Next Article in Journal
Efficacy of Pegylated Hyaluronic Acid Filler Enriched with Calcium Hydroxyapatite: A 24-Week Post-Market, Observational, Prospective, Open-Label, Single-Center Study
Next Article in Special Issue
Antibacterial Biomaterial Based on Bioglass Modified with Copper for Implants Coating
Previous Article in Journal
Fabrication and Evaluation of Porous dECM/PCL Scaffolds for Bone Tissue Engineering
Previous Article in Special Issue
Green Synthesis of Controlled Shape Silver Nanostructures and Their Peroxidase, Catalytic Degradation, and Antibacterial Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Engineering Antioxidant Surfaces for Titanium-Based Metallic Biomaterials

by
Jithin Vishnu
1,
Praveenkumar Kesavan
2,
Balakrishnan Shankar
1,
Katarzyna Dembińska
3,
Maria Swiontek Brzezinska
3 and
Beata Kaczmarek-Szczepańska
4,*
1
Department of Mechanical Engineering, Amrita Vishwa Vidyapeetham, Amritapuri Campus, Clappana 690525, India
2
Department of Materials Engineering, Indian Institute of Science, Bangalore 560012, India
3
Department of Environmental Microbiology and Biotechnology, Faculty of Biological and Veterinary Sciences, Nicolaus Copernicus University in Toruń, 87-100 Toruń, Poland
4
Department of Biomaterials and Cosmetic Chemistry, Faculty of Chemistry, Nicolaus Copernicus University in Toruń, 87-100 Toruń, Poland
*
Author to whom correspondence should be addressed.
J. Funct. Biomater. 2023, 14(7), 344; https://doi.org/10.3390/jfb14070344
Submission received: 2 June 2023 / Revised: 21 June 2023 / Accepted: 27 June 2023 / Published: 29 June 2023

Abstract

:
Prolonged inflammation induced by orthopedic metallic implants can critically affect the success rates, which can even lead to aseptic loosening and consequent implant failure. In the case of adverse clinical conditions involving osteoporosis, orthopedic trauma and implant corrosion-wear in peri-implant region, the reactive oxygen species (ROS) activity is enhanced which leads to increased oxidative stress. Metallic implant materials (such as titanium and its alloys) can induce increased amount of ROS, thereby critically influencing the healing process. This will consequently affect the bone remodeling process and increase healing time. The current review explores the ROS generation aspects associated with Ti-based metallic biomaterials and the various surface modification strategies developed specifically to improve antioxidant aspects of Ti surfaces. The initial part of this review explores the ROS generation associated with Ti implant materials and the associated ROS metabolism resulting in the formation of superoxide anion, hydroxyl radical and hydrogen peroxide radicals. This is followed by a comprehensive overview of various organic and inorganic coatings/materials for effective antioxidant surfaces and outlook in this research direction. Overall, this review highlights the critical need to consider the aspects of ROS generation as well as oxidative stress while designing an implant material and its effective surface engineering.

Graphical Abstract

1. Introduction

One of the key factors associated with inflammatory response is the oxidative stress, which is characterized by the imbalance/disparity between the generation of reactive oxygen species (ROS) and antioxidant defense system [1]. Osteoporosis, the most common bone disorder globally, is a systemic skeletal disorder associated with diminishing bone mass and micro-architectural bone tissue degradation with concomitant bone fragility and osteoporotic fracture [2,3,4]. Considered as one of the major global pandemics of the 21st century, osteoporosis induces more than 8.9 million bone fractures per annum, affecting about 200 million people, and in addition, poses a high risk specifically to post-menopausal women, with 40–50% prevalence in women older than 60 years [5,6,7]. Other leading causes for bone fracture includes road accidents, falls and sports injuries. Following the bone fracture, secondary healing ensues involving various stages such as hematoma formation, acute inflammation, callus formation and bone remodeling [8]. The fracture trauma results in blood vessel rupture in the region of fracture leading to hematoma [9]. The hematoma micro-environment in the fracture site is structurally unstable, hypoxic and acidic, which requires a cross-talk between inflammatory cells and cells related to bone healing in order to re-establish normal homeostatic state [9,10]. The bone remodeling involves a collective involvement of various bone cells such as osteoclasts (removal of damaged and old bones), osteoblasts (synthesis and secretion of osteoid matrix during mineralization) and osteocytes (regulating new bone formation and old bone resorption) [11,12,13]. Oxidative stress is a predominant factor which negatively affects the bone remodeling process resulting in a deteriorated bone mineral density, contributing in the etiology of osteoporosis [14,15,16].
Clinical intervention of bone fractures and defects considers the usage of orthopedic implants for the treatment of orthopedic trauma with minimal harm to the patients. Metallic, ceramic and polymeric biomaterials have been explored and researched for orthopedic implant applications, with each class of materials possessing its own advantages and disadvantages [17,18,19]. Thanks to their superior mechanical properties, metallic materials are the most widely used material for internal fracture fixation components. The three dominant material classes in this aspect are 316L stainless steel, Co-Cr alloys and Ti and its alloys [20,21,22,23]. However, metallic materials are prone to degradation due to corrosion-wear synergy (tribocorrosion) in complex physiological environments capable of eliciting the release of ions and debris in the peri-implant region [24,25]. Such wear-debris release from articulating components can result in the activation and senescence of resident cells including macrophages, fibroblasts, osteoclasts and osteoblasts, eventually leading to the production and release of pro-inflammatory cytokines, chemokines, ROS and reactive nitrogen species (RNS) [26,27]. This elicits chronic inflammatory cascades and oxidative stress reactions eventually resulting in bone resorption and osteolysis induced implant failure [26,28].
During normal healing process, osteoblasts express antioxidant enzymes such as superoxide dismutase (SOD) for inducing the conversion of ROS into O and H2O to induce osteoblast differentiation [29,30]. However, during adverse conditions as mentioned above, enhanced ROS activity results in oxidative stress reducing the bone mineralization density by affecting the remodeling process [31,32]. In spite of the presence of endogenous antioxidants, excessive generation of free radicals and inflammatory processes result in oxidative stress [33]. The occurrence of oxidative stress can be ascribed to abnormal activation of enzymes which generates ROS. ROS are highly reactive, short-lived molecules formed as by-products during molecular oxygen reduction which are capable of oxidative damage to macromolecules in biological cells [34,35,36,37,38]. ROS include radical and non-radical oxygen species such as superoxide anion (O2−), hydroxyl radical (OH) and hydrogen peroxide (H2O2) [39,40,41,42]. The mechanism of ROS formation via electrochemical corrosion reaction, radical transformation via Fenton and Haber–Weiss Reactions, light induction and surface catalytic reactions is elaborately reviewed by Kessier et al. [43].
Antioxidants are naturally occurring reducing agents which can hinder the generation of ROS via the phenomenon of scavenging free radicals and eradicating ROS derivatives. Hence, the origin of oxidative stress can be linked to the imbalance between ROS and antioxidants which encompass enzymatic antioxidants (e.g., polyphenols, carotenoids, glutathione, tocopherols) and antioxidant enzymes (SOD, catalase, glutathione peroxidase) [44]. An increment in antioxidant levels can be potentially harmful as it could induce molecular damages, apoptosis or necrosis, and oxidative stress is found to be associated with several diseases including cardiovascular, neurodegenerative, carcinoma, diabetes, ischemia/reperfusion injury, rheumatoid arthritis and aging [45]. Endogenous enzymatic antioxidants include SOD, catalase, glutathione peroxidase and glutathione reductase, whereas non-enzymatic endogenous antioxidants include glutathione and lipoic acid [46].
Metallic implant materials are widely used for bone-anchored therapy for orthopedic and dental treatments. Apart from the wear-induced oxidative stress as discussed above, metallic material insertion during surgical procedure induces large amount of ROS generation and is incapable of generating antioxidants, thereby critically influencing the healing process which elevates the healing time.
Implant surface plays a pivotal role in dictating the host response of the implanted material. In most cases, surface modification of implants alters the surface morphology, topography, chemistry and surface energy, particularly aimed at improving matrix protein adhesion, cellular adhesion and proliferation, to attain better osseointegration [47,48]. A variety of surface modification strategies involving surface texturing and surface coatings have been developed to improve the interfacial mechanical strength, wear resistance, tribo-corrosion resistance and biocompatibility in order to enhance the longevity of orthopedic implants [49,50]. Recently, surface modification of Ti implants has been gaining research attention to repair the impaired osseointegration by developing surfaces with antioxidant activity [51,52,53]. In summary, it is imperative to gain more insights into the advancements in this field to further improve the antioxidant activities of Ti implant surface by proper surface modification to improve its clinical efficiency. In view of these aspects, the present review is focused towards the various surface engineering techniques to combat the undesirable ROS generation associated with Ti-based metallic implants. Several review articles have been published reporting the underlying mechanism of ROS formation and antioxidative mechanisms [30,54,55]. In addition, review articles comprehensively describing surface modification techniques for Ti surface are published [56,57]. The novelty aspect of the present review lies in collating the available reported works in improving the antioxidant properties of Ti-based metallic implant surfaces via various organic and inorganic coatings. Even though several research articles have explored the antioxidant activity of surfaces developed for antibacterial and biocompatible applications, this review exempts these articles and is focused on research associated with surfaces/materials specifically developed for antioxidant purpose. This review initially presents an outline of ROS generation associated with the insertion of Ti implants. This is followed by sections describing organic- and inorganic-based coatings on Ti surfaces to ameliorate the antioxidant aspects along with prospective future perspectives. The major objective of the present review is to provide an overall idea about how surface modification can assist in improving the ROS scavenging activity and reduce oxidative damage to improve the clinical efficiency of Ti-based implants.

2. Titanium Alloys and Reactive Oxygen Species Metabolism

Titanium (Ti) and its alloys are the widely used material for a variety of load-bearing orthopedic implant applications thanks to the excellent mechanical aspects, lower modulus values, corrosion resistance and excellent biocompatibility [58]. Ti is a transition metal which exists in a hexagonal closed pack (hcp) crystal structure (α-Ti), which transforms into its allotropic form with a body-centered cubic (bcc) structure (β-Ti) above a temperature of 882 °C, which is retained up to its melting point (1688 °C). Several Ti-based alloys such as commercially pure Ti (cp-Ti, ASTM-F67), Ti-6Al-4V (ASTM-F136), Ti-6Al-7Nb (ASTM-F1472, F1295) and Ti-13Nb-13Zr (ASTM-F1713-08) have been explored for dental implants, bone fixation plates, screws and hip joint stems [59,60]. Current research focus is more shifted towards β-Ti alloys as they possess comparatively lower elastic modulus (as low as 46–55 GPa), high strength, good cold workability and, most importantly, the beneficial biocompatibility aspects due to β-phase stabilizing alloying additions (Nb, Ta, Mo, Mn, Fe etc.) [61,62,63]. In addition, Ti-based shape memory alloys are prospective materials for various biomedical applications owing to the shape memory and super-elasticity effects [64]. Despite these beneficial aspects, wear-induced aseptic loosening is a limiting factor hampering the efficiency of Ti-based orthopedic implants [65]. Wear-particle phagocytosis by macrophages can induce cytokine and free radical release, resulting in an aseptic inflammatory response, capable of promoting osteoclast resorption [66]. The role of high oxidative stress as one of the main causative factors in various inflammatory and degenerative disorders points towards the contribution of ROS towards aseptic loosening. As a response to the released metallic particles in a physiological condition, the immune system elicits an inflammation process, which involves generation of ROS through a series of enzyme-assisted biochemical reactions (schematic figure as shown in Figure 1) [67].
Superoxide radical generation is catalyzed by NADPH (nicotinamide adenine dinucleotide phosphate) oxidase (Equation (1)). Electrons from NADPH is accepted by the cytosolic domain of gp91phox (electron transferase of NADPH oxidase) and is transferred across membrane to O2 to generate superoxide radical (O2) as the primary product [68]. Gp91phox contains all the required co-factors to effectuate electron transfer reaction, in which electrons transfer from NADPH onto flavin adenine dinucleotide (FAD) and to the haem group in the following step, inducing reduction of O2 to O2 [69].
N A D P H + O 2 N A D P + + O 2 + H +
In response to this, antioxidant scavenging enzymes such as SOD promote dismutation to convert superoxide to hydrogen peroxide and an oxygen molecule (Equation (2)), which occurs spontaneously (rate constant = 5 × 105 M−1s−1 at neutral pH) [70]. This reaction is greatly accelerated by SOD, and the corresponding catalytic activity is attributed partly to the electrostatic interactions in active center of SOD protein [71].
2 O 2 + 2 H + O 2 + H 2 O 2
Stimulation of neutrophils results in oxygen consumption in a respiratory burst that produces O2 and H2O2. Simultaneous discharge of abundant myeloperoxidase enzyme occurs, which utilizes H2O2 to oxidize halides (chlorides, bromides) and thiocyanates to corresponding hypohalous acids and hypothiocyanite [72]. Myeloperoxidase, also called verdoperoxidase, is a heme-containing peroxidase generated mostly from polymorphonuclear neutrophils and found in primary granules of granulocytic cells [73]. The reaction between hydrogen peroxide with halides (such as Cl in physiological environment) is catalyzed by granule-localized myeloperoxidase to form hypochlorous acid (bleach) (Equation (3)).
H 2 O 2 + C l H O C l + O H
In addition, hydrogen peroxide can generate hydroxide and hydroperoxyl radicals by reacting with ferrous and ferric cations (Fenton reactions). Fenton chemistry can significantly enhance the degradation of many transition metals (including Ti alloys, Co-Cr alloys) [74]. Fenton reaction involves an initial electron transfer with neither bond formation nor breaking and the generation of hydroxyl radicals [75]. Haber Weiss reaction which makes use of Fenton chemistry involves vital mechanism in which highly reactive hydroxyl radical generation occurs [76]. Another possible cathodic reaction taking place at implant/bone interface is oxygen reduction to generate hydrogen peroxide (Equation (4)). The cathodic oxygen reduction can be sub-divided into several reactions, resulting in the generation of hydroxyl radicals and hydrogen peroxide.
O 2 + 2 H 2 O + 2 e H 2 O 2 + 2 O H
Hence, ROS are additional products of overall electrochemical reactions occurring in the implant interface other than the metallic ions and/or particles. The presence of ROS (hydroxyl radicals and hydrogen peroxide) can further promote degradation of Ti implants [77]. Among the various ROS molecules, hydrogen peroxide can mix with water and diffuse through membranes of peri-implant tissues, critically affecting intracellular redox status, thereby increasing the chances of implant failure [78].

3. Surface Modification for Antioxidant Ti Surfaces

Surface modification of Ti alloys offers an effective strategy to combat the limitations associated with ROS activity. To develop such surfaces/coatings, several surface modification techniques such as layer-by-layer technique, immersion/dip coating, spin coating, plasma immersion ion implantation and radiofrequency plasma-enhanced chemical vapor deposition (enlisted in Table 1) are being researched. A limited number of coating surfaces/materials have been explored to improve the antioxidant activity of Ti surfaces which can be conveniently categorized as organic and inorganic materials for surface modification.

3.1. Organic Materials for Surface Modification

3.1.1. Tannic Acid

Tannic acid is a water-soluble natural polyphenol compound, which is often present in tea, wine and fruits and possesses excellent antioxidant and antibacterial activity owing to the presence of pyrogallol and catechol groups [92]. The antioxidant activity of tannic acid is dependent on its capability to chelate metal ions such as Fe(II) and interfering one of the reaction steps in Fenton reaction, thereby retarding oxidation [93]. There are several published review articles pertaining to the surface modification aspects of tannic acid-based coatings for various applications [94,95]. Several techniques such as layer-by-layer deposition [96,97], electrodeposition, UV-assisted deposition [98] and immersion [53] have been used to deposit tannic acid coatings. The presence of catechol groups renders tannic acid substrate-independent adhesive properties. Polyphenol group interactions can occur via several catechol–surface interactions ranging from noncovalent interactions (hydrogen bonding, pi–pi interactions) to chemical bonding (coordination, covalent) [99]. In addition, polyphenol tannic acid is capable of forming functional coatings on various materials by means of an intrinsic auto-oxidative surface polymerization. Sebastian et al. investigated the deposition kinetics of tannic acid on Ti surfaces which revealed a multiphase layer generation [100]. An initial growth phase revealed build-up of layer which is compact as well as rigid (approx. 2 h), followed by adsorption of an increasingly dissipative layer (approx. 5 h). Following this, a coating discontinuation was observed which was corroborated with large particle precipitation in coating solutions.
In order to develop multifunctional coatings on Ti surface, tannic acid is often co-deposited along with other functional biomaterial coatings for prospective implant applications. Hydroxyapatite (Ca10(PO4)6(OH)2) is a bioactive material, the main inorganic bone component which possesses excellent osteoinduction and osteoconduction properties. In view of rendering Ti surfaces (which are bioinert) with bioactive and antioxidant properties, hydroxyapatite and tannic acid based composite coatings have been explored. A consistent and strong antioxidant activity was displayed by hydroxyapatite/tannic acid coatings deposited on Ti substrates modified by titania (TiO2) nanotubes (Figure 2a–c) [101]. Gelatin added to hydroxyapatite can improve the osteogenic aspects to enhance bone formation. However, gelatin-hydroxyapatite coatings failed in bone conduction function due to weak bonding between them. Tannic acid has been found to strongly adsorb to hydroxyapatite surface and firmly glued gelatin and hydroxyapatite [96]. The resultant tannic acid-hydroxyapatite-gelatin complex coating demonstrated significantly higher antioxidant activity and reduced cell damage/changes in the presence of H2O2. There are limitations reported with the adherence of tannic acid onto hydroxyapatite and salivary acquired pellicle peptide modified tannic acid exhibited better adsorption performance on hydroxyapatite surface [102]. Tightly adsorbed coating exhibited smooth, superhydrophilic surface with excellent antibacterial and antibiofouling performance.
In order to develop multifunctional antioxidant and antibacterial coatings, tannic acid is coated along with antibacterial elements which can be contact killing, release killing or anti-adhesive. Despite being widely explored for a wide spectrum of antibacterial applications, silver (Ag) usage for bio-surfaces is limited by dose dependent cytotoxicity. Hydroxyapatite-tannic acid coating developed by immersion technique on a Ag-loaded TiO2 nanotubular Ti surface demonstrated high antibacterial activity, improved cytocompatibility and revealed slow release of tannic acid from surface, which contributed towards persistent antioxidant activity as shown in Figure 2d–f [53]. Polyethylene glycol is a promising antifouling polymeric interface, an appropriate proton acceptor and can generate hydrogen bonds with tannic acid [103]. Simultaneous deposition of polyethylene glycol resulted in increased coating thickness and improved surface coverage [104]. A novel pH-bacteria triggered antibiotic release mechanism has been developed by layer-by layer deposition of tannic acid with cationic antibiotics such as tobramycin, gentamicin and polymyxin B [105]. Unlike linear polymer molecules which are incapable of retaining antibiotics, tannic acid through its hydrogen bonding and electrostatic interactions interacted well with the antibiotics. The interesting aspect is the non-eluting characteristic of the tannic acid/antibiotic coating which is capable of triggering antibiotic release created by pH reduction induced by bacterial pathogens. Hizal et al. reported an ultrathin tannic acid/gentamicin layer-by-layer film on 3D nano-pillared structures, which exhibited a 10-fold decrease in bacterial attachment due to larger surface area of nanostructured surface and lower bacterial adhesion forces on nanopillar tips [106]. Apart from these, strontium (Sr2+) incorporated tannic acid functionalized on Ti surface revealed enhanced osteoblast differentiation and reduced osteoclast activity [107].

3.1.2. Chitosan

Chitosan is a polycationic natural macromolecule (with a molecular structure of (1,4)-linked 2-amino-2-deoxy-β-d-glucan), which is capable of reacting with many physiologically relevant ROS [108,109,110,111]. Owing to its various beneficial aspects such as improvement of osseointegration and cellular interactions, minimal foreign body response, favorable degradation rate and, most importantly, due to antioxidant and free radical scavenging activities, this partly deacetylated form of chitin is a prospective material for surface modification [112,113]. Chitosan can form functional coatings on Ti surface owing to the existence of amine groups in chitosan polymer chains, which are capable of developing covalent bonds with Ti via silanization [114]. Reasonable mechanisms for antioxidative action of chitosan include presence of intra-molecular hydrogen bonding [115], residual-free amino groups in water-soluble chitosan which may induce metal chelation [116] and the ability of NH2 amino groups to react with hydroxyl groups (OH) to generate stable macromolecule radicals [117].
Lieder et al. studied the effect of degree of deacetylation of chitosan membranes coated on Ti surfaces which resulted in an improved fibronectin adsorption, cellular attachment and proliferation, but with no instigation of spontaneous osteogenic differentiation [114]. Chitosan coating (85–90% deacetylated) on porous Ti surface evidenced excellent antioxidant effect and favored osteoblast activity under diabetic conditions through reactivation of P13K/AKT pathway [118]. The study elucidated that chitosan can play a role in the reactivation of P13K/AKT pathway which mediates diabetes-induced ROS overproduction at bone-implant interface (Figure 3c,d). A multi-step layer-by-layer self-assembly was employed to deposit biofunctional composite coatings of chitosan and alginate enriched with caffeic acid on Ti-6Al-7Nb surface [119].
Multiple steps consisted of piranha solution treatment of Ti alloy surface, plasma chemical activation and dip coating. Antioxidant activity measured in terms of DPPH-scavenging activity was higher for chitosan coating due to its potent reducing activity by hydrogen-donating ability. Conjugation of caffeic acid on chitosan resulted in the generation of amide linkages, increasing the amount of electron-donating groups. Another chitosan-based composite coating consisting of chitosan-catechol, gelatin and hydroxyapatite nanofibres deposited on Ti substrates exhibited high level of ROS scavenging activity and decreased oxidative damage on cellular level as displayed in Figure 3a,b [51]. This coating was able to retain increased level of p-FAK (assists in cell spreading and migration) and p-Akt (control cell survival and apoptosis) compared to pure Ti. The developed multilayer coating improved cell matrix adhesion and intercellular adhesion, while attenuating ROS-induced damage by interfering expressions of integrin αv and β3, cadherin genes, anti-apoptotic and pro-apoptotic gene amounts. Electrophoretic deposition technique is also explored recently to coat chitosan-based composite coatings with hydroxyapatite, graphene and gentamycin [120].

3.1.3. Proanthocyanidin

Proanthocyanidin is condensed tannins (comprising of flavan-3-ol monomeric units), which belongs to the class of naturally occurring polyphenol flavonoid (non-thiol natural antioxidant), is found abundantly in berries and fruits [121,122]. Proanthocyanidin possesses excellent ROS scavenging activity, can regulate macrophage behavior and is capable of stimulating bone formation under oxidative stress conditions [123,124,125,126]. Tang et al. reported layer-by-layer self-assembly method to deposit hyaluronic acid/chitosan multilayers with proanthocyanidins [127]. The three-dimensional multilayered network of hyaluronic acid/chitosan on Ti surface facilitated proanthocyanidin incorporation into the micro interspaces between hyaluronic acid and chitosan, eventually leading to its controlled release. Proanthocyanidin incorporation is based on the electrostatic interaction between reactive OH radical in proanthocyanidin and positive amine groups in chitosan. Layer-by-layer assembly is a self-assembly technique based on the electrostatic attractions (polyanions and polycations) between the assembled components to generate polyelectrolyte multilayers. Layer-by-layer technique involves charging Ti substrates by conjugating polyethylenimine for the purpose of obtaining higher binding forces. A sustained release of proanthocyanidin for a prolonged period of 14 days and mitigation of ROS-mediated inflammatory response were inferred. In other work, layer-by-layer technique was employed to integrate collagen type-II with proanthocyanidin which assisted in accelerating proliferation and osteogenic differentiation via Wnt/b-catenin signaling pathway and improved bone generation in vivo [128]. A novel covalent-conjugation strategy is reported to immobilize chitosan-encapsulated proanthocyanidin on Ti surface based on coupling agents (3-aminopropyl) triethoxysilane and glutaraldehyde [129]. Effective attenuation of the inhibitory effect of oxidative stress was induced by proanthocyanidin by the decrease of p53 gene expression. This study also indicated the improved stability of covalently immobilized coatings with improved wear and compression resistances attributed to strong chemical bonding and possessed the advantage of using nanoparticles as roller bearings.

4. Inorganic Materials for Surface Engineering Antioxidative Properties

4.1. Ceria Based Coatings

Cerium is a rare earth metallic element in lanthanide series and can exist in either free metal or metallic oxide form. Cerium possesses dual oxidation states: trivalent cerium sesquioxide (cerous Ce3+) and tetravalent cerium dioxide (ceric Ce4+) forms. Cerium oxide nanoparticles have received widespread attention for biocompatibility improvement, ophthalmic applications [130], cardiovascular pathology, treating neurodegenerative disorders and spinal cord injury owing to its ROS-scavenging ability [131,132]. The role of cerium oxide nanoparticles to effectuate ROS-scavenging activity and antioxidant mimicking role has been extensively reviewed by Nelson et al. [133]. Cerium oxide nanoparticles exhibit rapid and expedient switches in oxidation state between Ce3+ and Ce4+ during redox reactions. Owing to its lower reduction potential, cerium oxide exhibits redox-cycling property.
Cerium oxide is one of the most interesting oxides due to the presence of oxygen vacancy defects (which can be quickly generated and eliminated), and it can act as an oxygen buffer. The presence of oxygen vacancy sites on nanoceria lattice is responsible for the unusual catalytic activity of this class of material which is dependent on the efficient supply of lattice oxygen at reaction sites governed by the formation of oxygen vacancy sites [134]. Catalytic reaction of cerium oxide nanoparticles with super oxide anion (O2−) and hydrogen peroxide (H2O2) mimics biological action of SOD-mimetic and catalase thereby protecting cells against ROS induced damage [135]. Multi-enzymatic antioxidant activity is based on the ability of cerium oxide to rapidly switch between the multiple valence states. SOD mimic activity is elicited by a shift from Ce3+ to Ce4+ (scavenging of O2−), and catalase mimic activity is induced by a shift from Ce4+ to Ce3+ (deactivating hydrogen peroxide) [135,136,137]. SOD and catalase mimic activity of cerium oxide nanoparticles is particularly relevant under physiological pH condition (pH-7.5), rendering ROS-scavenging properties and inhibiting inflammatory mediator production.
Plasma-sprayed cerium oxide coating with a hierarchical topography was developed for antioxidant surfaces to preserve the intracellular antioxidant defense system [138]. Ceria oxide coating was found to be successful in decreasing SOD activity, reducing ROS generation and suppressing malondialdehyde development in hydrogen peroxide-treated osteoblasts. Li et al. reported magnetron sputtering (2, 3 and 5 min, ≈104 Pa) and vacuum annealing (450 °C) to deposit tiny homogenously distributed cerium oxide nanoparticle coatings with varying surface Ce4+/Ce3+ ratios by tuning of deposition time [131]. Quite interestingly, the Ce4+/Ce3+ ratio in this work reported the opposite trend for SOD and catalase mimetic activity. This work also highlighted the effective antioxidative mechanism of cerium oxide only when SOD and catalase mimetic activities are coordinated (H2O2 decomposition rate ≥ generation rate). The observed Ce4+/Ce3+ ratio resulted in improved cytocompatibility, new bone formation, bone integration and upregulation of osteogenic genes and protein expressions (Figure 4a). Apart from the surface chemistry, the shape of ceria-based nanoparticles has also been reported to influence its ROS scavenging activity. Nanowire-shaped ceria is reported to occupy the extracellular space as its cellular internalization rate is slow [139]. Hence, nanowire-shaped ceria present on the cell surface can level down the hydrogen peroxide molecules and induce ROS consumption as schematically shown in Figure 4b. Spin coating represents a quick and facile surface modification technique to obtain coatings of thickness ranging from a few nanometers to few micrometers. Spin coating technique was used to deposit hydrothermally synthesized nano-CeO2 with varying morphologies (nanorod, nano-cube and nano-octahedra) which yielded uniform coatings in Ti surfaces [140]. Total antioxidant capacity was in the order of nano-octahedron > nano-cube > nanorod. The anti-inflammation ability was correlated to the relative Ce3+ or Ce4+ content (XPS results displayed in Figure 4c) which, in turn, was influenced by the particle size and exposed crystalline lattice planes. With decreased particles size, Ce3+ content increased and rendered nano-octahedron improved SOD mimetic activity.
High energy Ce plasma was used to develop cerium-modified Ti surface by using plasma immersion ion implantation technology [141]. A shift in the appearance of surface from nano-grains to nano-pits was noted, with treatment time increased from 30 to 60 and 90 min. Cerium implantation on Ti surface resulted in reducing the hydroxyl radical generation on Ti surface with increase in plasma immersion time. Reduction in fluorescence intensities and enhanced protection of E.coli model from oxidative stress are attributed to the improved corrosion resistance of the modified surface and the capability of the CeOx on Ti surface to consume hydroxyl radical and hydrogen peroxide. In a similar work, atmospheric plasma was used to deposit cerium oxide-incorporated calcium silicate coating on Ti-6Al-4V substrates [142]. The developed surfaces demonstrated good biochemical stability, cellular viability and antibacterial activity against E. faecalis. Recently metal organic framework (MOF) coating was developed in situ on Ti surface to develop bio-responsive Ce/Sr incorporated bio-MOF coating based on hydrothermal technique [143]. Hydrothermally treated Ce-MOF and Ce/Sr-MOF revealed a Ce4+/Ce3+ ratio of 1.186 and 2.76, respectively. Both of these Ce-containing surfaces revealed excellent H2O2 decomposition, superoxide anion disintegrating capacity, 80% of radical clearance during DPPH assay and persistence of antioxidant activity with TMB assay.

4.2. Silica

Various silicon based coatings have been explored for biocompatibility applications such as amorphous silicon oxygen thin films (a-SiOx) [144,145], calcium silicate [146], sol-gel-based silica bioglasses [147], silicon nitride [148], etc. Silicon is an important element which possesses an influential role in the activity of SOD to improve the ROS scavenging ability. Ilias et al. studied the plasma-enhanced chemical vapor deposition of amorphous silicon oxynitride in view of attaining rapid bone regeneration and fracture healing [149]. This study is the first of its kind to reveal the dependence of Si4+ on SOD1 to improve osteogenesis. For an effective bone healing, a sustained released of Si4+ should be ensured from the implant surface. Nitrogen incorporation into amorphous silica effectuated a continual Si4+ release which can be fine-tuned based on the surface chemistry (O/N ratio), and thickness of deposited film dictates the total release period. Plasma-enhanced chemical vapor deposition was similarly utilized by the same research group to develop coatings in the form of silicon oxynitrophosphide [150]. Up-regulation of SOD1 and cat-1 was observed in cells exposed to silicon oxynitrophosphide with varying oxygen and nitrogen contents. In other work, a radio frequency plasma-enhanced chemical vapor deposition (RF-PECVD) method which makes use of silane as Si source was used to deposit hydrogenated amorphous silicon coatings on Ti-6Al-4V substrate [151]. Hydrogen incorporation into the coating resulted in lower surface oxidation and amorphous silicon coating influenced fibroblasts with no significant effect on keratinocytes. A table enlisting the summary of advantages and limitations of different organic and inorganic coatings/materials described is provided in Table 2.

5. Conclusions and Future Perspectives

There are several titanium-surface modification techniques being used which can be classified as mechanical (polishing, blasting, peening), chemical (chemical treatment, sol-gel, anodic oxidation, chemical vapor deposition) and physical (thermal spraying, plasma spraying, physical vapor deposition, evaporation, ion plating, sputtering, glow discharge plasma, ion implantation and deposition) techniques [57,163]. In spite of the fact that several surface modification strategies are being researched with focus on antibacterial and cytocompatible surfaces, Ti surfaces with improved antioxidant properties require further research focus. The most common techniques explored on depositing functional molecules on the Ti surfaces are based on physical adsorption, based on weak hydrogen bonding and van der Waals forces. This is a limiting factor as it restricts the bond strength and coating life, which will potentially affect the efficacy of the implant. This can be tackled based on chemical immobilization via covalent bonding, in which case a more sustained release of functional molecules can be achieved as compared to physical adsorption techniques.
A critical limitation hampering the potentiality of Ti and its alloys is the inferior wear resistance to be used in articulating surfaces. In spite of the fact that various organic coatings on Ti are bioactive and can develop antioxidant activity, these coatings are mechanically instable, which is a particularly relevant aspect to be considered in terms of wear resistance. During surgical procedures, these implants often encounter mechanical forces of up to 15 N, which will critically affect the life of the coatings and sometimes can lead to coating spalling [129]. One way to tackle this will be the immobilization of such molecules on an already-modified surface layer [163]. Hence, a prospective idea is the development of bi-layer coating consisting of (a) an inner wear/corrosion resistant layer and (b) an outer bioactive layer with antioxidant activity. More research should be focused towards the extraction of exogenous antioxidants (mainly derived from food and medicinal plants, such as fruits, vegetables, cereals, mushrooms, beverages, flowers, spices and traditional medicinal herbs [164]) and its immobilization on Ti surface. Increasing the complexity of a surface modification process will render the process difficult and expensive for rapid commercialization.
One of the critical factors to be assessed while developing such surface is the effect of surface oxide layer on Ti surface. Ti and its alloys, when exposed to air, form a spontaneous native titanium dioxide (TiO2) layer with thickness in the range of 2–20 nm. This possesses a profound influence on the binding of molecules on the Ti surface and coating adhesion. Antioxidant release kinetics should also be given prior focus, as burst release in physiological environment can induce harmful enzymatic imbalances. More computational studies focused towards the stimulatory effect of various prospective coating materials on inducing oxidative stress and ROS need emphasis. Apart from these aspects, various factors to stimulate physiological conditions to assess the antioxidant activity of the developed surfaces shall be incorporated in studies, as synergetic influence of factors can alter the ROS scavenging activity.
Despite the beneficial aspects possessed by Ti and its alloy for load-bearing implant applications, there is a plenty of room for investigating the complex biological phenomena associated with ROS activity in the physiological environment. Most of the published works intended to improve the antioxidant properties of Ti surface are based on organic materials such as tannic acid, chitosan and proanthocyanidin and inorganic coatings based on ceria and silica. Future relevant research trends can be foreseen in improving the mechanical stability and controlled drug elution associated with organic coatings. It is also highly desirable that multifactorial aspects in a real physiological environment shall be considered while assessing the ROS scavenging activity of the developed surfaces. Overall, it is suggested to consider the ROS generation and antioxidant aspects with more research in this direction to develop an efficient implant surface of metallic biomaterials for improving the clinical efficiency. Most importantly, complex challenges associated with translation of lab research to clinical practice demands effective collaboration between material scientists, engineers, biologists and clinicians.

Author Contributions

Conceptualization, J.V. and B.K.-S.; writing—original draft preparation, J.V. and B.K.-S.; writing—review and editing, P.K., K.D., M.S.B. and B.S.; supervision, B.S. and B.K.-S.; funding acquisition, B.K.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

J.V. and B.S. would like to gratefully acknowledge the support and encouragement received from Amrita University, Amritapuri Campus.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohamad, N.-V.; Ima-Nirwana, S.; Chin, K.-Y. Are Oxidative Stress and Inflammation Mediators of Bone Loss due to Estrogen Deficiency? A Review of Current Evidence. Endocr. Metab. Immune Disord.-Drug Targets 2020, 20, 1478–1487. [Google Scholar] [CrossRef] [PubMed]
  2. Consensus development conference: Diagnosis, prophylaxis, and treatment of osteoporosis. Am. J. Med. 1993, 94, 646–650. [CrossRef] [PubMed]
  3. Harvey, N.; Dennison, E.; Cooper, C. Osteoporosis: Impact on health and economics. Nat. Rev. Rheumatol. 2010, 6, 99–105. [Google Scholar] [CrossRef] [PubMed]
  4. Salari, N.; Ghasemi, H.; Mohammadi, L.; Behzadi, M.H.; Rabieenia, E.; Shohaimi, S.; Mohammadi, M. The global prevalence of osteoporosis in the world: A comprehensive systematic review and meta-analysis. J. Orthop. Surg. Res. 2021, 16, 609. [Google Scholar] [CrossRef]
  5. Al Anouti, F.; Taha, Z.; Shamim, S.; Khalaf, K.; Al Kaabi, L.; Alsafar, H. An insight into the paradigms of osteoporosis: From genetics to biomechanics. Bone Rep. 2019, 11, 100216. [Google Scholar] [CrossRef]
  6. Johnell, O.; Kanis, J.A. An estimate of the worldwide prevalence and disability associated with osteoporotic fractures. Osteoporos. Int. 2006, 17, 1726–1733. [Google Scholar] [CrossRef]
  7. Ji, M.X.; Yu, Q. Primary osteoporosis in postmenopausal women. Chronic Dis. Transl. Med. 2015, 1, 9–13. [Google Scholar] [CrossRef] [Green Version]
  8. Loi, F.; Córdova, L.A.; Pajarinen, J.; Lin, T.-H.; Yao, Z.; Goodman, S.B. Inflammation, fracture and bone repair. Bone 2016, 86, 119–130. [Google Scholar] [CrossRef] [Green Version]
  9. Schell, H.; Duda, G.N.; Peters, A.; Tsitsilonis, S.; Johnson, K.A.; Schmidt-Bleek, K. The haematoma and its role in bone healing. J. Exp. Orthop. 2017, 4, 5. [Google Scholar] [CrossRef] [Green Version]
  10. Walters, G.; Pountos, I.; Giannoudis, P.V. The cytokines and micro-environment of fracture haematoma: Current evidence. J. Tissue Eng. Regen. Med. 2018, 12, e1662–e1677. [Google Scholar] [CrossRef]
  11. Raggatt, L.J.; Partridge, N.C. Cellular and Molecular Mechanisms of Bone Remodeling. J. Biol. Chem. 2010, 285, 25103–25108. [Google Scholar] [CrossRef] [Green Version]
  12. Bellido, T. Osteocyte-Driven Bone Remodeling. Calcif. Tissue Int. 2014, 94, 25–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Xiong, J.; O’Brien, C.A. Osteocyte RANKL: New insights into the control of bone remodeling. J. Bone Miner. Res. 2012, 27, 499–505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Marcucci, G.; Domazetovic, V.; Nediani, C.; Ruzzolini, J.; Favre, C.; Brandi, M.L. Oxidative Stress and Natural Antioxidants in Osteoporosis: Novel Preventive and Therapeutic Approaches. Antioxidants 2023, 12, 373. [Google Scholar] [CrossRef] [PubMed]
  15. Bădilă, A.E.; Rădulescu, D.M.; Ilie, A.; Niculescu, A.-G.; Grumezescu, A.M.; Rădulescu, A.R. Bone Regeneration and Oxidative Stress: An Updated Overview. Antioxidants 2022, 11, 318. [Google Scholar] [CrossRef] [PubMed]
  16. León-Reyes, G.; Argoty-Pantoja, A.D.; Becerra-Cervera, A.; López-Montoya, P.; Rivera-Paredez, B.; Velázquez-Cruz, R. Oxidative-Stress-Related Genes in Osteoporosis: A Systematic Review. Antioxidants 2023, 12, 915. [Google Scholar] [CrossRef]
  17. Hallab, N.J.; Jacobs, J.J. 2.5.4—Orthopedic Applications. In Biomaterials Science, 4th ed.; Wagner, W.R., Sakiyama-Elbert, S.E., Zhang, G., Yaszemski, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 1079–1118. [Google Scholar] [CrossRef]
  18. Ratner, B.D.; Zhang, G. 1.1.2—A History of Biomaterials. In Biomaterials Science, 4th ed.; Wagner, W.R., Sakiyama-Elbert, S.E., Zhang, G., Yaszemski, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 21–34. [Google Scholar] [CrossRef]
  19. Wagner, W.R. 1.3.1—The Materials Side of the Biomaterials Relationship. In Biomaterials Science, 4th ed.; Wagner, W.R., Sakiyama-Elbert, S.E., Zhang, G., Yaszemski, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 83–84. [Google Scholar] [CrossRef]
  20. Bai, L.; Gong, C.; Chen, X.; Sun, Y.; Zhang, J.; Cai, L.; Zhu, S.; Xie, S.Q. Additive Manufacturing of Customized Metallic Orthopedic Implants: Materials, Structures, and Surface Modifications. Metals 2019, 9, 1004. [Google Scholar] [CrossRef] [Green Version]
  21. Pilliar, R.M. Metallic Biomaterials. In Biomedical Materials; Narayan, R., Ed.; Springer International Publishing: Cham, Germany, 2021; pp. 1–47. [Google Scholar] [CrossRef]
  22. Prasad, K.; Bazaka, O.; Chua, M.; Rochford, M.; Fedrick, L.; Spoor, J.; Symes, R.; Tieppo, M.; Collins, C.; Cao, A.; et al. Metallic Biomaterials: Current Challenges and Opportunities. Materials 2017, 10, 884. [Google Scholar] [CrossRef]
  23. Chen, Q.; Thouas, G.A. Metallic implant biomaterials. Mater. Sci. Eng. R Rep. 2015, 87, 1–57. [Google Scholar] [CrossRef]
  24. Villanueva, J.; Trino, L.; Thomas, J.; Bijukumar, D.; Royhman, D.; Stack, M.M.; Mathew, M.T. Corrosion, Tribology, and Tribocorrosion Research in Biomedical Implants: Progressive Trend in the Published Literature. J. Bio-Tribo-Corros. 2016, 3, 1. [Google Scholar] [CrossRef] [Green Version]
  25. Eliaz, N. Corrosion of Metallic Biomaterials: A Review. Materials 2019, 12, 407. [Google Scholar] [CrossRef] [Green Version]
  26. Steinbeck, M.J.; Jablonowski, L.J.; Parvizi, J.; Freeman, T.A. The Role of Oxidative Stress in Aseptic Loosening of Total Hip Arthroplasties. J. Arthroplast. 2014, 29, 843–849. [Google Scholar] [CrossRef] [Green Version]
  27. Primožič, J.; Poljšak, B.; Jamnik, P.; Kovač, V.; Čanadi Jurešić, G.; Spalj, S. Risk Assessment of Oxidative Stress Induced by Metal Ions Released from Fixed Orthodontic Appliances during Treatment and Indications for Supportive Antioxidant Therapy: A Narrative Review. Antioxidants 2021, 10, 1359. [Google Scholar] [CrossRef]
  28. Goodman, S.B.; Gallo, J. Periprosthetic Osteolysis: Mechanisms, Prevention and Treatment. J. Clin.Med. 2019, 8, 2091. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, Y.; Branicky, R.; Noë, A.; Hekimi, S. Superoxide dismutases: Dual roles in controlling ROS damage and regulating ROS signaling. J. Cell Biol. 2018, 217, 1915–1928. [Google Scholar] [CrossRef] [Green Version]
  30. Atashi, F.; Modarressi, A.; Pepper, M.S. The role of reactive oxygen species in mesenchymal stem cell adipogenic and osteogenic differentiation: A review. Stem Cells Dev. 2015, 24, 1150–1163. [Google Scholar] [CrossRef] [Green Version]
  31. Cerqueni, G.; Scalzone, A.; Licini, C.; Gentile, P.; Mattioli-Belmonte, M. Insights into oxidative stress in bone tissue and novel challenges for biomaterials. Mater. Sci. Eng. C 2021, 130, 112433. [Google Scholar] [CrossRef] [PubMed]
  32. Domazetovic, V.; Marcucci, G.; Iantomasi, T.; Brandi, M.L.; Vincenzini, M.T. Oxidative stress in bone remodeling: Role of antioxidants. Clin. Cases Miner. Bone Metab. Off. J. Ital. Soc. Osteoporos. Miner. Metab. Skelet. Dis. 2017, 14, 209–216. [Google Scholar] [CrossRef] [PubMed]
  33. Pizzino, G.; Irrera, N.; Cucinotta, M.; Pallio, G.; Mannino, F.; Arcoraci, V.; Squadrito, F.; Altavilla, D.; Bitto, A. Oxidative Stress: Harms and Benefits for Human Health. Oxidative Med. Cell. Longev. 2017, 2017, 8416763. [Google Scholar] [CrossRef] [Green Version]
  34. Sies, H.; Belousov, V.V.; Chandel, N.S.; Davies, M.J.; Jones, D.P.; Mann, G.E.; Murphy, M.P.; Yamamoto, M.; Winterbourn, C. Defining roles of specific reactive oxygen species (ROS) in cell biology and physiology. Nat. Rev. Mol. Cell Biol. 2022, 23, 499–515. [Google Scholar] [CrossRef] [PubMed]
  35. Thannickal, V.J.; Fanburg, B.L. Reactive oxygen species in cell signaling. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2000, 279, L1005–L1028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Brieger, K.; Schiavone, S.; Miller, F.J., Jr.; Krause, K.-H. Reactive oxygen species: From health to disease. Swiss Med. Wkly. 2012, 142, w13659. [Google Scholar] [CrossRef] [PubMed]
  37. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef] [PubMed]
  38. Parham, S.; Kharazi, A.Z.; Bakhsheshi-Rad, H.R.; Nur, H.; Ismail, A.F.; Sharif, S.; RamaKrishna, S.; Berto, F. Antioxidant, Antimicrobial and Antiviral Properties of Herbal Materials. Antioxidants 2020, 9, 1309. [Google Scholar] [CrossRef] [PubMed]
  39. Collin, F. Chemical Basis of Reactive Oxygen Species Reactivity and Involvement in Neurodegenerative Diseases. Int. J. Mol.Sci. 2019, 20, 2407. [Google Scholar] [CrossRef] [Green Version]
  40. Murphy, M.P.; Bayir, H.; Belousov, V.; Chang, C.J.; Davies, K.J.A.; Davies, M.J.; Dick, T.P.; Finkel, T.; Forman, H.J.; Janssen-Heininger, Y.; et al. Guidelines for measuring reactive oxygen species and oxidative damage in cells and in vivo. Nat. Metab. 2022, 4, 651–662. [Google Scholar] [CrossRef]
  41. Andrés, C.M.; Pérez de la Lastra, J.M.; Andrés Juan, C.; Plou, F.J.; Pérez-Lebeña, E. Superoxide Anion Chemistry—Its Role at the Core of the Innate Immunity. Int. J. Mol.Sci. 2023, 24, 1841. [Google Scholar]
  42. Pervaiz, S.; Clement, M.-V. Superoxide anion: Oncogenic reactive oxygen species? Int. J. Biochem. Cell Biol. 2007, 39, 1297–1304. [Google Scholar] [CrossRef]
  43. Kessler, A.; Hedberg, J.; Blomberg, E.; Odnevall, I. Reactive Oxygen Species Formed by Metal and Metal Oxide Nanoparticles in Physiological Media&Mdash;A Review of Reactions of Importance to Nanotoxicity and Proposal for Categorization. Nanomaterials 2022, 12, 1922. [Google Scholar]
  44. Bešlo, D.; Golubić, N.; Rastija, V.; Agić, D.; Karnaš, M.; Šubarić, D.; Lučić, B. Antioxidant Activity, Metabolism, and Bioavailability of Polyphenols in the Diet of Animals. Antioxidants 2023, 12, 1141. [Google Scholar] [CrossRef]
  45. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M.T.D.; Mazur, M.; Telser, J. Free radicals and antioxidants in normal physiological functions and human disease. Int. J. Biochem. Cell Biol. 2007, 39, 44–84. [Google Scholar] [CrossRef] [PubMed]
  46. Kotha, R.R.; Tareq, F.S.; Yildiz, E.; Luthria, D.L. Oxidative Stress and Antioxidants&Mdash; A Critical Review on In Vitro Antioxidant Assays. Antioxidants 2022, 11, 2388. [Google Scholar]
  47. Smeets, R.; Stadlinger, B.; Schwarz, F.; Beck-Broichsitter, B.; Jung, O.; Precht, C.; Kloss, F.; Gröbe, A.; Heiland, M.; Ebker, T. Impact of Dental Implant Surface Modifications on Osseointegration. BioMed Res. Int. 2016, 2016, 6285620. [Google Scholar] [CrossRef] [Green Version]
  48. Zhu, G.; Wang, G.; Li, J.J. Advances in implant surface modifications to improve osseointegration. Mater. Adv. 2021, 2, 6901–6927. [Google Scholar] [CrossRef]
  49. Ghosh, S.; Abanteriba, S. Status of surface modification techniques for artificial hip implants. Sci. Technol. Adv. Mater. 2016, 17, 715–735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Liu, Y.; Rath, B.; Tingart, M.; Eschweiler, J. Role of implants surface modification in osseointegration: A systematic review. J. Biomed. Mater. Res. Part A 2020, 108, 470–484. [Google Scholar] [CrossRef] [Green Version]
  51. Chen, W.; Shen, X.; Hu, Y.; Xu, K.; Ran, Q.; Yu, Y.; Dai, L.; Yuan, Z.; Huang, L.; Shen, T.; et al. Surface functionalization of titanium implants with chitosan-catechol conjugate for suppression of ROS-induced cells damage and improvement of osteogenesis. Biomaterials 2017, 114, 82–96. [Google Scholar] [CrossRef]
  52. Dumitriu, C.; Ungureanu, C.; Popescu, S.; Tofan, V.; Popescu, M.; Pirvu, C. Ti surface modification with a natural antioxidant and antimicrobial agent. Surf. Coat. Technol. 2015, 276, 175–185. [Google Scholar] [CrossRef]
  53. Di, H.; Qiaoxia, L.; Yujie, Z.; Jingxuan, L.; Yan, W.; Yinchun, H.; Xiaojie, L.; Song, C.; Weiyi, C. Ag nanoparticles incorporated tannic acid/nanoapatite composite coating on Ti implant surfaces for enhancement of antibacterial and antioxidant properties. Surf. Coat. Technol. 2020, 399, 126169. [Google Scholar] [CrossRef]
  54. Clanton, T.L. Hypoxia-induced reactive oxygen species formation in skeletal muscle. J. Appl. Physiol. 2007, 102, 2379–2388. [Google Scholar] [CrossRef] [Green Version]
  55. Chapple, I.L.C. Reactive oxygen species and antioxidants in inflammatory diseases. J. Clin. Periodontol. 1997, 24, 287–296. [Google Scholar] [CrossRef]
  56. Han, X.; Ma, J.; Tian, A.; Wang, Y.; Li, Y.; Dong, B.; Tong, X.; Ma, X. Surface modification techniques of titanium and titanium alloys for biomedical orthopaedics applications: A review. Colloids Surf. B Biointerfaces 2023, 227, 113339. [Google Scholar] [CrossRef]
  57. Liu, X.; Chu, P.K.; Ding, C. Surface modification of titanium, titanium alloys, and related materials for biomedical applications. Mater. Sci. Eng. R Rep. 2004, 47, 49–121. [Google Scholar] [CrossRef] [Green Version]
  58. Geetha, M.; Singh, A.K.; Asokamani, R.; Gogia, A.K. Ti based biomaterials, the ultimate choice for orthopaedic implants—A review. Prog. Mater. Sci. 2009, 54, 397–425. [Google Scholar] [CrossRef]
  59. Kaur, M.; Singh, K. Review on titanium and titanium based alloys as biomaterials for orthopaedic applications. Mater. Sci. Eng. C 2019, 102, 844–862. [Google Scholar] [CrossRef] [PubMed]
  60. Quinn, J.; McFadden, R.; Chan, C.-W.; Carson, L. Titanium for Orthopedic Applications: An Overview of Surface Modification to Improve Biocompatibility and Prevent Bacterial Biofilm Formation. iScience 2020, 23, 101745. [Google Scholar] [CrossRef] [PubMed]
  61. Chen, L.-Y.; Cui, Y.-W.; Zhang, L.-C. Recent Development in Beta Titanium Alloys for Biomedical Applications. Metals 2020, 10, 1139. [Google Scholar] [CrossRef]
  62. Kolli, R.P.; Devaraj, A. A Review of Metastable Beta Titanium Alloys. Metals 2018, 8, 506. [Google Scholar] [CrossRef] [Green Version]
  63. Abdel-Hady Gepreel, M.; Niinomi, M. Biocompatibility of Ti-alloys for long-term implantation. J. Mech. Behav. Biomed. Mater. 2013, 20, 407–415. [Google Scholar] [CrossRef]
  64. Biesiekierski, A.; Wang, J.; Abdel-Hady Gepreel, M.; Wen, C. A new look at biomedical Ti-based shape memory alloys. Acta Biomater. 2012, 8, 1661–1669. [Google Scholar] [CrossRef]
  65. Vishnu, J.; Manivasagam, G. Surface Modification and Biological Approaches for Tackling Titanium Wear-Induced Aseptic Loosening. J. Bio-Tribo-Corros. 2021, 7, 32. [Google Scholar] [CrossRef]
  66. Peng, K.T.; Hsu, W.H.; Shih, H.N.; Hsieh, C.W.; Huang, T.W.; Hsu, R.W.W.; Chang, P.J. The role of reactive oxygen species scavenging enzymes in the development of septic loosening after total hip replacement. J. Bone Jt. Surg. Br. Vol. 2011, 93-B, 1201–1209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Prestat, M.; Thierry, D. Corrosion of titanium under simulated inflammation conditions: Clinical context and in vitro investigations. Acta Biomater. 2021, 136, 72–87. [Google Scholar] [CrossRef] [PubMed]
  68. Nguyen, G.T.; Green, E.R.; Mecsas, J. Neutrophils to the ROScue: Mechanisms of NADPH Oxidase Activation and Bacterial Resistance. Front. Cell. Infect. Microbiol. 2017, 7, 373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Groemping, Y.; Rittinger, K. Activation and assembly of the NADPH oxidase: A structural perspective. Biochem. J. 2005, 386, 401–416. [Google Scholar] [CrossRef] [Green Version]
  70. Fujii, J.; Homma, T.; Osaki, T. Superoxide Radicals in the Execution of Cell Death. Antioxidants 2022, 11, 501. [Google Scholar] [CrossRef]
  71. Getzoff, E.D.; Tainer, J.A.; Weiner, P.K.; Kollman, P.A.; Richardson, J.S.; Richardson, D.C. Electrostatic recognition between superoxide and copper, zinc superoxide dismutase. Nature 1983, 306, 287–290. [Google Scholar] [CrossRef]
  72. Paumann-Page, M.; Furtmüller, P.G.; Hofbauer, S.; Paton, L.N.; Obinger, C.; Kettle, A.J. Inactivation of human myeloperoxidase by hydrogen peroxide. Arch. Biochem. Biophys. 2013, 539, 51–62. [Google Scholar] [CrossRef] [Green Version]
  73. Khan, A.A.; Alsahli, M.A.; Rahmani, A.H. Myeloperoxidase as an Active Disease Biomarker: Recent Biochemical and Pathological Perspectives. Med. Sci. 2018, 6, 33. [Google Scholar] [CrossRef] [Green Version]
  74. Prousek, J. Fenton chemistry in biology and medicine. Pure Appl. Chem. 2007, 79, 2325–2338. [Google Scholar] [CrossRef]
  75. Winterbourn, C.C. Toxicity of iron and hydrogen peroxide: The Fenton reaction. Toxicol. Lett. 1995, 82–83, 969–974. [Google Scholar] [CrossRef]
  76. Kehrer, J.P. The Haber—Weiss reaction and mechanisms of toxicity. Toxicology 2000, 149, 43–50. [Google Scholar] [CrossRef] [PubMed]
  77. Peñarrieta-Juanito, G.; Sordi, M.B.; Henriques, B.; Dotto, M.E.R.; Teughels, W.; Silva, F.S.; Magini, R.S.; Souza, J.C.M. Surface damage of dental implant systems and ions release after exposure to fluoride and hydrogen peroxide. J. Periodontal Res. 2019, 54, 46–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kalbacova, M.; Roessler, S.; Hempel, U.; Tsaryk, R.; Peters, K.; Scharnweber, D.; Kirkpatrick, J.C.; Dieter, P. The effect of electrochemically simulated titanium cathodic corrosion products on ROS production and metabolic activity of osteoblasts and monocytes/macrophages. Biomaterials 2007, 28, 3263–3272. [Google Scholar] [CrossRef] [PubMed]
  79. Brown, P.S.; Bhushan, B. Mechanically durable, superomniphobic coatings prepared by layer-by-layer technique for self-cleaning and anti-smudge. J. Colloid Interface Sci. 2015, 456, 210–218. [Google Scholar] [CrossRef]
  80. Ariga, K.; Hill, J.P.; Ji, Q. Layer-by-layer assembly as a versatile bottom-up nanofabrication technique for exploratory research and realistic application. Phys. Chem. Chem. Phys. 2007, 9, 2319–2340. [Google Scholar] [CrossRef]
  81. Neacşu, I.A.; Nicoară, A.I.; Vasile, O.R.; Vasile, B.Ş. Chapter 9—Inorganic micro- and nanostructured implants for tissue engineering. In Nanobiomaterials in Hard Tissue Engineering; Grumezescu, A.M., Ed.; William Andrew Publishing: Norwich, NY, USA, 2016; pp. 271–295. [Google Scholar] [CrossRef]
  82. Alfieri, M.L.; Riccucci, G.; Ferraris, S.; Cochis, A.; Scalia, A.C.; Rimondini, L.; Panzella, L.; Spriano, S.; Napolitano, A. Deposition of Antioxidant and Cytocompatible Caffeic Acid-Based Thin Films onto Ti6Al4V Alloys through Hexamethylenediamine-Mediated Crosslinking. ACS Appl. Mater. Interfaces 2023, 15, 29618–29635. [Google Scholar] [CrossRef]
  83. Rased, N.H.; Vengadaesvaran, B.; Raihan, S.R.S.; Rahim, N.A. Chapter 6—Introduction to solar energy and its conversion into electrical energy by using dye-sensitized solar cells. In Energy Materials; Dhoble, S.J., Kalyani, N.T., Vengadaesvaran, B., Kariem Arof, A., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 139–178. [Google Scholar] [CrossRef]
  84. Sahu, N.; Parija, B.; Panigrahi, S. Fundamental understanding and modeling of spin coating process: A review. Indian J. Phys. 2009, 83, 493–502. [Google Scholar] [CrossRef] [Green Version]
  85. Hashizume, M.; Kunitake, T. Preparation of Self-Supporting Ultrathin Films of Titania by Spin Coating. Langmuir 2003, 19, 10172–10178. [Google Scholar] [CrossRef]
  86. Chen, Y.; Xu, C.; Wang, C.-H.; Bilek, M.M.M.; Cheng, X. An effective method to optimise plasma immersion ion implantation: Sensitivity analysis and design based on low-density polyethylene. Plasma Process. Polym. 2022, 19, 2100199. [Google Scholar] [CrossRef]
  87. Sotoudeh Bagha, P.; Paternoster, C.; Khakbiz, M.; Sheibani, S.; Gholami, N.; Mantovani, D. Surface Modification of an Absorbable Bimodal Fe-Mn-Ag Alloy by Nitrogen Plasma Immersion Ion Implantation. Materials 2023, 16, 1048. [Google Scholar] [CrossRef]
  88. Walschus, U.; Hoene, A.; Patrzyk, M.; Lucke, S.; Finke, B.; Polak, M.; Lukowski, G.; Bader, R.; Zietz, C.; Podbielski, A.; et al. A Cell-Adhesive Plasma Polymerized Allylamine Coating Reduces the In Vivo Inflammatory Response Induced by Ti6Al4V Modified with Plasma Immersion Ion Implantation of Copper. J. Funct. Biomater. 2017, 8, 30. [Google Scholar] [CrossRef] [Green Version]
  89. Terasawa, T.-O.; Saiki, K. Growth of graphene on Cu by plasma enhanced chemical vapor deposition. Carbon 2012, 50, 869–874. [Google Scholar] [CrossRef]
  90. Gan, Z.; Wang, C.; Chen, Z. Material Structure and Mechanical Properties of Silicon Nitride and Silicon Oxynitride Thin Films Deposited by Plasma Enhanced Chemical Vapor Deposition. Surfaces 2018, 1, 59–72. [Google Scholar] [CrossRef] [Green Version]
  91. Vasudev, M.C.; Anderson, K.D.; Bunning, T.J.; Tsukruk, V.V.; Naik, R.R. Exploration of Plasma-Enhanced Chemical Vapor Deposition as a Method for Thin-Film Fabrication with Biological Applications. ACS Appl. Mater. Interfaces 2013, 5, 3983–3994. [Google Scholar] [CrossRef]
  92. Kaczmarek, B. Tannic Acid with Antiviral and Antibacterial Activity as A Promising Component of Biomaterials—A Minireview. Materials 2020, 13, 3224. [Google Scholar] [CrossRef]
  93. Lopes, G.K.B.; Schulman, H.M.; Hermes-Lima, M. Polyphenol tannic acid inhibits hydroxyl radical formation from Fenton reaction by complexing ferrous ions1This study is dedicated to the memory of Botany Professor Luiz F.G. Labouriau (1921–1996).1. Biochim. Et Biophys. Acta BBA-Gen. Subj. 1999, 1472, 142–152. [Google Scholar] [CrossRef]
  94. Sathishkumar, G.; Gopinath, K.; Zhang, K.; Kang, E.-T.; Xu, L.; Yu, Y. Recent progress in tannic acid-driven antibacterial/antifouling surface coating strategies. J. Mater. Chem. B 2022, 10, 2296–2315. [Google Scholar] [CrossRef]
  95. Wang, Z.; Gao, J.; Zhu, L.; Meng, J.; He, F. Tannic acid-based functional coating: Surface engineering of membranes for oil-in-water emulsion separation. Chem. Commun. 2022, 58, 12629–12641. [Google Scholar] [CrossRef]
  96. Zhu, Y.; Zhou, D.; Zan, X.; Ye, Q.; Sheng, S. Engineering the surfaces of orthopedic implants with osteogenesis and antioxidants to enhance bone formation in vitro and in vivo. Colloids Surf. B Biointerfaces 2022, 212, 112319. [Google Scholar] [CrossRef]
  97. Iqbal, M.H.; Schroder, A.; Kerdjoudj, H.; Njel, C.; Senger, B.; Ball, V.; Meyer, F.; Boulmedais, F. Effect of the Buffer on the Buildup and Stability of Tannic Acid/Collagen Multilayer Films Applied as Antibacterial Coatings. ACS Appl. Mater. Interfaces 2020, 12, 22601–22612. [Google Scholar] [CrossRef]
  98. He, X.; Gopinath, K.; Sathishkumar, G.; Guo, L.; Zhang, K.; Lu, Z.; Li, C.; Kang, E.-T.; Xu, L. UV-Assisted Deposition of Antibacterial Ag–Tannic Acid Nanocomposite Coating. ACS Appl. Mater. Interfaces 2021, 13, 20708–20717. [Google Scholar] [CrossRef]
  99. Saiz-Poseu, J.; Mancebo-Aracil, J.; Nador, F.; Busqué, F.; Ruiz-Molina, D. The Chemistry behind Catechol-Based Adhesion. Angew. Chem. Int. Ed. 2019, 58, 696–714. [Google Scholar] [CrossRef] [PubMed]
  100. Geißler, S.; Barrantes, A.; Tengvall, P.; Messersmith, P.B.; Tiainen, H. Deposition Kinetics of Bioinspired Phenolic Coatings on Titanium Surfaces. Langmuir 2016, 32, 8050–8060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Qiaoxia, L.; Yujie, Z.; Meng, Y.; Yizhu, C.; Yan, W.; Yinchun, H.; Xiaojie, L.; Weiyi, C.; Di, H. Hydroxyapatite/tannic acid composite coating formation based on Ti modified by TiO2 nanotubes. Colloids Surf. B Biointerfaces 2020, 196, 111304. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, X.; Huang, P.; Wang, H.; Cai, S.; Liao, Y.; Mo, Z.; Xu, X.; Ding, C.; Zhao, C.; Li, J. Antibacterial and anti-biofouling coating on hydroxyapatite surface based on peptide-modified tannic acid. Colloids Surf. B Biointerfaces 2017, 160, 136–143. [Google Scholar] [CrossRef] [PubMed]
  103. Camós Noguer, A.; Olsen, S.M.; Hvilsted, S.; Kiil, S. Long-term stability of PEG-based antifouling surfaces in seawater. J. Coat. Technol. Res. 2016, 13, 567–575. [Google Scholar] [CrossRef] [Green Version]
  104. Guo, L.L.; Cheng, Y.F.; Ren, X.; Gopinath, K.; Lu, Z.S.; Li, C.M.; Xu, L.Q. Simultaneous deposition of tannic acid and poly(ethylene glycol) to construct the antifouling polymeric coating on Titanium surface. Colloids Surf. B Biointerfaces 2021, 200, 111592. [Google Scholar] [CrossRef]
  105. Zhuk, I.; Jariwala, F.; Attygalle, A.B.; Wu, Y.; Libera, M.R.; Sukhishvili, S.A. Self-Defensive Layer-by-Layer Films with Bacteria-Triggered Antibiotic Release. ACS Nano 2014, 8, 7733–7745. [Google Scholar] [CrossRef]
  106. Hizal, F.; Zhuk, I.; Sukhishvili, S.; Busscher, H.J.; van der Mei, H.C.; Choi, C.-H. Impact of 3D Hierarchical Nanostructures on the Antibacterial Efficacy of a Bacteria-Triggered Self-Defensive Antibiotic Coating. ACS Appl. Mater. Interfaces 2015, 7, 20304–20313. [Google Scholar] [CrossRef] [PubMed]
  107. Steffi, C.; Shi, Z.; Kong, C.H.; Chong, S.W.; Wang, D.; Wang, W. Use of Polyphenol Tannic Acid to Functionalize Titanium with Strontium for Enhancement of Osteoblast Differentiation and Reduction of Osteoclast Activity. Polymers 2019, 11, 1256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Lin, W.; Qi, X.; Guo, W.; Liang, D.; Chen, H.; Lin, B.; Deng, X. A barrier against reactive oxygen species: Chitosan/acellular dermal matrix scaffold enhances stem cell retention and improves cutaneous wound healing. Stem Cell Res. Ther. 2020, 11, 383. [Google Scholar] [CrossRef]
  109. Khokon, M.A.R.; Uraji, M.; Munemasa, S.; Okuma, E.; Nakamura, Y.; Mori, I.C.; Murata, Y. Chitosan-Induced Stomatal Closure Accompanied by Peroxidase-Mediated Reactive Oxygen Species Production in Arabidopsis. Biosci. Biotechnol. Biochem. 2010, 74, 2313–2315. [Google Scholar] [CrossRef] [Green Version]
  110. Banerjee, M.; Mallick, S.; Paul, A.; Chattopadhyay, A.; Ghosh, S.S. Heightened Reactive Oxygen Species Generation in the Antimicrobial Activity of a Three Component Iodinated Chitosan−Silver Nanoparticle Composite. Langmuir 2010, 26, 5901–5908. [Google Scholar] [CrossRef] [PubMed]
  111. Parham, S.; Kharazi, A.Z.; Bakhsheshi-Rad, H.R.; Kharaziha, M.; Ismail, A.F.; Sharif, S.; Razzaghi, M.; RamaKrishna, S.; Berto, F. Antimicrobial Synthetic and Natural Polymeric Nanofibers as Wound Dressing: A Review. Adv. Eng. Mater. 2022, 24, 2101460. [Google Scholar] [CrossRef]
  112. Kumari, S.; Tiyyagura, H.R.; Pottathara, Y.B.; Sadasivuni, K.K.; Ponnamma, D.; Douglas, T.E.L.; Skirtach, A.G.; Mohan, M.K. Surface functionalization of chitosan as a coating material for orthopaedic applications: A comprehensive review. Carbohydr. Polym. 2021, 255, 117487. [Google Scholar] [CrossRef]
  113. Abinaya, B.; Prasith, T.P.; Ashwin, B.; Viji Chandran, S.; Selvamurugan, N. Chitosan in Surface Modification for Bone Tissue Engineering Applications. Biotechnol. J. 2019, 14, 1900171. [Google Scholar] [CrossRef]
  114. Lieder, R.; Darai, M.; Thor, M.B.; Ng, C.H.; Einarsson, J.M.; Gudmundsson, S.; Helgason, B.; Gaware, V.S.; Másson, M.; Gíslason, J.; et al. In vitro bioactivity of different degree of deacetylation chitosan, a potential coating material for titanium implants. J. Biomed. Mater. Res. Part A 2012, 100A, 3392–3399. [Google Scholar] [CrossRef]
  115. Tomida, H.; Fujii, T.; Furutani, N.; Michihara, A.; Yasufuku, T.; Akasaki, K.; Maruyama, T.; Otagiri, M.; Gebicki, J.M.; Anraku, M. Antioxidant properties of some different molecular weight chitosans. Carbohydr. Res. 2009, 344, 1690–1696. [Google Scholar] [CrossRef]
  116. Xue, C.; Yu, G.; Hirata, T.; Terao, J.; Lin, H. Antioxidative Activities of Several Marine Polysaccharides Evaluated in a Phosphatidylcholine-liposomal Suspension and Organic Solvents. Biosci. Biotechnol. Biochem. 1998, 62, 206–209. [Google Scholar] [CrossRef] [Green Version]
  117. Xie, W.; Xu, P.; Liu, Q. Antioxidant activity of water-soluble chitosan derivatives. Bioorganic Med. Chem. Lett. 2001, 11, 1699–1701. [Google Scholar] [CrossRef] [PubMed]
  118. Li, X.; Ma, X.-Y.; Feng, Y.-F.; Ma, Z.-S.; Wang, J.; Ma, T.-C.; Qi, W.; Lei, W.; Wang, L. Osseointegration of chitosan coated porous titanium alloy implant by reactive oxygen species-mediated activation of the PI3K/AKT pathway under diabetic conditions. Biomaterials 2015, 36, 44–54. [Google Scholar] [CrossRef] [PubMed]
  119. Jabłoński, P.; Kyzioł, A.; Pawcenis, D.; Pucelik, B.; Hebda, M.; Migdalska, M.; Krawiec, H.; Arruebo, M.; Kyzioł, K. Electrostatic self-assembly approach in the deposition of bio-functional chitosan-based layers enriched with caffeic acid on Ti-6Al-7Nb alloys by alternate immersion. Biomater. Adv. 2022, 136, 212791. [Google Scholar] [CrossRef] [PubMed]
  120. Stevanović, M.; Djošić, M.; Janković, A.; Kojić, V.; Stojanović, J.; Grujić, S.; Bujagić, I.M.; Rhee, K.Y.; Mišković-Stanković, V. The chitosan-based bioactive composite coating on titanium. J. Mater. Res. Technol. 2021, 15, 4461–4474. [Google Scholar] [CrossRef]
  121. Rauf, A.; Imran, M.; Abu-Izneid, T.; Iahtisham Ul, H.; Patel, S.; Pan, X.; Naz, S.; Sanches Silva, A.; Saeed, F.; Rasul Suleria, H.A. Proanthocyanidins: A comprehensive review. Biomed. Pharmacother. 2019, 116, 108999. [Google Scholar] [CrossRef]
  122. de la Iglesia, R.; Milagro, F.I.; Campión, J.; Boqué, N.; Martínez, J.A. Healthy properties of proanthocyanidins. BioFactors 2010, 36, 159–168. [Google Scholar] [CrossRef] [PubMed]
  123. Park, Y.S.; Jeon, M.H.; Hwang, H.J.; Park, M.R.; Lee, S.H.; Kim, S.G.; Kim, M. Antioxidant activity and analysis of proanthocyanidins from pine (Pinus densiflora) needles. Nutr. Res. Pract. 2011, 5, 281–287. [Google Scholar] [CrossRef] [Green Version]
  124. Yang, L.; Xian, D.; Xiong, X.; Lai, R.; Song, J.; Zhong, J. Proanthocyanidins against Oxidative Stress: From Molecular Mechanisms to Clinical Applications. BioMed Res. Int. 2018, 2018, 8584136. [Google Scholar] [CrossRef] [Green Version]
  125. Andersen-Civil, A.I.S.; Leppä, M.M.; Thamsborg, S.M.; Salminen, J.-P.; Williams, A.R. Structure-function analysis of purified proanthocyanidins reveals a role for polymer size in suppressing inflammatory responses. Commun. Biol. 2021, 4, 896. [Google Scholar] [CrossRef]
  126. Tenkumo, T.; Aobulikasimu, A.; Asou, Y.; Shirato, M.; Shishido, S.; Kanno, T.; Niwano, Y.; Sasaki, K.; Nakamura, K. Proanthocyanidin-rich grape seed extract improves bone loss, bone healing, and implant osseointegration in ovariectomized animals. Sci. Rep. 2020, 10, 8812. [Google Scholar] [CrossRef]
  127. Tang, J.; Chen, L.; Yan, D.; Shen, Z.; Wang, B.; Weng, S.; Wu, Z.; Xie, Z.; Shao, J.; Yang, L.; et al. Surface Functionalization with Proanthocyanidins Provides an Anti-Oxidant Defense Mechanism That Improves the Long-Term Stability and Osteogenesis of Titanium Implants. Int. J. Nanomed. 2020, 15, 1643–1659. [Google Scholar] [CrossRef] [Green Version]
  128. Bai, Z.; Hu, K.; Shou, Z.; Yu, J.; Meng, H.; Zhou, H.; Chen, L.; Yu, T.; Lu, R.; Li, N.; et al. Layer-by-layer assembly of procyanidin and collagen promotes mesenchymal stem cell proliferation and osteogenic differentiation in vitro and in vivo. Regen. Biomater. 2023, 10, rbac107. [Google Scholar] [CrossRef] [PubMed]
  129. Zhou, Q.; Wu, T.; Bai, Z.; Hong, G.; Bian, J.; Xie, H.; Chen, C. A silane-based coupling strategy for enhancing the mechanical properties of proanthocyanidin nanocoatings on Ti dental implants. Appl. Surf. Sci. 2022, 602, 154400. [Google Scholar] [CrossRef]
  130. Maccarone, R.; Tisi, A.; Passacantando, M.; Ciancaglini, M. Ophthalmic Applications of Cerium Oxide Nanoparticles. J. Ocul. Pharmacol. Ther. 2019, 36, 376–383. [Google Scholar] [CrossRef]
  131. Li, J.; Wen, J.; Li, B.; Li, W.; Qiao, W.; Shen, J.; Jin, W.; Jiang, X.; Yeung, K.W.K.; Chu, P.K. Valence State Manipulation of Cerium Oxide Nanoparticles on a Titanium Surface for Modulating Cell Fate and Bone Formation. Adv. Sci. 2018, 5, 1700678. [Google Scholar] [CrossRef] [PubMed]
  132. Dhall, A.; Self, W. Cerium Oxide Nanoparticles: A Brief Review of Their Synthesis Methods and Biomedical Applications. Antioxidants 2018, 7, 97. [Google Scholar] [CrossRef] [Green Version]
  133. Nelson, B.C.; Johnson, M.E.; Walker, M.L.; Riley, K.R.; Sims, C.M. Antioxidant Cerium Oxide Nanoparticles in Biology and Medicine. Antioxidants 2016, 5, 15. [Google Scholar] [CrossRef] [Green Version]
  134. Esch, F.; Fabris, S.; Zhou, L.; Montini, T.; Africh, C.; Fornasiero, P.; Comelli, G.; Rosei, R. Electron Localization Determines Defect Formation on Ceria Substrates. Science 2005, 309, 752–755. [Google Scholar] [CrossRef]
  135. Heckert, E.G.; Karakoti, A.S.; Seal, S.; Self, W.T. The role of cerium redox state in the SOD mimetic activity of nanoceria. Biomaterials 2008, 29, 2705–2709. [Google Scholar] [CrossRef] [Green Version]
  136. Korsvik, C.; Patil, S.; Seal, S.; Self, W.T. Superoxide dismutase mimetic properties exhibited by vacancy engineered ceria nanoparticles. Chem. Commun. 2007, 1056–1058. [Google Scholar] [CrossRef]
  137. Pirmohamed, T.; Dowding, J.M.; Singh, S.; Wasserman, B.; Heckert, E.; Karakoti, A.S.; King, J.E.S.; Seal, S.; Self, W.T. Nanoceria exhibit redox state-dependent catalase mimetic activity. Chem. Commun. 2010, 46, 2736–2738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Li, K.; Xie, Y.; You, M.; Huang, L.; Zheng, X. Plasma sprayed cerium oxide coating inhibits H2O2-induced oxidative stress and supports cell viability. J. Mater. Sci. Mater. Med. 2016, 27, 100. [Google Scholar] [CrossRef] [PubMed]
  139. Mahapatra, C.; Singh, R.K.; Lee, J.-H.; Jung, J.; Hyun, J.K.; Kim, H.-W. Nano-shape varied cerium oxide nanomaterials rescue human dental stem cells from oxidative insult through intracellular or extracellular actions. Acta Biomater. 2017, 50, 142–153. [Google Scholar] [CrossRef] [PubMed]
  140. Li, X.; Qi, M.; Sun, X.; Weir, M.D.; Tay, F.R.; Oates, T.W.; Dong, B.; Zhou, Y.; Wang, L.; Xu, H.H.K. Surface treatments on titanium implants via nanostructured ceria for antibacterial and anti-inflammatory capabilities. Acta Biomater. 2019, 94, 627–643. [Google Scholar] [CrossRef] [PubMed]
  141. Zhang, H.; Qiu, J.; Liu, X. Enhanced antioxidant capability and osteogenic property of medical titanium by cerium plasma immersion ion implantation. Surf. Interfaces 2021, 26, 101402. [Google Scholar] [CrossRef]
  142. Qi, S.; Wu, J.; Xu, Y.; Zhang, Y.; Wang, R.; Li, K.; Xu, Y. Chemical Stability and Antimicrobial Activity of Plasma-Sprayed Cerium Oxide–Incorporated Calcium Silicate Coating in Dental Implants. Implant Dent. 2019, 28, 564–570. [Google Scholar] [CrossRef] [PubMed]
  143. Chen, M.; Wang, D.; Li, M.; He, Y.; He, T.; Chen, M.; Hu, Y.; Luo, Z.; Cai, K. Nanocatalytic Biofunctional MOF Coating on Titanium Implants Promotes Osteoporotic Bone Regeneration through Cooperative Pro-osteoblastogenesis MSC Reprogramming. ACS Nano 2022, 16, 15397–15412. [Google Scholar] [CrossRef]
  144. Mandracci, P.; Mussano, F.; Ceruti, P.; Pirri, C.F.; Carossa, S. Reduction of bacterial adhesion on dental composite resins by silicon–oxygen thin film coatings. Biomed. Mater. 2015, 10, 015017. [Google Scholar] [CrossRef]
  145. Mandracci, P.; Ceruti, P.; Ricciardi, C.; Mussano, F.; Carossa, S. a-SiOx Coatings Grown on Dental Materials by PECVD: Compositional Analysis and Preliminary Investigation of Biocompatibility Improvements. Chem. Vap. Depos. 2010, 16, 29–34. [Google Scholar] [CrossRef]
  146. Alves Silva, E.C.; Tanomaru-Filho, M.; da Silva, G.F.; Delfino, M.M.; Cerri, P.S.; Guerreiro-Tanomaru, J.M. Biocompatibility and Bioactive Potential of New Calcium Silicate–based Endodontic Sealers: Bio-C Sealer and Sealer Plus BC. J. Endod. 2020, 46, 1470–1477. [Google Scholar] [CrossRef]
  147. Arcos, D.; Vallet-Regí, M. Sol–gel silica-based biomaterials and bone tissue regeneration. Acta Biomater. 2010, 6, 2874–2888. [Google Scholar] [CrossRef]
  148. Heimann, R.B. Silicon Nitride, a Close to Ideal Ceramic Material for Medical Application. Ceramics 2021, 4, 208–223. [Google Scholar] [CrossRef]
  149. Ilyas, A.; Odatsu, T.; Shah, A.; Monte, F.; Kim, H.K.W.; Kramer, P.; Aswath, P.B.; Varanasi, V.G. Amorphous Silica: A New Antioxidant Role for Rapid Critical-Sized Bone Defect Healing. Adv. Healthc. Mater. 2016, 5, 2199–2213. [Google Scholar] [CrossRef]
  150. Monte, F.A.D.; Awad, K.R.; Ahuja, N.; Kim, H.K.W.; Aswath, P.; Brotto, M.; Varanasi, V.G. Amorphous Silicon Oxynitrophosphide-Coated Implants Boost Angiogenic Activity of Endothelial Cells. Tissue Eng. Part A 2019, 26, 15–27. [Google Scholar] [CrossRef] [PubMed]
  151. Mussano, F.; Genova, T.; Laurenti, M.; Munaron, L.; Pirri, C.F.; Rivolo, P.; Carossa, S.; Mandracci, P. Hydrogenated amorphous silicon coatings may modulate gingival cell response. Appl. Surf. Sci. 2018, 436, 603–612. [Google Scholar] [CrossRef] [Green Version]
  152. Bigham, A.; Rahimkhoei, V.; Abasian, P.; Delfi, M.; Naderi, J.; Ghomi, M.; Dabbagh Moghaddam, F.; Waqar, T.; Nuri Ertas, Y.; Sharifi, S.; et al. Advances in tannic acid-incorporated biomaterials: Infection treatment, regenerative medicine, cancer therapy, and biosensing. Chem. Eng. J. 2022, 432, 134146. [Google Scholar] [CrossRef]
  153. Kaczmarek-Szczepańska, B.; Polkowska, I.; Paździor-Czapula, K.; Nowicka, B.; Gierszewska, M.; Michalska-Sionkowska, M.; Otrocka-Domagała, I. Chitosan/Phenolic Compounds Scaffolds for Connective Tissue Regeneration. J. Funct. Biomater. 2023, 14, 69. [Google Scholar] [CrossRef] [PubMed]
  154. Widsten, P.; Salo, S.; Niemelä, K.; Helin, H.; Salonen, M.; Alakomi, H.-L. Tannin-Based Microbicidal Coatings for Hospital Privacy Curtains. J. Funct. Biomater. 2023, 14, 187. [Google Scholar] [CrossRef]
  155. Amarowicz, R. Tannins: The new natural antioxidants? Eur. J. Lipid Sci. Technol. 2007, 109, 549–551. [Google Scholar] [CrossRef]
  156. Sanguedolce, M.; Saffioti, M.R.; Rotella, G.; Curcio, F.; Cassano, R.; Umbrello, D.; Filice, L. The Effects of Substrate Material on Chitosan Coating Performance for Biomedical Application. Procedia CIRP 2022, 108, 817–820. [Google Scholar] [CrossRef]
  157. Ganesh, S.S.; Anushikaa, R.; Swetha Victoria, V.S.; Lavanya, K.; Shanmugavadivu, A.; Selvamurugan, N. Recent Advancements in Electrospun Chitin and Chitosan Nanofibers for Bone Tissue Engineering Applications. J. Funct. Biomater. 2023, 14, 288. [Google Scholar] [CrossRef]
  158. Oe, T.; Dechojarassri, D.; Kakinoki, S.; Kawasaki, H.; Furuike, T.; Tamura, H. Microwave-Assisted Incorporation of AgNP into Chitosan–Alginate Hydrogels for Antimicrobial Applications. J. Funct. Biomater. 2023, 14, 199. [Google Scholar] [PubMed]
  159. Yu, K.; Song, Y.; Lin, J.; Dixon, R.A. The complexities of proanthocyanidin biosynthesis and its regulation in plants. Plant Commun. 2023, 4, 100498. [Google Scholar] [CrossRef] [PubMed]
  160. Banavar, S.; Deshpande, A.; Sur, S.; Andreescu, S. Ceria nanoparticle theranostics: Harnessing antioxidant properties in biomedicine and beyond. J. Phys. Mater. 2021, 4, 042003. [Google Scholar] [CrossRef]
  161. Yokel, R.A.; Hussain, S.; Garantziotis, S.; Demokritou, P.; Castranova, V.; Cassee, F.R. The yin: An adverse health perspective of nanoceria: Uptake, distribution, accumulation, and mechanisms of its toxicity. Environ. Sci. Nano 2014, 1, 406–428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Huang, Y.; Li, P.; Zhao, R.; Zhao, L.; Liu, J.; Peng, S.; Fu, X.; Wang, X.; Luo, R.; Wang, R.; et al. Silica nanoparticles: Biomedical applications and toxicity. Biomed. Pharmacother. 2022, 151, 113053. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, L.-C.; Chen, L.-Y.; Wang, L. Surface Modification of Titanium and Titanium Alloys: Technologies, Developments, and Future Interests. Adv. Eng. Mater. 2020, 22, 1901258. [Google Scholar] [CrossRef]
  164. Xu, D.-P.; Li, Y.; Meng, X.; Zhou, T.; Zhou, Y.; Zheng, J.; Zhang, J.-J.; Li, H.-B. Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources. Int. J. Mol. Sci. 2017, 18, 96. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic depicting the ROS generation associated with Ti implants with associated biochemical reactions.
Figure 1. Schematic depicting the ROS generation associated with Ti implants with associated biochemical reactions.
Jfb 14 00344 g001
Figure 2. Antioxidant activity of HAP/tannic acid composite coating. (a) Mechanism illustrating antioxidant activity; (b,c) antioxidant activity as a function of incubation time of HAP/tannic acid coating compared to Trolox; (c) antioxidant activity of (I) negative control (deionized water), (II–IV) 10, 30 and 50 g/L of tannic acid, (V) positive control (Trolox). Data reported as means ± standard deviations, n = 3 (* p < 0.05, ** p < 0.01, *** p < 0.001). Reprinted from [101] with permission from Elsevier. (d) Formation mechanism depiction of Ag nanoparticles; (e) antioxidant activity with respect to incubation time of Ag/HAP/tannic acid coating compared to Trolox; and (f) antioxidant activity of (I) negative control (deionized water), (II–IV) 0.05, 0.1 and 0.2 M AgNO3, (V) positive control (Trolox). Reprinted from [53] with permission from Elsevier.
Figure 2. Antioxidant activity of HAP/tannic acid composite coating. (a) Mechanism illustrating antioxidant activity; (b,c) antioxidant activity as a function of incubation time of HAP/tannic acid coating compared to Trolox; (c) antioxidant activity of (I) negative control (deionized water), (II–IV) 10, 30 and 50 g/L of tannic acid, (V) positive control (Trolox). Data reported as means ± standard deviations, n = 3 (* p < 0.05, ** p < 0.01, *** p < 0.001). Reprinted from [101] with permission from Elsevier. (d) Formation mechanism depiction of Ag nanoparticles; (e) antioxidant activity with respect to incubation time of Ag/HAP/tannic acid coating compared to Trolox; and (f) antioxidant activity of (I) negative control (deionized water), (II–IV) 0.05, 0.1 and 0.2 M AgNO3, (V) positive control (Trolox). Reprinted from [53] with permission from Elsevier.
Jfb 14 00344 g002
Figure 3. Antioxidant activity of chitosan based surfaces on Ti. (a) Antioxidant activity measured over a period of 8 weeks; (b) osteoblast interaction with and without treatment of hydrogen peroxide. Reprinted from [51] with permission from Elsevier. (c) intracellular ROS activity measured by DCF fluorescence intensity and (d) intracellular GSH-Px activity to quantify the rate of oxidation of the reduced glutathione to the oxidized glutathione by H2O2 (* p < 0.05 vs. TI + NS; # p < 0.05 vs. TI + DS; ** p < 0.01). Reprinted from [118] with permission from Elsevier.
Figure 3. Antioxidant activity of chitosan based surfaces on Ti. (a) Antioxidant activity measured over a period of 8 weeks; (b) osteoblast interaction with and without treatment of hydrogen peroxide. Reprinted from [51] with permission from Elsevier. (c) intracellular ROS activity measured by DCF fluorescence intensity and (d) intracellular GSH-Px activity to quantify the rate of oxidation of the reduced glutathione to the oxidized glutathione by H2O2 (* p < 0.05 vs. TI + NS; # p < 0.05 vs. TI + DS; ** p < 0.01). Reprinted from [118] with permission from Elsevier.
Jfb 14 00344 g003
Figure 4. Influence of ceria surface chemistry and shape on the ROS scavenging activity. (a) Influence of Ce4+/Ce3+ ratio on MSC and macrophage. Reprinted from [131] with permission from John Wiley & Sons. (b) X-ray-photoelectron spectroscopy analysis of nanorod, nano-cube and nano-octahedron-shaped ceria with varying Ce4+/Ce3+ ratio. Reprinted from [139] with permission from Elsevier. (c) Schematic depicting the scavenging of extracellular matrix ROS by nanowire-shaped ceria present at the cell surface. Reprinted from [140] with permission from Elsevier.
Figure 4. Influence of ceria surface chemistry and shape on the ROS scavenging activity. (a) Influence of Ce4+/Ce3+ ratio on MSC and macrophage. Reprinted from [131] with permission from John Wiley & Sons. (b) X-ray-photoelectron spectroscopy analysis of nanorod, nano-cube and nano-octahedron-shaped ceria with varying Ce4+/Ce3+ ratio. Reprinted from [139] with permission from Elsevier. (c) Schematic depicting the scavenging of extracellular matrix ROS by nanowire-shaped ceria present at the cell surface. Reprinted from [140] with permission from Elsevier.
Jfb 14 00344 g004
Table 1. Various reported techniques used for developing antioxidant surfaces on Ti/Ti alloy surfaces.
Table 1. Various reported techniques used for developing antioxidant surfaces on Ti/Ti alloy surfaces.
TechniqueDescriptionRef.
Layer-by-layer techniqueBottom-up adsorption technique which involves the development of multi-layered (layers of oppositely charged species) thin films bound together through electrostatic interactions. [79,80]
Immersion/dip coatingSolution-based deposition method which involves the immersion of substrate into a solution of material to be coated which depends on parameters such as dwelling time, substrate-withdrawal speed, number of dip-coating cycles and coating evaporation factor.[81,82]
Spin coatingA technique which uses centrifugal force for deposition, in which a suspension is dropped from top into the rotating substrate, and the resulting centrifugal force will assist in spreading out the coating on the substrate, thereby coating it. The process is dependent on parameters such as dispense volume, spin speed, solution viscosity, solution concentration, spin time, etc.[83,84,85]
Plasma immersion ion implantation (PIII)Method to improve biocompatibility aspects of material surfaces by immersing in a plasma environment and applying negative-high-voltage pulsed bias. Compared to traditional plasma techniques, PIII can extend to some tens of nanometers beneath sample surface and can treat complex geometries.[86,87,88]
Plasma enhanced chemical vapor depositionLow temperature chemical vapor deposition in which plasma is used to drive chemical reactions between plasma-generated-reactive species and substrate instead of high temperatures.[89,90,91]
Table 2. Summary of various coatings used for antioxidant properties.
Table 2. Summary of various coatings used for antioxidant properties.
Type of Coating/SurfaceAntioxidant MechanismAdvantagesLimitationsRef.
Tannic acidAbility to chelate metal ions such as Fe(II), thereby interfering with one of the reaction steps in the Fenton reaction and thereby slowing oxidationAntibacterial, antioxidant, high hemostatic efficiency, anticancer property, regenerative potentialWeak lipid solubility, low bioavailability, and short half-life, release rate should be controlled to exclude potential cytotoxicity, unstable adhesion[152,153,154,155]
ChitosanReasonable mechanisms include presence of intra-molecular hydrogen bonding, metal chelation, ability of NH2 amino groups to react with hydroxyl groups Biological activity, antimicrobial activity, hydrophilicity, and biodegradabilityDelamination, unstable adhesion[156,157,158]
ProanthocyanidinBy scavenging free radicals and by modifying signaling pathways, including those involving nuclear factor erythroid 2-related factor 2 (Nrf2), mitogen-activated protein kinase (MAPK), nuclear factor-kappaB (NF-κB), and phosphoinositide 3-kinase (PI3K)/AktAntioxidant, anticancer, antidiabetic, neuroprotective, and antimicrobialHigh cost, low chemical stability and limited binding sites, difficulties in resolving the chemical labeling pattern of PAs with their proposed biosynthetic pathway, and defining the subcellular sites of biosynthesis[121,124,159]
CeriaAbility to rapidly switch between multiple valence states. SOD mimic activity is elicited by a shift from Ce3+ to Ce4+ (scavenging of O2−) and catalase mimic activity is induced by a shift from Ce4+ to Ce3+ (deactivating hydrogen peroxide)Antioxidant, anticancer and anti-inflammatory properties, biosensorsToxicity associated with small-sized nano-ceria[160,161]
SilicaHydroxylation degree, By regulation of antioxidants enzyme activityAccelerated bone fracture healing, biomineral synthesisLipid peroxidation induced toxicity[149,162]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vishnu, J.; Kesavan, P.; Shankar, B.; Dembińska, K.; Swiontek Brzezinska, M.; Kaczmarek-Szczepańska, B. Engineering Antioxidant Surfaces for Titanium-Based Metallic Biomaterials. J. Funct. Biomater. 2023, 14, 344. https://doi.org/10.3390/jfb14070344

AMA Style

Vishnu J, Kesavan P, Shankar B, Dembińska K, Swiontek Brzezinska M, Kaczmarek-Szczepańska B. Engineering Antioxidant Surfaces for Titanium-Based Metallic Biomaterials. Journal of Functional Biomaterials. 2023; 14(7):344. https://doi.org/10.3390/jfb14070344

Chicago/Turabian Style

Vishnu, Jithin, Praveenkumar Kesavan, Balakrishnan Shankar, Katarzyna Dembińska, Maria Swiontek Brzezinska, and Beata Kaczmarek-Szczepańska. 2023. "Engineering Antioxidant Surfaces for Titanium-Based Metallic Biomaterials" Journal of Functional Biomaterials 14, no. 7: 344. https://doi.org/10.3390/jfb14070344

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

Article Metrics

Back to TopTop