014.12.021. [DOI] [PubMed] [Google Scholar]
  160. Lai Y. S.; Chen W. C.; Huang C. H.; Cheng C. K.; Chan K. K.; Chang T. K. The effect of graft strength on knee laxity and graft in-situ forces after posterior cruciate ligament reconstruction. PLoS One 2015, 10, e0127293 10.1371/journal.pone.0127293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  161. Swain S.; Misra R. D.; You C. K.; Rautray T. R. TiO2 nanotubes synthesised on Ti-6Al-4V ELI exhibits enhanced osteogenic activity: A potential next-generation material to be used as medical implants. Mater. Technol. 2021, 36, 393–399. 10.1080/10667857.2020.1760510. [DOI] [Google Scholar]
  162. Karacan I.; Ben-Nissan B.; Wang H. A.; Juritza A.; Swain M. V.; Müller W. H.; Chou J.; Stamboulis A.; Macha I. J.; Taraschi V. Mechanical testing of antimicrobial biocomposite coating on metallic medical implants as drug delivery system. Mater. Sci. Eng., C 2019, 104, 109757. 10.1016/j.msec.2019.109757. [DOI] [PubMed] [Google Scholar]
  163. Ibrahim M. Z.; Sarhan A. A.; Kuo T. Y.; Yusof F.; Hamdi M.; Lee T. M. Developing a new laser cladded FeCrMoCB metallic glass layer on nickel-free stainless-steel as a potential superior wear-resistant coating for joint replacement implants. Surf. Coat. Technol. 2020, 392, 125755. 10.1016/j.surfcoat.2020.125755. [DOI] [Google Scholar]
  164. Cheng K. Y.; Gopal V.; McNallan M.; Manivasagam G.; Mathew M. T. Enhanced tribocorrosion resistance of hard ceramic coated Ti-6Al-4V alloy for hip implant application: in-vitro simulation study. ACS Biomater. Sci. Eng. 2019, 5, 4817–4824. 10.1021/acsbiomaterials.9b00609. [DOI] [PubMed] [Google Scholar]
  165. Goldberg J. R.; Gilbert J. L. The electrochemical and mechanical behavior of passivated and TiN/AlN-coated CoCrMo and Ti6Al4V alloys. Biomaterials 2004, 25, 851–864. 10.1016/S0142-9612(03)00606-9. [DOI] [PubMed] [Google Scholar]
  166. Nakagawa M.; Matsuya S.; Shiraishi T.; Ohta M. Effect of fluoride concentration and pH on corrosion behavior of titanium for dental use. J. Dent. Res. 1999, 78, 1568–1572. 10.1177/00220345990780091201. [DOI] [PubMed] [Google Scholar]
  167. Nagarajan S.; Rajendran N. Surface characterisation and electrochemical behaviour of porous titanium dioxide coated 316L stainless steel for orthopaedic applications. Appl. Surf. Sci. 2009, 255, 3927–3932. 10.1016/j.apsusc.2008.10.058. [DOI] [Google Scholar]
  168. Starikov V. V.; Starikova S. L.; Mamalis A. G.; Lavrynenko S. N.; Ramsden J. J. The application of niobium and tantalum oxides for implant surface passivation. J. Biol. Phys. Chem. 2007, 7 (4), 141–145. 10.4024/40704.jbpc.07.04. [DOI] [Google Scholar]
  169. Sowa M.; Kazek-Kesik A.; Krzakała A.; Socha R. P.; Dercz G.; Michalska J.; Simka W. Modification of niobium surfaces using plasma electrolytic oxidation in silicate solutions. J. Solid State Electrochem. 2014, 18, 3129–3142. 10.1007/s10008-013-2341-7. [DOI] [Google Scholar]
  170. Souza J. C.; Henriques M.; Oliveira R.; Teughels W.; Celis J. P.; Rocha L. A. Do oral biofilms influence the wear and corrosion behavior of titanium?. Biofouling 2010, 26, 471–478. 10.1080/08927011003767985. [DOI] [PubMed] [Google Scholar]
  171. Souza J. C.; Barbosa S. L.; Ariza E. A.; Henriques M.; Teughels W.; Ponthiaux P.; Celis J. P.; Rocha L. A. How do titanium and Ti6Al4V corrode in fluoridated medium as found in the oral cavity? An in vitro study. Mater. Sci. Eng., C 2015, 47, 384–393. 10.1016/j.msec.2014.11.055. [DOI] [PubMed] [Google Scholar]
  172. Ramel C. F.; Lüssi A.; Özcan M.; Jung R. E.; Hämmerle C. H.; Thoma D. S. Surface roughness of dental implants and treatment time using six different implantoplasty procedures. Clin. Oral Implants Res. 2016, 27, 776–781. 10.1111/clr.12682. [DOI] [PubMed] [Google Scholar]
  173. Lin T. H.; Tamaki Y.; Pajarinen J.; Waters H. A.; Woo D. K.; Yao Z.; Goodman S. B. Chronic inflammation in biomaterial-induced periprosthetic osteolysis: NF-κB as a therapeutic target. Acta Biomater. 2014, 10, 1–10. 10.1016/j.actbio.2013.09.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Kuwabara T.; Ishikawa F.; Kondo M.; Kakiuchi T. The role of IL-17 and related cytokines in inflammatory autoimmune diseases. Mediators Inflamm. 2017, 2017, 3908061. 10.1155/2017/3908061. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Weng J.; Liu Q.; Wolke J. G.; Zhang X.; de Groot K. Formation and characteristics of the apatite layer on plasma-sprayed hydroxyapatite coatings in simulated body fluid. Biomaterials 1997, 18, 1027–1035. 10.1016/S0142-9612(97)00022-7. [DOI] [PubMed] [Google Scholar]
  176. Kokubo T.; Takadama H. How useful is SBF in predicting in vivo bone bioactivity?. Biomaterials 2006, 27, 2907–2915. 10.1016/j.biomaterials.2006.01.017. [DOI] [PubMed] [Google Scholar]
  177. Kokubo T. Design of bioactive bone substitutes based on biomineralization process. Mater. Sci. Eng., C 2005, 25, 97–104. 10.1016/j.msec.2005.01.002. [DOI] [Google Scholar]
  178. Kokubo T.; Matsushita T.; Takadama H.; Kizuki T. Development of bioactive materials based on surface chemistry. J. Eur. Ceram. Soc. 2009, 29, 1267–1274. 10.1016/j.jeurceramsoc.2008.08.004. [DOI] [Google Scholar]
  179. Singh G.; Singh H.; Sidhu B. S. Corrosion behavior of plasma sprayed hydroxyapatite and hydroxyapatite-silicon oxide coatings on AISI 304 for biomedical application. Appl. Surf. Sci. 2013, 284, 811–818. 10.1016/j.apsusc.2013.08.013. [DOI] [Google Scholar]
  180. Ponsonnet L.; Reybier K.; Jaffrezic N.; Comte V.; Lagneau C.; Lissac M.; Martelet C. Relationship between surface properties (roughness, wettability) of titanium and titanium alloys and cell behaviour. Mater. Sci. Eng., C 2003, 23, 551–560. 10.1016/S0928-4931(03)00033-X. [DOI] [Google Scholar]
  181. Harvey A. G.; Hill E. W.; Bayat A. Designing implant surface topography for improved biocompatibility. Expert Rev. Med. Devices 2013, 10, 257–267. 10.1586/erd.12.82. [DOI] [PubMed] [Google Scholar]
  182. Xie W.; Wang J.; Berndt C. C. Ethylene methacrylic acid (EMAA) single splat morphology. Coatings 2013, 3, 82–97. 10.3390/coatings3020082. [DOI] [Google Scholar]
  183. Jung U. W.; Hwang J. W.; Choi D. Y.; Hu K. S.; Kwon M. K.; Choi S. H.; Kim H. J. Surface characteristics of a novel hydroxyapatite-coated dental implant. J. Periodontal Implant Sci. 2012, 42, 59–63. 10.5051/jpis.2012.42.2.59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Shekaran A.; García A. J. Extracellular matrix-mimetic adhesive biomaterials for bone repair. J. Biomed. Mater. Res., Part A 2011, 96, 261–272. 10.1002/jbm.a.32979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Chai F.; Ochsenbein A.; Traisnel M.; Busch R.; Breme J.; Hildebrand H. F. Improving endothelial cell adhesion and proliferation on titanium by sol-gel derived oxide coating. J. Biomed. Mater. Res., Part A 2009, 92A, 754–765. 10.1002/jbm.a.32399. [DOI] [PubMed] [Google Scholar]
  186. Rohanová D.; Boccaccini A. R.; Horkavcová D.; Bozdechová P.; Bezdička P.; Castorálová M. Is non-buffered DMEM solution a suitable medium for in vitro bioactivity tests?. J. Mater. Chem. B 2014, 2, 5068–5076. 10.1039/C4TB00187G. [DOI] [PubMed] [Google Scholar]
  187. Shokouhinejad N.; Nekoofar M. H.; Razmi H.; Sajadi S.; Davies T. E.; Saghiri M. A.; Gorjestani H.; Dummer P. M. Bioactivity of EndoSequence root repair material and bioaggregate. Int. Endod. J. 2012, 45, 1127–1134. 10.1111/j.1365-2591.2012.02083.x. [DOI] [PubMed] [Google Scholar]
  188. Vitorino R.; Lobo M. J.; Ferrer-Correira A. J.; Dubin J. R.; Tomer K. B.; Domingues P. M.; Amado F. M. Identification of human whole saliva protein components using proteomics. Proteomics 2004, 4, 1109–1115. 10.1002/pmic.200300638. [DOI] [PubMed] [Google Scholar]
  189. Tamaki N.; Tada T.; Morita M.; Watanabe T. Comparison of inhibitory activity on calcium phosphate precipitation by acidic proline-rich proteins, statherin, and histatin-1. Calcif. Tissue Int. 2002, 71, 59. 10.1007/s00223-001-1084-0. [DOI] [PubMed] [Google Scholar]
  190. Eanes E.; Gillessen I.; Posner A. Intermediate states in the precipitation of hydroxyapatite. Nature 1965, 208, 365–367. 10.1038/208365a0. [DOI] [PubMed] [Google Scholar]
  191. Katta P. P.; Nalliyan R. Corrosion resistance with self-healing behavior and biocompatibility of Ce incorporated niobium oxide coated 316L SS for orthopedic applications. Surf. Coat. Technol. 2019, 375, 715–726. 10.1016/j.surfcoat.2019.07.042. [DOI] [Google Scholar]
  192. Pradeep PremKumar K.; Duraipandy N.; Manikantan Syamala K.; Rajendran N. Antibacterial effects, biocompatibility and electrochemical behavior of zinc incorporated niobium oxide coating on 316L SS for biomedical applications. Appl. Surf. Sci. 2018, 427, 1166–1181. 10.1016/j.apsusc.2017.08.221. [DOI] [Google Scholar]
  193. Sharkeev Y.; Komarova E.; Sedelnikova M.; Sun Z. M.; Zhu Q. F.; Zhang J.; Tolkacheva T.; Uvarkin P. Structure and properties of micro-arc calcium phosphate coatings on pure titanium and Ti-40Nb alloy. Trans. Nonferrous Met. Soc. China 2017, 27, 125–133. 10.1016/S1003-6326(17)60014-1. [DOI] [Google Scholar]
  194. Enayati M.; Fathi M.; Zomorodian A. Characterisation and corrosion properties of novel hydroxyapatite niobium plasma sprayed coating. Surf. Eng. 2009, 25, 338–342. 10.1179/026708408X341214. [DOI] [Google Scholar]
  195. Jacobs A.; Renaudin G.; Forestier C.; Nedelec J.-M.; Descamps S. Biological properties of copper-doped biomaterials for orthopedic applications: A review of antibacterial, angiogenic and osteogenic aspects. Acta Biomater. 2020, 117, 21–39. 10.1016/j.actbio.2020.09.044. [DOI] [PubMed] [Google Scholar]
  196. Chellan P.; Sadler P. J. The elements of life and medicines. Philos. Trans. Royal Soc. 2015, 373, 20140182. 10.1098/rsta.2014.0182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Lowe N. M.; Fraser W. D.; Jackson M. J. Is there a potential therapeutic value of copper and zinc for osteoporosis?. Proc. Nutr. Soc. 2002, 61, 181–185. 10.1079/PNS2002154. [DOI] [PubMed] [Google Scholar]
  198. Rojas P.; Rodil S. Corrosion behaviour of amorphous niobium oxide coatings. Int. J. Electrochem. Sci. 2012, 7, 1443–1458. [Google Scholar]
  199. Arsova I.; Prusi A.; Grčev T.; Arsov L. Electrochemical characterization of the passive films formed on niobium surfaces in H2SO4 solutions. J. Serbian Chem. 2006, 71, 177–187. 10.2298/JSC0602177A. [DOI] [Google Scholar]
  200. Gagnon J.; Fromm K. M. Toxicity and protective effects of cerium oxide nanoparticles (nanoceria) depending on their preparation method, particle size, cell type, and exposure route. Eur. J. Inorg. Chem. 2015, 2015, 4510–4517. 10.1002/ejic.201500643. [DOI] [Google Scholar]
  201. Zhou G.; Gu G.; Li Y.; Zhang Q.; Wang W.; Wang S.; Zhang J. Effects of cerium oxide nanoparticles on the proliferation, differentiation, and mineralization function of primary osteoblasts in vitro. Biol. Trace Elem. Res. 2013, 153, 411–418. 10.1007/s12011-013-9655-2. [DOI] [PubMed] [Google Scholar]
  202. Niu J.; Wang K.; Kolattukudy P. E. Cerium oxide nanoparticles inhibits oxidative stress and nuclear factor-κB activation in H9c2 cardiomyocytes exposed to cigarette smoke extract. J. Pharmacol. Exp. Ther. 2011, 338, 53–61. 10.1124/jpet.111.179978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Korsvik C.; Patil S.; Seal S.; Self W. T. Superoxide dismutase mimetic properties exhibited by vacancy engineered ceria nanoparticles. Chem. Commun. 2007, 2007, 1056–1058. 10.1039/b615134e. [DOI] [PubMed] [Google Scholar]
  204. Hu H.; Zhang W.; Qiao Y. Q.; Jiang X. Y.; Liu X.; Ding C. Antibacterial activity and increased bone marrow stem cell functions of Zn-incorporated TiO2 coatings on titanium. Acta Biomater. 2012, 8, 904–915. 10.1016/j.actbio.2011.09.031. [DOI] [PubMed] [Google Scholar]
  205. Applerot G.; Lipovsky A.; Dror R.; Perkas N.; Nitzan Y.; Lubart R.; Gedanken A. Enhanced antibacterial activity of nanocrystalline ZnO due to increased ROS-mediated cell injury. Adv. Funct. Mater. 2009, 19, 842–852. 10.1002/adfm.200801081. [DOI] [Google Scholar]
  206. Rai M.; Kon K.; Gade A.; Ingle A.; Nagaonkar D.; Paralikar P.; da Silva S. S. Antibiotic Resistance: Can Nanoparticles Tackle the Problem? In Antibiotic Resistance; Academic Press, 2016; pp 121–143, 10.1016/B978-0-12-803642-6.00006-X. [DOI] [Google Scholar]
  207. Devrim B.; Bozkır A. Nanocarriers and Their Potential Application as Antimicrobial Drug Delivery. In Nanostructures for Antimicrobial Therapy; Elsevier, 2017; pp 169–202. [Google Scholar]
  208. Grigore M. E.; Biscu E. R.; Holban A. M.; Gestal M. C.; Grumezescu A. M. Methods of synthesis, properties and biomedical applications of CuO nanoparticles. Pharmaceuticals 2016, 9, 75. 10.3390/ph9040075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Perreault F.; Melegari S. P.; da Costa C. H.; de Oliveira Franco Rossetto A. L.; Popovic R.; Matias W. G. Genotoxic effects of copper oxide nanoparticles in Neuro 2A cell cultures. Sci. Total Environ. 2012, 441, 117–124. 10.1016/j.scitotenv.2012.09.065. [DOI] [PubMed] [Google Scholar]
  210. Baek Y. W.; An Y. J. Microbial toxicity of metal oxide nanoparticles (CuO, NiO, ZnO, and Sb2O3) to Escherichia coli, Bacillus subtilis, and Streptococcus aureus. Sci. Total Environ. 2011, 409, 1603–1608. 10.1016/j.scitotenv.2011.01.014. [DOI] [PubMed] [Google Scholar]
  211. Ruiz P.; Katsumiti A.; Nieto J. A.; Bori J.; Jimeno-Romero A.; Reip P.; Arostegui I.; Orbea A.; Cajaraville M. P. Short-term effects on antioxidant enzymes and long-term genotoxic and carcinogenic potential of CuO nanoparticles compared to bulk CuO and ionic copper in mussels Mytilus galloprovincialis. Mar. Environ. Res. 2015, 111, 107–120. 10.1016/j.marenvres.2015.07.018. [DOI] [PubMed] [Google Scholar]
  212. Sikka M. P.; Midha V. K. The Role of Biopolymers and Biodegradable Polymeric Dressings in Managing Chronic Wounds. In Advanced Textiles for Wound Care; Elsevier, 2019; pp 463–488. [Google Scholar]
  213. Gera S.; Kankuri E.; Kogermann K. Antimicrobial peptides-Unleashing their therapeutic potential using nanotechnology. Pharmacol. Ther. 2021, 107990. 10.1016/j.pharmthera.2021.107990. [DOI] [PubMed] [Google Scholar]
  214. Schafhauser B. H.; Kristofco L. A.; de Oliveira C. M.; Brooks B. W. Global review and analysis of erythromycin in the environment: occurrence, bioaccumulation and antibiotic resistance hazards. Environ. Pollut. 2018, 238, 440–451. 10.1016/j.envpol.2018.03.052. [DOI] [PubMed] [Google Scholar]
  215. Munckhof W. J.; Borlace G.; Turnidge J. D. Postantibiotic suppression of growth of erythromycin A-susceptible and-resistant gram-positive bacteria by the ketolides telithromycin (HMR 3647) and HMR 3004. Antimicrob. Agents Chemother. 2000, 44, 1749–1753. 10.1128/AAC.44.6.1749-1753.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Pokkunuri I.; Champney W. S. Characteristics of a 50S ribosomal subunit precursor particle as a substrate for ermE methyltransferase activity and erythromycin binding in Staphylococcus aureus. RNA boil. 2007, 4, 147–153. 10.4161/rna.4.3.5346. [DOI] [PubMed] [Google Scholar]

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES