Next Article in Journal
Fatigue Estimation Using Inverse Stamping
Next Article in Special Issue
A First-Time Investigation into Ecofriendly and Biocompatible Mg-Se Binary System for a Greener Earth
Previous Article in Journal
The Activation Energy of Viscous Flow and Liquid–Liquid Structure Transition in Co-B Alloys
Previous Article in Special Issue
Dependence of Electrochemical Characteristics of a Biodegradable Fe-30Mn-5Si wt.% Alloy on Compressive Deformation in a Wide Temperature Range
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Corrosion of Fixed Orthodontic Appliances: Causes, Concerns, and Mitigation Strategies

by
António Fróis
1,2,3,
Ana Cristina Santos
1,2,3 and
Cristina Santos Louro
1,*
1
Department of Mechanical Engineering, CEMMPRE, ARISE, University of Coimbra, Rua Luis Reis Santos, 3030-177 Coimbra, Portugal
2
Faculty of Medicine, Biophysics Institute, Coimbra Institute for Clinical and Biomedical Research/Centre for Innovative Biomedicine and Biotechnology (iCBR/CIBB), University of Coimbra, 3000-548 Coimbra, Portugal
3
Area of Environment Genetics and Oncobiology (CIMAGO), University of Coimbra, 3000-548 Coimbra, Portugal
*
Author to whom correspondence should be addressed.
Metals 2023, 13(12), 1955; https://doi.org/10.3390/met13121955
Submission received: 16 October 2023 / Revised: 17 November 2023 / Accepted: 24 November 2023 / Published: 29 November 2023
(This article belongs to the Special Issue Feature Papers in Biobased and Biodegradable Metals)

Abstract

:
The orthodontic supply market is a prosperous billion-dollar industry, driven by an increasing demand for orthodontic appliances. The supremacy of metallic first-generation biomaterials is evident for manufacturing brackets, archwires, bands, and other components due to their well-recognized chemical inertness, spontaneous passivation, biocompatibility, and favorable mechanical properties combination. However, the oral cavity is the ultimate corrosion-promoting environment for any metallic material. In this work, the general picture of the intraoral degradation of fixed orthodontic appliances is first addressed, from the causes to the harmful effects and their oral clinical implications. Current mitigation strategies are also pointed out, including the alloys’ bulk composition adjustment combined with new and advanced manufacturing processes and/or their surface treatment or coating deposition. The versatile use of thin films and coatings stands out with different deposition technologies: Many in vivo and in vitro efforts have been devoted to oral aging, from monolithic to composite architectures and micro- to nano-scale materials, to meet the best and safest oral practice demands. Unfortunately, literature data suggest that even the existing commercially available protective coatings have drawbacks and are fallible. Further multidisciplinary research is still required to effectively mitigate the corrosion behavior of fixed orthodontic appliances.

1. Introduction

Orthodontics may be defined as the
branch of dentistry that is concerned with the supervision, guidance and correction of the growing and mature dentofacial structures. It includes the diagnosis, prevention, interception and treatment of all forms of malocclusion of the teeth and associated alterations in their surrounding structures”.
[1]
Malocclusions—usually referred to as “crooked” or “misaligned teeth”—are a worldwide dental problem [2,3,4,5]. Technically, a malocclusion is not a disease, but rather aesthetical and/or functional misalignments between the dental arches or teeth irregularities (beyond what is considered a normal biological variation), as shown in Figure 1. Still, malocclusions can cause susceptibility to trauma and periodontal diseases [2,4,6,7,8,9]. Standard treatments for dental malocclusions involve removable or fixed orthodontic appliances.
Fixed appliances are, in general, more effective than removable ones—especially for more complex situations and/or for adult patients—and incorporate brackets, archwires, tubes, and/or bands, tightened by metallic or polymeric ligatures [10,11] (Figure 2). During treatment, a constant load is transferred from the brackets to the teeth, by using orthodontic archwires (attached to the brackets), obtaining tooth movement while adjacent bone and tissue are remodeled [12]. Typical loads for tooth movement using fixed appliances are summarized in Table 1, usually involving values lower than 1 N.
Figure 1. Some common malocclusion types: (a) open bite; (b) missing tooth; (c) crowding; (d) spacing; (e) impacted tooth; (f) overjet; (g) overbite; (h) underbite or anterior crossbite; (i) posterior crossbite. Adapted from [13,14,15,16,17,18,19,20,21]. Sources: (a) reproduced with permission of Oxford University Press on behalf of the European Orthodontic Society, from [13]; (b) reproduced with permission of Dental Press Publishing from [14]; (c) reproduced with permission from [15]; (d) reproduced with permission from [16], © Georg Thieme Verlag KG; (e) reproduced with permission of Dental Press Publishing from [17]; (f) reproduced with permission from [18]; (g) reprinted from [19], Copyright (2007), with permission from Elsevier; (h) reproduced with permission of Journal of Clinical Orthodontics from [20]; (i) reproduced with permission of Dental Press Publishing from [21].
Figure 1. Some common malocclusion types: (a) open bite; (b) missing tooth; (c) crowding; (d) spacing; (e) impacted tooth; (f) overjet; (g) overbite; (h) underbite or anterior crossbite; (i) posterior crossbite. Adapted from [13,14,15,16,17,18,19,20,21]. Sources: (a) reproduced with permission of Oxford University Press on behalf of the European Orthodontic Society, from [13]; (b) reproduced with permission of Dental Press Publishing from [14]; (c) reproduced with permission from [15]; (d) reproduced with permission from [16], © Georg Thieme Verlag KG; (e) reproduced with permission of Dental Press Publishing from [17]; (f) reproduced with permission from [18]; (g) reprinted from [19], Copyright (2007), with permission from Elsevier; (h) reproduced with permission of Journal of Clinical Orthodontics from [20]; (i) reproduced with permission of Dental Press Publishing from [21].
Metals 13 01955 g001
A standard comprehensive orthodontic treatment may last approximately 2 years [23] and involves three sequential phases: (1st) leveling and aligning; (2nd) correction of molar relationship and space closure; and (3th) detailing and finishing [24].
In contemporary orthodontics, the market supply entails a worldwide billion-dollar industry that is expected to grow in the next few years [25]. Metallic materials are still the first choice for manufacturing fixed appliances due to their balanced set of mechanical, biological, and chemical properties [26]. Up to now, the most commonly used metallic alloys include stainless steel (SS), pure titanium (Ti) and its alloys—especially nickel–titanium (NiTi)—and cobalt–chromium (CoCr) alloys. Other metallic materials can also be found in fixed orthodontic appliances, but with a lower application range.
Table 1. Typical loads for orthodontic tooth movement (adapted with permission of Elsevier, from [27]; permission conveyed through Copyright Clearance Center, Inc.).
Table 1. Typical loads for orthodontic tooth movement (adapted with permission of Elsevier, from [27]; permission conveyed through Copyright Clearance Center, Inc.).
MovementDescriptionLoad (N)
TippingPredominant movement of the dental crown in the opposite direction.0.34–0.59
Bodily movement (translation)Movement that tilts the tooth until its root is in the vertical direction.0.69–1.18
Root uprightingPredominant movement of the root.0.49–0.98
RotationRotation of the tooth around its long axis.0.34–0.59
ExtrusionMoving the tooth in the opposite direction to the supporting alveolar bone.0.34–0.59
IntrusionMoving the tooth into the supporting alveolar bone.0.10–0.20
A clinical concern during orthodontic treatments is intraoral corrosion. Always associated with metallic ion release into the oral cavity, corrosion can be intensified by dental plaque accumulation and/or mechanical actions such as friction and fatigue stress. Several important consequences of this undesirable degradation may arise, namely enamel discoloration and demineralization, hypersensitivity, inflammatory reactions and local pain, and, in more severe cases, toxicity effects [28,29,30,31,32].
The need to modify the orthodontic alloys has been identified. Current research guidelines point in two main directions: (i) to adjust the alloys’ bulk composition combined with new and advanced manufacturing processes; or (ii) to modify their surface, while taking advantage of the excellent mechanical properties of the bulk. The composition and microstructure of the surface can be altered by using chemical or physical methods, either by treatment or coating deposition.
The present overview is schematized in Figure 3 and partitioned into five sections. After this introductory section (Section 1), Section 2 focuses on metallic corrosion, the oral environment’s aggressiveness, and its impact on orthodontic alloys. Section 3 comprises an overview of the harmful effects and clinical impact of intraoral corrosion on fixed appliances. A brief condensation concerning the mitigation strategies is presented in Section 4, describing current modifications of the orthodontic alloys. Finally, due to the utmost importance of the surface properties of any biomaterial, Section 5 provides a comprehensive overview of the protective coatings in orthodontics, from metals to ceramics and polymers, as well as their composite architectures with different reinforcement materials.

2. Metallic Corrosion

This chapter focuses on the main alloys used for the manufacturing of orthodontic appliances, the characteristics of the oral environment, and their effects on the corrosion behavior of metallic alloys.

2.1. Orthodontic Alloys

Metals and alloys thrive in the medical field and are more employed as biomaterials than any other material type [26]. Today, the major metallic alloys used in orthodontic applications include stainless steel (SS), pure titanium (Ti) and its alloys—especially the nickel–titanium (NiTi)—and cobalt–chromium (CoCr) alloys. Some of the main characteristics of these bioalloys, in comparison to human molar tooth enamel, are summarized in Table 2.
Stainless steels are iron (Fe)-based alloys containing at least 12% chromium (Cr) and a maximum of 1.2% carbon (C), according to the European Standard EN 10088-1 [33] (Table 3). SS are outstanding materials for manufacturing brackets, bands, tubes, and ligatures [11,34,35], namely the austenitic 3xx series-AISI (American Iron and Steel Institute: 302, 303, 304L, and 316L), the precipitation hardening (PH) steels, as well as the duplex steels (SAF 2205) [10,36,37,38,39,40]. Together with Ti alloys, SS archwires are frequently used in an orthodontic treatment, especially during the 2nd and 3rd phases [41,42].
Table 2. Main characteristics of bioalloys used for manufacturing orthodontic components [37,43,44,45,46].
Table 2. Main characteristics of bioalloys used for manufacturing orthodontic components [37,43,44,45,46].
Main CompositionYoung’s Modulus (GPa)Yield Strength (MPa)
Human molar tooth enamelCalcium phosphate hydroxyapatite70–115
Stainless steel
(AISI 316L)
Fe–Cr–Ni160–187960–1500
Cobalt–chromiumCo–Cr–Fe–Ni150–217830–1200
α-TitaniumTi (grade 4)104550
β-TitaniumTi–Mo–Sn–Zr60–68620–690
Ti-6Al-4VTi–Al–V (grade 5)100–110830–1070
Nickel–titaniumNi–Ti32–36200–500
The major advantages of these Fe–Cr alloys include their good corrosion resistance combined with their outstanding biomechanical behavior and affordable price. The key feature of the corrosion behavior is the Cr content, which is between 16 and 25 wt.% for austenitic (face-centered cubic structure, FCC) Fe–Ni–Cr alloys (Table 3). The Cr element in the solid solution phase of SS alloys allows the development of the typical external protective chromium oxide (Cr2O3) thin film. Other bulk alloying elements of SS include molybdenum (Mo) and nickel (Ni > 8%), which improve the corrosion resistance effectiveness: while Ni promotes the formation of the FCC structure, Mo stabilizes the Cr-based passive layer.
However, some concerns regarding oral corrosion resistance, despite the presence of a small molybdenum (Mo) content (Table 3), and the overall biocompatibility led to the emergence of alternatives [10,11,37,42]. The high Ni nominal content in SS alloys can cause contact dermatitis (see Section 3), which has been encouraging for the development of new Ni-“free” austenitic stainless steels (see Section 4).
CoCr-based alloys have been used in orthodontics since the 1960s for manufacturing brackets and archwires [37,42,48]. With higher Cr content (>20%), these alloys surpass the SS ones in corrosion resistance—mainly in chloride environments due to the Cr-rich oxide passive layer—and biocompatibility, with higher wear resistance [42]; yet, improved ductility and resilience may be achieved (Table 2). The foremost drawbacks reported in the literature include additional heat treatments to improve mechanical performance and a more complex soldering process [10]. Currently, CoCr-based wires are commercially available in four color-coded variations according to the heat treatment applied; the blue one (“soft”) is the most used due to its low yield strength compared to stainless steels [37] (Table 2).
Ti and its alloys are among the most biocompatible materials and were introduced in orthodontics in the 1980s, gaining popularity for brackets, tubes, and archwires production [11,49]. This class of metallic materials presents outstanding mechanical properties, excellent corrosion resistance (better than SS), in addition to low density (4.5 against 7.8 g/cm3 for SS), providing a very high strength-to-weight ratio and non-eliciting allergic responses. Commercially pure titanium (α-Ti, Grade 4) and/or Ti-6Al-4V (Grade 5) brackets and β-Ti (including titanium molybdenum alloy—TMA) archwires are examples of some Ni-free components with outstanding corrosion resistance and biocompatibility [10,50]. Ti-based brackets and tubes reduce bonding failure to enamel, whereas TMA wires are ideal for certain (but not all) orthodontic situations due to the right balance of mechanical properties (e.g., low stiffness and high stringback and formability) and weldability [10,37,42]. The low elastic modulus supports the selection of β-Ti and/or NiTi alloys (Table 2) for orthodontic wires. High manufacturing cost is the most negative drawback [37].
Particular attention should be given to additional Ti-based alloys, such as NiTi and Cu–NiTi alloys, due to its high Ni nominal content. Nitinol®—which stands for “Nickel Titanium Naval Ordnance Laboratory”, with near-equiatomic Ni and Ti concentrations—revolutionized orthodontics since its introduction into clinical practice in 1972 [37,51,52]. Due to its distinct mechanical properties, such as shape memory (shape memory alloy—SMA) and superelasticity behavior, this class of metals is now extensively used for the manufacture of orthodontic wires. [42,49,53]. In fact, the initial leveling stage of the orthodontic treatment (Section 1) usually involves NiTi archwires [41,42]. While the shape memory effect allows for the spontaneous recovery of the component form after being subjected to deformation higher than its elastic limit (by heating), the superelasticity tolerates a constant stress as the strain increases. After the initial elastic stress region and the stress/strain release, the NiTi alloy springs back to its original shape. Thus, high elasticity, spring back, and stored energy (Table 2) enable low-force delivery, even when malocclusions involve extreme teeth crowding. To further increase the alloys’ strength and reduce energy loss, NiTi alloys have been chemically modified by copper addition (5–6% Cu)—the Cu–NiTi alloys—by acquiring a thermally activated behavior [54,55,56]. These wires yield lower loads on the teeth and also on deformation percentage; thus, teeth movement proceeds in a more physiological manner, preventing necrosis, hyalinization areas, and the probability of root resorption [57,58]. Other elements, such as Fe and Cr, are also added to Ni–Ti-based SMA alloys to modify their mechanical properties [59]. All Ti-based alloys spontaneously passivate by generating a titanium oxide protective film that provides good oral corrosion resistance.
Other metallic alloys can be found in fixed orthodontic appliances, but with a lower application range. The use of gold (Au)—precious metal-based alloys—for instance, was widespread before 1950 due to its higher corrosion resistance compared with alternative alloys at that time [10,60]. However, high cost and poor mechanical properties (low hardness) undermine its use, even though Au-based and Au-coated aesthetic components are still available today [10,60,61,62,63,64].

2.2. Intraoral Environment

The human body is an extreme environment for any metallic biomaterial [65], and the mouth is its “portal entry” [66]—an “open ecosystem” [67] in which variations in intraoral parameters are frequent and complex, leading to a unique corrosion-promoting medium.
Human saliva—99.5% water, 0.3% proteins, and 0.2% organic compounds—plays multiple important physiological functions, not only in taste, digestion, and speech but also in teeth and tissue lubrication/protection, pH buffering, and microbiological control [68,69,70,71]. The main functions of saliva and its constituents are presented in Table 4. This summary intends to reflect the saliva complexity, which is further exacerbated by other factors mentioned during this review.
Proteins and glycoproteins from saliva rapidly adhere to teeth enamel and any other surface placed inside the oral cavity to form a thin layer (70–100 nm), making them an important natural lubricant and oral protective film [69,73,74]. The most relevant chemical components of saliva include inorganic ions (e.g., N+, K+, Cl, F, CHO3, PO43−, …), antimicrobial factors, nitrogenous compounds, enzymes, immunoglobulins, albumin and other proteins, and glucose, among others [66,69,71,72,75]. Moreover, the chemical composition, temperature, and pH of human saliva vary between individuals and along the course of the day (circadian rhythms), also depending on the person’s lifestyle, diet, and health/disease conditions [72,73,76,77,78,79,80].
Intraoral mean temperature usually ranges around 33–37 °C [76], but abrupt variations up to 65 °C can occur (e.g., drinking a hot coffee after eating an ice cream or drinking a glass of ice water) [77,81]. The pH of non-stimulated saliva—i.e., without consuming food or drinks—usually varies between 6 and 7, but may also oscillate from 5.3 to 7.8 [66,76,82]. A pH value below 5.5 facilitates the development of dental caries [82,83,84]. An acidic diet can also reduce intraoral pH to 3, for instance, due to acidic soft drinks and fruit juices (pH from 1 to 6) [68,82,85,86]. Another possible contributor to salivary pH fluctuation is regurgitated stomach acid, which has a typical pH value of 1.2 [68]—one of the intraoral problems of bulimic people and oncological patients.
The oral environment is additionally ideal for the inevitable proliferation of microorganisms. So far, over 700 bacterial species have been identified, as well as numerous fungi and viruses [66,67,73,87]. The oral microbiota co-evolved with humans in a mutualistic or even symbiotic manner: While the host provides excellent physiochemical and nutritional conditions, microorganisms (especially bacteria) play important physiological roles, including digestion, oral mucosa cell differentiation, and protection against exogenous pathogens [88,89]. Table 5 provides a summary of the main recognized oral microbiomes.
Planktonic (i.e., non-attached, free-floating) bacteria are 1000 times more vulnerable to antimicrobials than when aggregated. Therefore, some species—the primary colonizers—soon physically associate with and then adhere to the glycoprotein-based film over the teeth and biomaterials’ surfaces. The mechanism of oral biofilm adhesion is described in Figure 4. Other bacterial species adhere and proliferate along with primary colonizers, forming microcolonies imbedded in an extracellular polysaccharide matrix. At this point, an oral biofilm (the dental plaque) grows: Complex groups of microcolonies positively interact with each other and even form a “primitive circulatory system” [67]. The grown (mature) oral biofilm is therefore advantageous to its inhabitants by providing nutrients to and protecting both aerobic and anaerobic colonizers—even against drugs, antimicrobial factors from saliva, and phagocytic cells [67,89,90].
While dental plaque consumes remaining food inside the mouth and protects the teeth against mechanical and chemical injuries (e.g., enamel demineralization), caries and periodontitis may occur if the host/dental plaque relationship is disturbed [73,88,89]. Some species are pathogens, and the microbiological activity of dental plaque releases several by-products into the oral cavity that can modify the chemical composition, oxygenation, and oral pH values [73,88,89]. Saliva and self-cleansing by the cheeks and tongue can naturally control biofilm growth to a certain extent. Nevertheless, oral hygiene procedures are crucial for removing dental plaque, including mechanical brushing with fluoride-containing toothpastes and mouth rinsing with fluorinated mouthwashes and elixirs [78,88,89,91,92].
In short, the intraoral environment is a highly dynamic and complex system—an ultimate degradation-promoting scenario for any biometallic material. Corrosion is a necessary (but not sufficient) condition for causing adverse biologic effects during the use of fixed orthodontic appliances.

2.3. Corrosion of Metallic Alloys

Metallic corrosion can be expressed as a “physicochemical interaction between a metal and the environment that results in changes in the properties of the metal, and which may lead to significant impairment of the function of the metal, the environment, or the technical system, of which these form a part” [93].
In an aqueous environment, such as the intraoral cavity, corrosion initiates through electrochemical reactions in the metal/solution interface, involving the anodic dissolution of the metal, M e to M e a q 2 + (oxidation, Equation (1)), and the cathodic reduction of an oxidant from the solution, O x a q to R e d e r e d o x a q (reduction, Equation (2)), that is [94]:
M e M e a q 2 + + 2 e m e t a l
2 O x a q + 2 e m e t a l 2 R e d e r e d o x a q
The general charge-transfer reaction for a divalent metal can be written according to Equation (3) [94]:
M e + 2 O x a q M e a q 2 + + 2 R e d e r e d o x a q
Dissolved oxygen is usually the cathodic reactant—e.g., according to Equation (4)—with the production of hydroxide ions. However, other mechanisms could be more relevant under acidic conditions since protons may accept electrons produced during the anodic reaction. Typical cathodic reactions under low pH include both Equations (5) and (6), producing water or hydrogen gas, respectively [26,42,90].
1 2 O 2 + H 2 O + 2 e 2 O H
O 2 + 4 H + + 4 e 2 H 2 O
2 H + + 2 e H 2
The end result of corrosion is the release of metallic ions into the electrolyte, whose extent depends on the electrolyte’s nature, including its chemical composition, dissolved oxygen, and pH [42,95]. Moreover, the type of metal or alloy, its manufacturing process, and surface finishing will also influence corrosion [28,39,95,96,97,98].
Most bioalloys—mainly those containing Cr and Ti—rely on the spontaneous formation of a surface protective film. This thin oxide-based layer (some nm thick) may act as a barrier to the movement of ions—a passive film—protecting the metallic substrate against additional electrochemical corrosion [42,93,94]. Passivation is well established, being a spontaneous equilibrium of precipitation and dissolution of ions, with both active and passive films in contact with the electrolyte [99,100].
A simple passivation model [97] is presented in Figure 5. After surface hydroxylation, both metallic ions (from the alloy) and oxide ions (from the aqueous saliva) migrate through the forming film to the electrolyte and to the metal/oxide layer interface, respectively (reactions A and B, Figure 5). The electrons produced both from passivation and corrosion processes will be consumed by oxidants through redox mechanisms (reaction C, Figure 5). Therefore, electronic conduction will be essential to maintain the passivation equilibrium and the oxide film thickness. In other words, the dissolution rate must be lower than the precipitation rate [97,99]. In general, the anodic reaction of oxidation responsible for the passive film formation can be represented by Equation (7):
n M e + m H 2 O M e n O m + 2 m H + + 2 m e
where M e n O m is the top metallic oxide [100].
It is well recognized that some elements—such as manganese (Mn), Mo, Co, and Ni—considering the normal composition of metallic alloys do not participate in passive film formation and are preferentially released into aqueous medium. If those metallic ions bond with biomolecules in vivo, they may present a certain degree of toxicity [99]. Others, however, such as Cr and Ti, are responsible for the growth of oxide films and are most likely to be less toxic [99]. The high corrosion resistance of Fe-based alloys containing Cr (16 < Cr < 25%) is due to the formation of a Cr (III) oxide-hydroxide passive layer [37,101]. The insoluble Cr2O3 therefore prevents iron dissolution. Similarly, TiO2-based passive layers spontaneously form on surface Ti and its alloys, being well recognized for their excellent stability [10,26,102].
Unfortunately, passive layers are fallible [95], that is, top oxide film dissolution and/or disruption may occur, driven by several environmental parameters, including fluctuations in the electrolyte chemical composition. Complexing anions can capture metallic ions from the oxide surface region (reaction D1, in Figure 5), facilitate their migration, dissolve the passive film, and release metallic ions into the saliva (reaction D2, in Figure 5) [97,103]. Aggressive ions such as chlorine (Cl), fluoride (F) [29,94,100,104], or sulphide (S2−) [105], as well as mechanical damages [26,39,73,106], can disrupt the external passive oxide layer, exposing the active metallic surface for corrosion reactions to proceed. The deleterious effect of the fluoride ions—common in fluorinated prophylactic gels and mouthwash elixirs—on both SS- and Ti-based alloy passivation can be described by the following reactions, Equations (8) and (9), respectively [107].
C r 2 O 2 + 12 F + 8 H + 12 C r H F 6 2 + 3 H 2 O
T i O 2 + 6 F + 4 H + T i F 6 2 + 2 H 2 O
Furthermore, both the chemical composition and the presence of defects and impurities [100] influence the thickness, stability, and composition of the external passive films [94]. Those variables, in turn, are determined by the manufacturing process and the corresponding surface finishing—mechanical/physicochemical (polishing, electro-erosion, plasma, acid/alkaline treatments, electrochemical deposition) or biochemical treatments (biological molecules to mimic biointerfaces). In the end, the balance between passive layer damage and its reconstitution will dictate the overall final corrosion resistance behavior of the biometallic alloys [94,100].

2.4. Oral Corrosion Forms

Intraoral degradation of metallic appliances is inevitable. Until today, nine basic corrosion types have been reported: uniform, galvanic, crevice, pitting, intergranular, selective leaching, stress, erosion, and microbiologically induced corrosion (MIC), which will be briefly addressed as follows.

2.4.1. Uniform Attack

Uniform corrosion is recognized as general corrosion. This degradation form seems to occur uniformly over the entire surface due to surface electrochemical reactions, almost at the same rate [81,93]. It is the most common type of oral corrosion, affecting all metallic materials at different rates [95], but sometimes it is difficult to detect—only when a significant amount of material is dissolved. All parts of a metallurgical and compositionally uniform surface should be accessible to the electrolyte (saliva) [29].

2.4.2. Galvanic Corrosion

Galvanic corrosion is observed when two different metallic surfaces—with different corrosion potentials—are joined or sufficiently close in an electrolyte solution by establishing a galvanic coupling. The more electropositive (less noble) metal or alloy becomes the anode and preferentially corrodes [26,30,81,93,95].
In orthodontics, contact between dissimilar metallic surfaces might occur in two situations: By simple contact [29,108] or through bonding processes [10,29,109]. In the first case, bracket/wire interactions are inevitable during orthodontic treatment with fixed appliances. Predictably, in certain combinations, such as in the so-common NiTi wire/SS bracket, galvanic corrosion is susceptible to occurring [29], especially in a fluoride-rich environment [108]. Recent research work [60] reported no evidence of galvanic coupling between SS lingual brackets and SS archwires, but the authors suggested caution when using fluoride-containing products during fixed orthodontic treatment with SS brackets and NiTi archwires.
Different parts of brackets or posted archwires are often made of dissimilar alloys, leading to galvanic corrosion [110]. Furthermore, brazing alloys can be used during the manufacturing of orthodontic components [10,29]. Mechanically active welding joints [95] may be reactive, increasing galvanic corrosion susceptibility accompanied by toxic metallic ion release, particularly for silver (Ag)-, copper (Cu)-, and zinc (Zn)-based welding materials [29,111,112]. A recent in vitro study [113] demonstrated that Ag ion release from Ag-soldered SS bands was an order of magnitude higher than other non-soldered SS orthodontic appliances. The authors assigned this effect to the manufacturing process used (welding).
It is generally accepted that galvanic cells can also occur in different locations of the same metallic surface due to non-uniform surface finish (e.g., roughness and chemical composition) and mechanical properties (e.g., work hardening) or even dissimilar properties of the electrolyte (pH and chemical composition) [95]. In the oral cavity, saliva is the main electrolyte, but extracellular fluids such as blood or gingival fluid are also present. Galvanic currents may take place due to the contact of the metallic surface to different biological fluids [73].

2.4.3. Crevice Corrosion

Crevice corrosion is a localized attack occurring in or near constricted places (crevices) formed by two surfaces, of which at least one is metallic [93]. This leads to a local enhancement of aggressive species and depletion of oxygen, in addition to the consequent acidification of the crevice solution due to the hydrolysis of the dissolving metal ions. Generally, metallic materials that show an affinity to pitting also suffer from crevice corrosion. The main causes include differences in metallic ions, fluoride concentration, or oxygenation between the crevice and its surroundings [28,81], associated with a pH decrease and a chloride ion concentration increase [10,29], which deteriorate the protective passive layer—especially on SS alloys [114].
Elastomeric or metallic ligatures are frequently used to fix orthodontic archwires to brackets, establishing ideal sites for crevice attack on brackets [109] (including on 316L SS alloy): Deep craters, fissures, and pores have been detected after long intraoral exposure, as well as extensive deterioration and perforation of the resin-fixed bracket base [28,30,81]. Daems et al. [106] also noticed this type of corrosion at regions of bracket/archwire contact or with plaque and food remnants. Other factors that reportedly cause crevice corrosion comprise the recycling process of the components [29]—not recommended in several countries—surface defects or irregularities [95,106]—including those caused by handling the components by the orthodontist during treatment (Figure 6)—and the presence of welding areas [50,115].

2.4.4. Pitting Corrosion

Pitting corrosion is a localized type of corrosion that initiates on metallic surfaces when the protective passive film disrupts due to mechanical and/or electrochemical attack [93,100], leading to the formation of pit holes and/or cavities. This attack has been associated with other corrosion types [114,116], such as the one caused by the well-adherent biofilm that forms during orthodontic treatments (Figure 7)—discussed further in Section 2.4.9. Aggressive ions in saliva, such as chlorine and fluoride [29,42,117]—especially under acidic conditions [101,104,118]—and food additives, such as certain spices [119,120], effectively damage surface protective oxides. Manufacturing defects on orthodontic metallic components may also increase pitting corrosion susceptibility for both SS and NiTi alloys [28,29,30].
The main strategy to improve the pitting resistance of SS alloys is to increase the Cr and Mo nominal content. However, the presence of non-metallic inclusions, such as manganese sulphide (MnS), is of major importance since pits usually initiate at these precipitates [121,122,123]. Usually, the CoCr-based alloys are resistant to pitting; the dissolution of the protective Cr2O3 layer into soluble ions (CrO42−) takes place by oxidation at potentials below the oxygen evolution range [124].

2.4.5. Intergranular Corrosion

As the name suggests, intergranular corrosion occurs in microstructural planar defects along grain boundaries or in the immediate near zones, with minimal or no attack on the alloy grain itself [93,114]. The net result is an alloy fracture along these grain boundaries. SS alloys—used for manufacturing orthodontic brackets and archwires—are particularly vulnerable to this corrosion form, leading to surface staining, weakening the mechanical behavior (strength and ductility), or even failure [112,120]. Special attention should be given to heat treatment of steels [37] (or brazing/welding [95,115,125]—termed weld decay). For a prolonged period above the sensitization temperature [28,29], the formation of small precipitate particles of chromium carbide (Cr23C6) occurs [114]. Two major consequences arise: the SS brittleness increases and its corrosion resistance decreases, both due to the Cr-depleted zone adjacent to the grain boundary [37,95].

2.4.6. Selective Leaching

Selective leaching or dealloying is found in solid solution alloys, such as Ni–Cr-based or binary alloys containing calcium (Ca) and zinc (Zn) [26], occurring when one element is preferentially removed during the corrosion process [81,114]. This preferential release of a more reactive element from an alloy, regardless of its chemical composition [93], can occur in vivo [26,28]. Still, the effect of selective leaching seems negligible in dentistry [26,81].

2.4.7. Stress Corrosion

Stress corrosion, sometimes termed stress corrosion cracking (SCC), develops due to the influence of both applied tensile stress and a corrosive environment [28,95]. Some alloys that are virtually inert in a particular corrosive medium can become susceptible to this type of corrosion when under loads. This can seriously compromise the mechanical integrity of the material, and failure may eventually occur under low stress levels (compared with alloys in non-corrosive environments) [81,93,114]. Nitinol archwires bonded to brackets are exposed to compressive and tensile stress and might fracture during orthodontic treatment [29,126].

2.4.8. Erosion Corrosion

Erosion corrosion refers to the deterioration of a metallic material due to mechanical abrasion or wear with the combined action of the chemical attack of the corrodent fluid motion. Three subtypes are well known: Erosion, cavitation, and fretting [114]. In orthodontics, fretting corrosion is the most relevant form [81], due to the slight relative motion (vibration and slip) of two contacting metallic surfaces under load [93,114]. Surfaces of both archwires and bracket slots experience load and may undergo a cold-welding phenomenon. In addition, the required small displacements could disrupt the passive films and, consequently, increase corrosion susceptibility (e.g., by pitting) [28,29,95].

2.4.9. Microbiologically Induced Corrosion

As previously mentioned, oral microorganisms can directly or indirectly degrade metallic materials in vivo, either by metabolizing metal from the surface or by modifying the surrounding electrolyte with their metabolic by-products, respectively [95]. This form of corrosion is known as microbiologically induced corrosion (MIC) [73,114,127]. Zarasvand and Rai [90] extensively studied the MIC mechanisms, while Mystkowska et al. [73] described the intraoral process. Accordingly, oral biofilms create differential concentration cells on the metallic surfaces of three main types: oxygen concentration cells, metal concentration cells, and active–passive cells (Figure 7).
Figure 7. (a) SS brackets after two years of intraoral use. Scanning electron microscopy (SEM) and optical micrographs coupled to energy dispersive (EDS) chemical compositions of two distinct zones: SS alloy and pitting morphology underneath a biofilm layer. Adapted from [116]; (b,c) pitting corrosion induced by sulphate-reducing bacteria: confocal scanning microscope images after a 56-day in vitro test using D. nigrificans bacteria on 316LV steel; and (d) representation of the proposed corrosion mechanism by [128] (reproduced from [73]).
Figure 7. (a) SS brackets after two years of intraoral use. Scanning electron microscopy (SEM) and optical micrographs coupled to energy dispersive (EDS) chemical compositions of two distinct zones: SS alloy and pitting morphology underneath a biofilm layer. Adapted from [116]; (b,c) pitting corrosion induced by sulphate-reducing bacteria: confocal scanning microscope images after a 56-day in vitro test using D. nigrificans bacteria on 316LV steel; and (d) representation of the proposed corrosion mechanism by [128] (reproduced from [73]).
Metals 13 01955 g007
Oxygenation cells appear due to a non-uniform biofilm layer—in terms of thickness, ratio of aerobic (oxygen-consuming)/anaerobic microorganisms, or due to the presence of layers of corrosion products—that cause differences in oxygenation throughout the surface. Regions with high oxygen concentrations favor cathodic reactions (Equation (2) and Figure 8a), and the metallic surface below becomes the cathode. Conversely, in a poorly oxygenated environment, the anodic reaction is enhanced (Equation (1)), and the surface becomes the anode and corrodes. Differences in metallic ion concentration on different sites (Figure 8b) also occur due to the nature of the extracellular matrix, which has diverse composition and functional groups with different affinities to metallic ions. Under biofilm regions with low affinity to metallic ions, cathodic reactions further progress, whereas anodic dissolution of the metal increases under high-affinity biofilm sites. Finally, if a dense biofilm layer is mechanically or chemically disrupted (Figure 8c), the exposed metallic surface corrodes (the anode), while biofilm-covered regions behave as cathodes [73,90]—active–passive cells [93].
Certain anaerobic microorganisms, such as sulfur-reducing bacteria (SRB), release corrosive metabolic products that degrade metallic alloys. The SRB can produce hydrogen gas (H2), hydrogen sulfide (H2S), and sulfur difluoride (F2S, a strong local cathode), while other Gram-negative bacteria release butyric acid (C4H8O2) and carbon dioxide (CO2) [73,90,114]. Besides weakening and retarding the passivation mechanism of the metallic surface, H2S is highly toxic to cells [129] and reacts with metals to form metal sulfides and atomic hydrogen. Metal sulfides may precipitate on the surface, generating new active–passive cells, while released atomic hydrogen can cause SCC [73,90,93].
Recovered SS brackets [110] and archwires [106] unequivocally revealed highly adherent biofilm deposits (Figure 9a), under which pitting or crevice corrosion occurred [106,116], as presented in Figure 7 for brackets after two years of intraoral use. Calcium-precipitating bacteria on SS orthodontic wires cause surface calcium deposits, into which chlorine ions penetrate and induce pitting [130].
Abundant biofilm layers were also observed on NiTi archwires [116] (Figure 9b), in spite of the good resistance against MIC corrosion [131,132] of the Ti-based alloys. Still, dental plaque accumulates and might enhance crevice corrosion in the welding gaps of Ti-based brackets [50,98]. Some research works [105,133] suggest that the corrosion susceptibility of Ti and its alloys decreases in the presence of bacteria, namely by inducing oxygenation cells due to bacterial metabolism.

3. Harmful Effects and Clinical Implications

The main consequences of intraoral aging of orthodontic metallic alloys are briefly presented in this chapter, namely the release of metallic ions into the oral cavity, the friction effect between components, and the consequences of using fluoride-based products during treatment with fixed orthodontic appliances.
Aging of metallic alloys is an important issue in orthodontics since both structural and morphological modifications can occur and thus negatively affect the normal clinical treatment progression. Corrosion of metallic surfaces and the presence of biofilm promote metallic ion release and roughness, which may increase friction between brackets and archwires [134,135,136,137,138] and extend the treatment time. Another pointed aging implication is the friction enhancement between the appliance and the mucosa, which causes oral mucosa lesions (from minor wounds to large ulcers), resulting in patients’ pain and discomfort [68,139,140,141]. Moreover, aging decreases the resistance to fracture of metallic alloys under repeated cyclic loading [30,136]—fatigue—and could lead to premature failure of archwires [95,142] ruining the in vivo function of the biomaterial.

3.1. Release of Metallic Ions

Corrosion processes ultimately cause the release of metallic ions and particles into the oral cavity [28,99,143,144], which may interact with oral tissues and move to the gastrointestinal tract; even so, their impact on health is not yet fully understood [42,73,95,145]. Biocompatibility concerns raised among clinicians and researchers as hazardous species such as nickel, chromium, cobalt, copper, and vanadium (Ni, Cr, Co, Cu, and V, respectively) can be released from metallic appliances [26,29,144,146,147,148]. Back in 1975, Samitz and Katz [149]—who reviewed data related to Ni released from implanted prostheses—concluded that solubilized metal was found in tissues near implants in laboratory animals.
Multiple researchers have been trying to quantify the release of metallic ions from orthodontic appliances to assess if the concentrations can reach toxic levels for humans, both in vitro and in vivo [144,150,151,152]. Table 6 compiles an update of the 59 studies found in the literature regarding the in vivo measurements of metallic ions released from orthodontic appliances, sorted out by publication year.
The first study found dates back to 1991, by Gjerdet et al. [153], and measured the Ni and Fe contents in patients’ saliva up to 3 months of usage. The authors found an initially higher salivary metal content that decreased over treatment time, but values were small when compared with those from dietary intake. Nonetheless, they were already alerted to the large interindividual variability found, as well as to Ni-sensitive patients [153].
Through time, Ni and Cr concentrations are almost always focused, but most studies concluded that the salivary metallic ion concentration is well below toxicity levels. In fact, dietary studies conducted in different countries obtained a daily intake of nickel between 100 and 300 μg/day from food and drinking water. Consuming Ni-enriched food (e.g., processed food) may increase this value up to 900 μg/day [95,154,155,156,157,158,159,160]. Haber et al. [161] estimated a toxicity reference value for Ni-sensitized populations of 4 µg Ni/kg of body weight per day, in addition to Ni in food. Concerning chromium, an average daily intake of 50–280 μg has been proposed [95,162]. However, some authors who analyzed different matrixes (oral mucosa cells, dental plaque, bone, gingiva, hair, and internal organs) found evidence of bioaccumulation that may provoke toxic effects, including DNA damage. In fact, Eliades and Athanasios [28] argued that in vivo studies measuring urinary or serum concentrations of metallic ions in orthodontic patients may give falsely lower Ni levels due to its accumulation in an organ. Further research should therefore persist.
Among the metallic ions released into the oral cavity, Ni raises special health concerns and has been systematically studied [26,154,158,159,161,163,164,165,166,167,168,169], including in orthodontics [143,144,168,170,171,172,173,174,175,176,177,178,179,180,181,182]. The European Union (EU) currently forbids the use of Ni [183]:
  • “in any post-assemblies which are inserted into pierced ears and other pierced parts of the human body unless the rate of Nickel release from such post-assemblies is less than 0.2 µg/cm2/week (migration limit)”;
  • “in articles intended to come into direct and prolonged contact with the skin (…) if the rate of Nickel release from the parts of these articles coming into direct and prolonged contact with the skin is greater than 0.5 µg/cm2/week”;
  • “in articles referred to in point 2 where these have a non-nickel coating unless such coating is sufficient to ensure that the rate of nickel release from those parts of such articles coming into direct and prolonged contact with the skin will not exceed 0.5 µg/cm2/week for a period of at least two years of normal use of the article” [183].
Unfortunately, biometallic alloys lie outside of this EU regulation regarding this matter. The American Academy of Pediatrics also expressed concerns regarding the use of Ni-containing alloys, urging the adoption of regulations similar to the EU nickel directive [168]. Dental biomaterials must still comply with several standards and regulations [184].
This transition metal (Ni) is a well-known allergen [185,186], a strong immunologic sensitizer capable of inducing delayed hypersensitive reactions [169,187], triggering cytotoxic, carcinogenic, and mutagenic effects [144,164,188,189], and affecting several cellular functions by long-term exposure to a small amount [144]. Moreover, emphasis has been given to Ni-induced genetic effects, including DNA damage and the inhibition of enzymes involved in DNA reparation [144,180]. The International Agency for Research on Cancer (IARC) classifies Ni (II) and its compounds as carcinogenic or potentially carcinogenic to humans [165].
Chromium is another well-known toxic element. Between the two most stable oxidation states, Cr(III) and Cr(VI), its hexavalent form is toxic and exhibits mutagenic, cytotoxic, and carcinogenic effects in humans [144]. Reportedly, both oxidation states were found in vitro after the corrosion of SS orthodontic brackets in artificial saliva [190].
Ni carcinogenicity, genotoxicity, and allergy are controversial in orthodontics [95,191,192,193,194,195,196,197]. Nonetheless, released Ni from orthodontic components can accumulate in the oral mucosa cells (see studies in Table 6) and decrease cell viability [180], while systemic toxicity should not be ignored [198]. Moreover, Kochanowska et al. [199] showed the in vivo effect of long-term exposure to metal orthodontic appliances on both the metallothionein gene expression and the induction of protein synthesis by using animal models (pigs).
Several subtle to severe intra- and/or extra-oral symptoms of allergic reactions to nickel have been reported due to the use of metallic appliances [31,147,192,200,201,202,203,204,205,206], as exemplified in Figure 10. Symptoms include burning sensation, stomatitis, angioedema, severe gingivitis without dental plaque, gingival hyperplasia, generalized urticaria, and widespread eczema [31,147,203,206,207,208,209]. Besides discomfort and pain for patients, orthodontists may need to replace high Ni-containing components, interrupt the treatment, and/or refer the patient to an allergologist or other specialist for further examination [201,203,204,205].
Ni allergy—namely extreme hypersensitive reactions—is (fortunately) rare in orthodontics [32,191,206], but may be ineffectively diagnosed: Subtle signs are easily misinterpreted as mimicking mechanical injuries or microbiologic activity [95,192,202,210]. Schuster and colleagues [192] reported allergy symptoms related to the presence of fixed appliances during treatment without intraoral signs. Corrosion products induce enamel demineralization, metallic ion incorporation, and color change [211], as well as pain and swelling of oral soft tissues, leading to secondary infections [29]. Pazzini et al. [212] concluded that patients treated with Ni-“free” (0.5–4% Ni) appliances had better gingival health and smaller blood changes when compared with those wearing conventional metallic components (13% Ni). Another possible negative effect is the increase in antibiotic resistance of some bacteria exposed to metals and their potential transfer to medically relevant pathogens [95].
For further comprehension of the toxic effects of metallic ions released during a fixed orthodontic treatment, the reading of the outstanding review works conducted by Martín-Caméan and colleagues [144,213], and by Downarowicz and Mikulewicz [214], is recommended.

3.2. Friction in Orthodontics

Resistance to sliding is present when two surfaces come into contact with each other (e.g., bracket/wire and wire/ligature) [12,37], which is clinically relevant in orthodontics since reduced resistance to sliding can decrease treatment time [215]. Kusy and Whitley [216,217] partitioned resistance to sliding into three components: (i) Friction, “a force that opposes every action that an orthodontist takes to move the teeth”; (ii) binding, when the angle between the bracket slot and the archwire is high enough to promote contact between the bracket corners and the archwire; and (iii) notching, when a permanent deformation of the wire (or bracket) occurs [215,216,217,218,219]. Figure 11 schematizes these three contributions to resistance to sliding in a bracket/archwire contact.
Table 6. Overview of in vivo studies reporting the quantification of metallic ions release from orthodontic appliances. DL: detection limit; d.h.m.: dry hair mass; AAS: atomic absorption spectroscopy; ICP-AES/ICP-OES: inductively coupled plasma atomic/optical emission spectroscopy; ICP-MS: inductively coupled plasma mass spectroscopy; XRF: X-ray fluorescence.
Table 6. Overview of in vivo studies reporting the quantification of metallic ions release from orthodontic appliances. DL: detection limit; d.h.m.: dry hair mass; AAS: atomic absorption spectroscopy; ICP-AES/ICP-OES: inductively coupled plasma atomic/optical emission spectroscopy; ICP-MS: inductively coupled plasma mass spectroscopy; XRF: X-ray fluorescence.
Sample SizeAppliancesMatrix and SamplingElements and Detection ModeMean/Median
Concentrations
Main ResultsReference
34 patientsSS brackets and bands; NiTi archwires.Saliva. Sampling before treatment, right away or more than 3 weeks after application, and 3–5 weeks after removal.Ni and Fe. AAS.Without appliances:
Ni: 8.2 ppb,
Fe: 148 ppb.
Immediately after placement:
Ni: 67.6 ppb,
Fe: 488 ppb.
More than 3 weeks after placement:
Ni: 7.8 ppb,
Fe: 172 ppb.
Significant increase in Ni and Fe concentrations and absolute masses right after placement, but not after 3 months of usage. [153] Gjerdet et al. (1991)
31 patientsSS brackets and archwires; NiTi archwires.Blood. Sampling before treatment, and after 3 months (with NiTi archwires) and 4–5 months (with SS archwires).Ni. AAS. No significant or consistent increase in Ni blood level during orthodontic treatment. [220] Bishara et al. (1993)
47 patientsBrackets, bands and archwires.Saliva. Sampling before and during treatment (1–2 days, 1 week, and 1 month).Ni and Cr. AAS.Before treatment:
Ni: 68 ppb,
Cr: 68 ppb.
During treatment:
Ni: 55–74 ppb,
Cr: 69–90 ppb.
Ni and Cr concentrations were not significantly affected.[221] Kerosuo et al. (1997)
45 patients + 15 controlsMetallic brackets, tubes, and bands; NiTi archwires.Saliva. Sampling before and during treatment (1 week, 1 month, and 2 months).Ni and Cr. AAS.Before treatment:
Controls:
Ni: 1.16 µg/mL,
Cr: 2.20 µg/mL.
Patients:
Ni: 0.53–0.54 µg/mL,
Cr: 1.35–1.41 µg/mL.
During treatment:
Controls:
Ni: 1.33–1.46 µg/mL,
Cr: 2.50–3.43 µg/mL.
Patients:
Ni: 0.49–0.67 µg/mL,
Cr: 0.49–1.98 µg/mL.
Fixed orthodontic appliances do not seem to significantly affect Ni and Cr concentrations in saliva.[177] Kocadereli et al. (2000)
100 patientsSS brackets, bands and archwires; NiTi archwires.Saliva and Serum. Sampling before and during treatment (1 week, 1 month, 1 year and 2 years).Ni and Cr. AAS.Saliva:
Before treatment:
Ni: 4.45 ppb,
Cr: 0.75 ppb.
During treatment:
Ni: 4.12–11.53 ppb,
Cr: 0.53–1.53 ppb.

Serum:
Before treatment:
Ni: 8.36 ppb,
Cr: 6.21 ppb.
During treatment:
Ni: 7.87–10.27 ppb,
Cr: 6.16–10.98 ppb.
The maximum levels of Ni and Cr in saliva were recorded 1 month after starting the orthodontic treatment. Below toxicity levels.
The maximum levels of Ni and Cr in serum were recorded 2 months after starting the orthodontic treatment. Below toxicity levels.
[172] Aǧaoǧlu et al. (2001)
17 patients + 7 controlsSS brackets and archwires.Saliva. Before and after rinsing with distilled water.Ni, Cr and Fe. ICP-AES.Controls before:
Ni: 18 ppb,
Cr: 20 ppb,
Fe: 21 ppb.
Controls after:
Ni: 11 ppb,
Cr and Fe: < DL.
Patients before:
Ni and Cr: < DL,
Fe: 14.
Patients after:
Ni: 10 ppb,
Cr: 27 ppb,
Fe: 17 ppb.
No significant difference between controls and patients. Below toxic levels.[222] Eliades et al. (2003)
55 patients + 30 controlsSS brackets, bands and archwires; NiTi and CoCr archwires.Oral mucosa epithelial cells.Ni, Co. ICP-MS.Controls:
Ni: 0.725 ppb,
Co: 0.202 ppb.
Patients:
Ni: 2.521 ppb,
Co: 0.568 ppb.
Statistically significant differences between controls and patients. DNA damage (Comet assays).[223] Faccioni et al. (2003)
24 patients + 24 controlsSS brackets, bands and archwires; NiTi archwires.Saliva and dental plaque (on enamel and metallic surfaces). Sampling at approximately 16 months after starting the orthodontic treatment.Ni. AAS.Controls:
Filtered saliva Ni: 0.004 µg/g,
Saliva sediment Ni: 14.85 µg/g.
Dental plaque (all tested surfaces) Ni: 0.380–0.875 µg/g.

Patients:
Filtered saliva Ni: 0.005 µg/g
Saliva sediment Ni: 25.25 µg/g
Dental plaque (all tested surfaces) Ni: 0.685–2.690 µg/g
Statistically significant Ni increase in both filter-retained fraction (saliva sediments) and dental plaque.[178] Fors and Persson (2006)
15 patients + controlsOrthodontic appliances (not specified).Hair. During the orthodontic treatment (not specified).Ni. AAS.Controls:
Ni: 0.64 µg/g.
Patients:
Ni: 0.50 µg/g,
No statistically significant different between patients and controls.[224] Levrini et al. (2006)
10 patientsBrackets, bands, and tubes; NiTi and SS archwires.Hyperplastic and healthy gingiva. After 2–4 year-treatments with fixed orthodontic appliances.Ni. AAS.Ni in healthy gingiva: 1.81 μg/g;
Ni in hyperplastic gingiva: 1.32 μg/g.
Non-significant differences between Ni content in healthy and hyperplastic gingiva, but histological differences (toxicity).[210] Gursoy et al. (2007)
21 patientsSS brackets and bandsUrine. Before and 2 months after placementNi. AASBefore placement:
Ni: 17.67 µg/L.
After placement:
Ni: 19.89 µg/L.
Statistically significant increase in urinary Ni 2 months after placement.[225] Menezes et al. (2007)
10 patientsSS brackets, bands, and ligatures; NiTi archwires.Saliva. Sampling before and during treatment (1 and 3 weeks).Ni and Cr. AAS.Before treatment:
Ni: 31.62 ppb,
Cr: 38.82 ppb.
After 1 week:
Ni: 113.20 ppb,
Cr: 83.15 ppb.
After 3 weeks:
Ni: 65.24 ppb,
Cr: 64.21 ppb.
Statistically significant increase in salivary Ni and Cr concentrations during the orthodontic treatment when compared with the basal levels. [181] Singh et al. (2008)
30 patients + 30 controlsSS brackets, bands and archwires; NiTi archwires,Saliva and mucosa cells. 1 sampling,Ni, Cr, and Co. AAS.Controls:
Ni: 12.26 ppb,
Cr: 3.46 ppb,
Co: 0.44 ppb.
Patients:
Ni: 21.74 ppb,
Cr: 4.24 ppb,
Co: 0.84 ppb.
Significantly higher Ni concentrations in patients when compared with controls. No differences regarding Cr and Co levels.[226] Amini et al. (2008)
30 patientsSS bracketsSaliva. Sampling before and after 10 min, 24 h, 7 days, 30 days, and 60 days of usage.Ni, Cr and Fe. AAS.Before treatment:
Ni: 5.25 µg/L,
Cr: 0.64 µg/L,
Fe: 94.03 µg/L.
During treatment:
Ni: 1.69–16.01 µg/L,
Cr: 0.52–1.72 µg/L,
Fe: 28.31–103.58 µg/L.
Concentration peak for Ni and Cr 10 min after placing the orthodontic appliances, but no significant variations for all metals throughout the study time.[171] De Souza and Menezes (2008)
18 patientsSS brackets and bands; NiTi archwires.Saliva. Before and during the orthodontic treatment (immediately after placing the SS components; immediately before and after placing the NiTi archwires; and after 4 and 8 weeks).Ni. ICP-MS.Before placing any component:
Ni: 34 µg/L.
After placing SS components:
Ni: 78 µg/L.
After placing the archwires:
Ni: 56 µg/L.
During the remaining study time points:
Ni: 28–34 µg/L.
Statistically significant increase in salivary Ni immediately after placing the SS and NiTi components.[143] Petoumeno et al. (2009) and [131] Petoumeno et al. (2008)
15 patients, divided according to the bracket typeSS, Ti, and Ni-free brackets and tubes.Oral mucosa cells. Sampling before and 30 days after placing the appliances.Ni, Cr, Fe, Ti, Co, Mn, and Mo. ICP-MS.Before treatment:
Ni: 3.44 µg/L,
Cr: 0.00 µg/L,
Fe: 1.95 µg/L,
Ti: 0.98 µg/L,
Co: 0.00 µg/L,
Mn: 0.32 µg/L,
Mo: 0.13 µg/L.
After 30 days:
Ni: 0.00–0.04 µg/L,
Cr: 0.00–0.34 µg/L,
Fe: 1.24–5.36 µg/L,
Ti: 0.82–3.04 µg/L,
Co: 0.00 µg/L,
Mn: 0.58–1.08 µg/L,
Mo: 0.00 µg/L.
Increased Ti and Mn in cells exposed to the SS components. Higher Cr and Fe detected in cells exposed to Ni-free components. Increased Mn in cells exposed to Ti components. Ti components are the most biocompatible.[227,227] Fernández-Miñano et al. (2011)
20 patients + 20 controlsSS brackets and bands; SS and NiTi archwires.Oral mucosa cells. Sampling after debonding (removal) of the orthodontic appliances, and 30 days later. Minimum treatment time of 18 months.Ni and Cr. ICP-MS.Immediately after debonding:
Controls Ni: 3.86 ppb,
Controls Cr: 2.71 ppb,
Patients Ni: 4.09 ppb,
Patients Cr: 3.63 ppb.
30 days after debonding:
Controls Ni: 3.48 ppb,
Controls Cr: 2.26 ppb,
Patients Ni: 3.83 ppb,
Patients Cr: 2.94 ppb.
No statistically significant differences between the groups regarding Ni and Cr concentrations. Genotoxic damage during orthodontic treatment reverted.[228] Natarajan et al. (2011)
28 patients divided in 4 groups according to the brackets/archwires type combination+ 18 controlsSS brackets and bands; SS and Ti brackets; SS and NiTi archwires.Oral mucosa cells. Sampling before and during treatment (3 and 6 months after placement).Ni and Cr. AAS.Before treatment:
Ni: 0.52 ppb,
Cr: 0.31 ppb.
After 3 months:
Ni: 0.68 ppb,
Cr: 0.41 ppb.
After 6 months:
Ni: 0.78 ppb,
Cr: 0.78 ppb.
Statistically significant increase in Cr content fat 3 months, as well as in Ni and Cr content after 6 months. DNA damage. SS brackets/SS archwire combination shows higher biocompatibility, whereas the Ti brackets/NiTi archwires are the less biocompatible.[180] Hafez et al. (2011)
20 patients divided according to the bracket typeSS conventional or self-ligating brackets and bands; NiTi archwires.Saliva. Sampling before and during treatment (1, 7 and 30 days after placement).Ni and Cr. AAS.Before treatment:
Ni: 0.68 µg/L,
Cr: 5.19–6.06 µg/L.
After 1 day:
Conventional—Ni: 1.95 µg/L, Cr: 21.78 µg/L.
Self-ligating—Ni: 2.72 µg/L, Cr: 10.65 µg/L.
After 7 days:
Conventional—Ni: 2.89 µg/L, Cr: 36.69 µg/L.
Self-ligating—Ni: 4.95 µg/L; Cr: 14.34 µg/L.
After 30 days:
Conventional—Ni: 1.18 µg/L; Cr: 8.98 µg/L.
Self-ligating—Ni: 1.12 µg/L; Ce: 6.31 µg/L
Non-significant increase in Ni and Cr concentrations during the orthodontic treatment. Below dietary levels.[170] Sahoo et al. (2011)
30 patients + 30 controlsHyrax appliances (which include 4 orthodontic bands with Ag-based soldering).Saliva. Sampling before and after placement (10 min, 24 h, 7 days, 30 days, and 60 days).Cd, Cu, Zn and Ag. AAS.Controls:
Cd: 0.15–0.18 µg/L,
Cu: 3.81–6.54 µg/L,
Zn: 0.01 µg/L,
Ag: < DL.
Patients before placement:
Cd: 0.14 µg/L,
Cu: 16.98 µg/L,
Zn: 0.05 µg/L,
Ag: 0.80 µg/L.
Patients after placement:
Cd: 0.18–0.71 µg/L,
Cu: 12.63–70.60 µg/L,
Zn: 0.05–0.20 µg/L,
Ag: 2.01–11.53 µg/L.
All ions showed significant increases 10 min after placing the orthodontic appliances.[229] Freitas et al. (2011)
28 patients + 18 controlsOrthodontic appliances (non-specified).Hair. 1.5–2 years orthodontic treatment.Ni, Cr, Mn, and Fe. ICP-OES.Controls:
Ni: 0.3642 µg/g d.h.m,
Cr: 0.1298 µg/g d.h.m,
Mn: 0.4850 µg/g d.h.m,
Fe: 11.74 µg/g d.h.m.
Patients:
Ni: 0.5073 µg/g d.h.m,
Cr: 0.1331 µg/g d.h.m,
Mn: 0.5739 µg/g d.h.m,
Fe: 12.22 µg/g d.h.m.
Non-significant differences in the hair metal contents between controls and patients.[230] Mikulewicz et al. (2011c)
28 patients + 28 controlsSS brackets, and bands archwires.Saliva. 12–18 months.Ni and Cr. AAS.Patients:
Ni: 18.5 ng/mL, Cr: 2.6 ng/mL.
Control:
Ni: 11.9 ng/mL, Cr: 2.2 ng/mL.
Statistically significant difference for Ni between the two studied groups. Below toxic levels.[231] Amini et al. (2012)
16 patientsSS brackets and tubes; NiTi archwires.Saliva. Before the orthodontic treatment, immediately after and 8 weeks after placing the NiTi archwires.Ni. ICP-MS.Before treatment:
Ni: 32 µg/L.
Just after placing the NiTi archwire:
Ni: 51 µg/L.
8 weeks after placing the NiTi archwire:
Ni: 34 µg/L.
Statistically significant Ni increase just after placing the NiTi archwire. Non-significant difference after 8 weeks.[232] Ousehal and Lazrak (2012)
20 patientsSS brackets and bands; SS and NiTi archwiresSaliva. Sampling before and during the treatment (6 and 12 months).Ni and Cr. AAS.Before treatment:
Ni: 9.75 µg/L,
Cr: 3.86 µg/L.
After 6 months:
Ni: 10.37 µg/L,
Cr: 4.60 µg/L.
After 12 months:
Ni: 8.32 µg/L,
Cr: 2.04 µg/L.
Statistically insignificant change in Ni concentrations throughout the treatment time. Small but significant decrease in salivary Cr after 1 year.[233] Amini et al. (2012)
40 patients + 50 controlsSS brackets and bands; SS and NiTi archwires.Saliva. Sampling during the orthodontic treatment (ranging from 1 to 32 months after start, 1 sample per patients).Ni and Cr. ICP-MS and ICP-OES.Controls:
Ni: 2.29 µg/L,
Cr: 3.23 µg/L.
Patients:
Ni: 4.19 µg/L,
Cr: 2.83 µg/L.
Statistically significant increase in salivary Ni and decrease in Cr between patients and controls. Non-toxic levels.[234] Talic et al. (2013)
32 patientsSS brackets and bands; SS and NiTi archwires.Saliva. Sampling before and during the treatment (20 days, and 3 and 6 months after starting).Ni and Cr. AAS.Before treatment:
Ni: 5.76 µg/L,
Cr: 2.6 µg/L.
After 20 days:
Ni: 6.54 µg/L,
Cr: 3.68 µg/L.
After 3 months:
Ni: 5.13 µg/L,
Cr: 3.41 µg/L.
After 6 months:
Ni: 5.61 µg/L,
Cr: 3.39 µg/L.
Salivary Ni and Cr concentration did not significantly change.[174] Yassaei et al. (2013)
24 patients + controlsSS brackets and NiTi archwires.Hair. Sampling 16 weeks after starting the orthodontic treatment.Ni. AAS.Before treatment:
Controls Ni: 0.245 µg/g,
Patients Ni: 0.350 µg/g.
After 16 weeks:
Controls Ni: 0.382 µg/g,
Patients Ni: 0.637 µg/g.
Statistically significant differences between controls and patients after 16 weeks.[235] Abtahi et al. (2013)
30 patientsSS brackets and bands; NiTi archwires.Saliva. Before placing the orthodontic treatment, 3 months before introducing stress; and 15 and 30 min after introducing stress.Ni, Cr. AAS.Before stress:
Ni: 11.9–12.4 µg/L,
Cr: 4.1–4.4 µg/L.
After stress:
Ni: 1.6–14.4 µg/L,
Cr: 4.8–5.1 µg/L.
Significant increase in Ni concentrations after stress. No significant alteration in Cr concentrations.[236] Amini et al. (2013)
20 patients + 20 controlsSS brackets, bands, tubes, and ligatures; NiTi and SS archwiresOral mucosa cells. Sampling between 13–15 months after starting the orthodontic treatment.Ti, V, and Zr. ICP-MS.Controls:
Ti: 5.14 ng/g,
Zr: < DL.

Patients:
Ti: 5.23 ng/g,
Zr: 0.54 ng/g.
Non-significant differences in both groups.[237] Martín-Cameán et al. (2014)
20 patients + 20 controlsSS brackets, bands, tubes, and ligatures; NiTi and SS archwires,Oral mucosa cells. Sampling between 13–15 months after starting the orthodontic treatment.Ni, Cr, Co, and Cu. ICP-MS.Controls:
Ni: 4.3 µg/L,
Cr: 2.3 µg/L,
Co: 0.6 µg/L,
Cu: 4.9 µg/L.
Patients:
Ni: 24.8 µg/L,
Cr: 17.5 µg/L,
Co: 11.6 µg/L,
Cu: 8.5 µg/L.
Significantly higher values for all metals for patients when compared with controls.[238] Martín-Cameán et al. (2014)
70 patients + 56 controlsSS brackets, bands, tubes, and ligatures; NiTi and SS archwires.Hair. Minimum duration of the orthodontic treatment of 24 months.Ni, Cr, Fe, Mn, Cu. AAS.Controls:
Ni: 0.36 µg/g,
Cr: 0.36 µg/g,
Fe: 25.3 µg/g,
Cu: 33 µg/g,
Mn: 0.23 µg/g.
Patients:
Ni: 0.33 µg/g,
Cr: 0.33 µg/g,
Fe: 24.86 µg/g,
Cu: 24 µg/g,
Mn: 0.42 µg/g.
Significant increase for Mn concentrations only.[239] Martín-Cameán et al. (2014)
24 pigs (12 controls)SS plates simulating orthodontic appliances.Hair, kidneys, liver, lungs, aorta, and oral mucosa. Up to 6 months.Ni, Cr, Fe, Cd, Co, Cu, Mn, Mo, Si and Zn. ICP-OES.Higher increase in toxic metals registered in the aorta (for Ni), cheek (for Ni) and hair (for Cr).Products of corrosion passed into selected tissues of pigs. Below toxicity levels.[240] Mikulewicz et al. (2014)
47 patientsSS brackets and ligatures; NiTi archwires.Hair. Sampling before and during the orthodontic treatment (4, 8 and 12 months).Ni, Cr and Fe. ICP-OES.Before treatment:
Ni: 0.275 mg/kg
Cr: 0.0201 mg/kg
Fe: 13.2 mg/kg
After 1 year:
Ni: 0.422 mg/kg
Cr: 0.158 mg/kg
Fe: 14.2 mg/kg
Statistically significant increase in Cr content only. Below toxicity levels.[241] Mikulewicz et al. (2015)
30 patients, divided in two groups according to the type of brackets Conventional or MIM SS brackets and tubes; NiTi archwires.Saliva. Immediately before and 60 days after starting the treatment.Ni and Cr. AAS.Before the treatment:
Conventional—Ni: 7.12 µg/L, Cr: 0.25 µg/L.
MIM—Ni: 8.62 µg/L, Cr: 0.42 µg/L.
60 days after treatment:
Conventional—Ni: 12.57 µg/L;,Cr: 0.35 µg/L.
MIM—Ni: 8.86 µg/L, Cr: 0.26 µg/L.
Significant increase in Ni content in each group. Not significant differences between the groups. [242] Amini et al. (2015)
24 patientsSS brackets and bands; NiTi and SS archwires.Hair. Immediately before and 6 months after starting the treatment.Ni and Cr. AAS.Before the treatment:
Ni: 0.1380 µg/g d.h.m,
Cr: 0.1455 µg/g d.h.m.
After 6 months:
Ni: 0.6715 µg/g d.h.m,
Cr: 0.1683 µg/g d.h.m.
Ni and Cr content in hair significantly increased (387 and 16%, respectively).[243] Amini et al. (2015)
13 patientsSS brackets, bands, tubes, lingual sheath, transpalatal arch and archwires; NiTi archwires.Saliva. Sampling before and during the orthodontic treatment (1 week, and 1 and 3 months after placement).Ni and Cr. AAS.Before treatment:
Ni: 1.156 µg/L,
Cr: 11.570 µg/L.
After 1 week:
Ni: 6.841 µg/L,
Cr: 70.386 µg/L.
After 1 month:
Ni: 3.403 µg/L,
Cr: 21.254 µg/L.
After 3 months:
Ni: 3.124 µg/L,
Cr: 20.002 µg/L.
Significant increase in salivary Ni and Cr after starting the orthodontic treatment. Peak concentrations 1 week after placement.[176] Dwivedi et al. (2015)
30 patientsSS brackets and bands; NiTi and SS archwires.Saliva. Sampling before and during treatment (after aligning phase and 10–12 months after placement).Ni and Cr. ICP-MS.Before treatment:
Ni: 48.78 ppb,
Cr: 69.74 ppb.
After aligning stage:
Ni: 59.19 ppb,
Cr: 102.68 ppb.
10–12 months after start:
Ni: 46.33 ppb,
Cr: 87.07 ppb.
Significant increase in salivary Ni and Cr after the initial aligning phase.[182] Nayak et al. (2015)
50 patients with 4–6 months of fixed orthodontic treatmentFixed orthodontic appliances (not specified).Saliva. Sampling after 1 week without using mobile phone (controls); then after 1 week of regular usage (experimental group).Ni. ICP-MS.Experimental group:
Ni: 16.22 ng/L.
Controls:
Ni: 12.84 ng/L.
Statistically significant increase in Ni release for the experimental group when compared with the controls.[244] Saghiri et al. (2015)
1 patientFixed appliances (not specified).Saliva, alveolar bone, and gingiva. Collection during periodontal cosmetic surgery and exostosis removal.Ni. AAS.Ni in saliva: 986.4 ppb.
Ni in bone: 779.5 ppb.
Ni in gingiva: 620.5 ppb.
High Ni accumulation in each sample type.[245] Arcila et al. (2015)
30 patientsSS bands and closed coil springs; SS self-ligating brackets with NiTi clip; NiTi archwire and open springs.Saliva. Sampling before and during the orthodontic treatment (immediately after placing brackets and bands and 2 weeks later; immediately after placing the archwires and 4 and 8 weeks later).Ni. ICP-MS.Before treatment:
Ni: 21.85 µg/L.
Immediately after placing brackets and bands:
Ni: 85.34 µg/L.
Immediately after placing the NiTi archwires:
Ni: 57.74 µg/L.
For the remaining sampling times:
Ni: 13.73–19.83 µg/L.
Significant increase in salivary Ni concentrations after brackets and bands insertion, as well as after placing the archwire. Return to basal levels after 4 weeks. Below dietary intake.[246] Gölz et al. (2016)
24 patientsSS brackets; NiTi archwires.Gingival crevicular fluid. Sampling before and during treatment (1 and 6 months).Ni and Cr. AAS.Before treatment:
Ni: 3.894 µg/g,
Cr: 1.978 µg/g.
1 month after starting:
Ni: 5.913 µg/g,
Cr: 4.135 µg/g.
6 months after starting:
Ni: 19.810 µg/g,
Cr: 13.760 µg/g.
Significant increase in Ni and Cr (up to 510 and 700%, respectively) during the treatment, as well as gingival inflammation promotion.[247] Amini et al. (2016)
42 patients, divided according to the bracket typeMIM tubes and SS brackets; NiTi, Cu-NiTi, or epoxy-coated NiTi archwire.Saliva. Sampling before and 2 months after starting the orthodontic treatment.Ni. AAS.Before treatment:
Ni: 10.4571 µg/L.
After 2 months:
Ni: 11.0799 µg/L.
Statistically significant increase in salivary Ni concentration but depends on the archwire type.[173] Masjedi et al. (2016)
10 patientsSS brackets and ligatures; bands with tubes and lingual sheaths (not specified); NiTi archwires.Saliva. Sampling before and during treatment (10 days and 1 month after placement).Ni and Cr. ICP-OES.Before treatment:
Ni: 0.0039 mg/L,
Cr: 0.0024 mg/L.
After 10 days:
Ni: 0.0288 mg/L,
Cr: 0.0037 mg/L.
After 30 days:
Ni: 0.0370 mg/L,
Cr: 0.0103 mg/L.
Statistically significant increase in salivary Ni after 10 and 30 days when compared with controls, as well as for salivary Cr between the 10th and the 30th day.[248] Kumar et al. (2016)
47 patients with different dietary habits (coffee, yoghurt, juice and vinegar consumption)Fixed orthodontic appliances (not specified).Human hair. Sampling at the beginning of the treatment, and after 4, 8 and 12 months.Ni and Cr. ICP-OES.At the beginning:
Ni: 0.131–0.331 mg/kg,
Cr: 0.00578–0.0338 mg/kg.
After 4 months:
Ni: 0.222–0.505 mg/kg,
Cr: 0.0620–0.446 mg/kg.
After 8 months:
Ni: 0.252–0.444 mg/kg,
Cr: 0.0862–0.292 mg/kg.
After 12 months:
Ni: 0.207–0.500 mg/kg,
Cr: 0.124–0.191 mg/kg.
Consuming foods and drinks with low pH can intensify metal release of Cr and Ni during the orthodontic treatment.[249] Wołowiec et al. (2017)
30 patientsFixed orthodontic appliances (not specified).Saliva, biofilm, and oral mucosa cells. Before and during orthodontic treatment (1 week and 6 months).Ni. AAS.Before treatment:
Saliva—Ni: 2.213 ppm,
Biofilm—Ni: 4.943 ppm,
Oral mucosa—Ni: 3.327 ppm.
After 1 week:
Saliva—Ni: 2.627 ppm,
Biofilm—Ni: 5.75 ppm,
Oral mucosa—Ni: 3.683 ppm.
After 6 months:
Saliva—Ni: 3.03 ppm,
Biofilm—Ni: 6.917 ppm,
Oral mucosa—Ni: 3.143 ppm.
Significant increase in Ni levels, especially in biofilm samples.[250] Causado-Vitola et al. (2017)
46 patients, divided according to the bracket typeSS MIM or conventional brackets and tubes; NiTi and SS archwires.Hair. Sampling before and 6 months after starting the orthodontic treatment.Ni and Cr. AAS.Before treatment:
Ni: 0.1600 µg/g d.h.m,
Cr: 0.1657 µg/g d.h.m.
After 6 months:
Ni: 0.3199 µg/g d.h.m,
Cr: 0.3066 µg/g d.h.m.
Statistically significant Ni and Cr content increase, regardless the bracket type.[251] Masjedi et al. (2017)
37 patientsMetallic Fixed appliances (not specified).Saliva. Sampling immediately before placing the fixed appliances, and after 1 and 24 weeks.Ni. AAS. Increased salivary Ni concentration, probably responsible for modifying the oxidative/antioxidative balance of saliva.[252] Buckzo et al. (2017)
60 patients + 30 controlsConventional appliances: SS brackets and bands, and NiTi archwires; aesthetic appliances: polycarbonate brackets and tubes and Rh-coated NiTi archwires.Saliva. Sampling from patients undergoing an orthodontic treatment for 1–6 months. One collection per patient.Ni, Cr, Fe, and Cu. Total reflection XRF.Controls:
Ni: 4.14 µg/L,
Cr: 10.32 µg/L,
Fe: 32.04,
Cu: 11.40.
Patients with conventional appliances:
Ni: 22.20 µg/L,
Cr: 89.45 µg/L,
Fe: 517.77 µg/L,
Cu: 15.10 µg/L.
No significant differences regarding Ni and Cr concentrations between conventional or aesthetical and control groups. Ni and Cr influenced by the type of appliances. No differences in Fe and Cu between groups.[253,253] Lages et al. (2017)
42 patients, divided in two groups (mobile phone users and nun-users)Fixed orthodontic appliances (not specified).Saliva. 6–9 months after placement.Ni. ICP-OES.Mobile phone users:
Ni: 0.012 ppb.
Non-users:
Ni: 0.0083 ppb.
Mobile phone radiations can influence Ni ion release, but a statistically non-significant difference was obtained.[254] Nanjannawar et al. (2017)
30 patientsSS brackets and NiTi archwireGingival crevicular fluid. Sampling before and during treatment (1 and 6 month)Ni and Cr. AASBefore treatment:
Ni: 3.2 µg/g
Cr: 4.1 µg/g
1 month after starting:
Ni: 4.5 µg/g
Cr: 4.9 µg/g
6 months after starting:
Ni: 14.2 µg/g
Cr: 21.4 µg/g
Significant increase in Ni and Cr concentrations during the orthodontic treatment.[255] Bhasin et al. (2017)
24 patients + 28 controls Scalp hair. Sampling before and 1 year after placement.Ni and Cr. AAS.Before:
Controls—Ni: 0.085 µg/g, Cr: 0.299 µg/g.
Patients—Ni: 0.061 µg/g, Cr: 0.304 µg/g.
1 year after:
Controls—Ni: 0.086 µg/g, Cr: 0.258 µg/g.
Patients—Ni: 0.149 µg/g, Cr: 0.339 µg/g.
Statistically significant differences in both Ni and Cr contents in scalp hair between the groups after 1 year of orthodontic treatment.[256] Jamshidi et al. (2018)
42 patients with metallic brackets + 42 with ceramic bracketsSS or ceramic brackets; NiTi archwires.Saliva. Sampling before and 6 months after starting the orthodontic treatment.Ni, Cr, Ti, Co, Cu, and Zn. ICP-MS.Metallic brackets:
Before:
Ni: 4.24 µg/L,
Cr: 1.95 µg/L,
Ti: 1.68 µg/L,
Co: 0.46 µg/L,
Cu: 23.31; µg/L,
Zn: 220.67 µg/L.
After 6 months:
Ni: 5.04 µg/L,
Cr: 1.01 µg/L,
Ti: 9.29 µg/L,
Co: 0.32 µg/L,
Zn: 168.45 µg/L.
Statistically significant increase in salivary Ti, and statistically significant decrease in Cr and Zn. Non-significant differences difference between both groups.[257] Jurela et al. (2018)
20 healthy patients + 20 periodontal patientsBrackets (not specified) and NiTi archwires.Saliva. Sampling 2 months after starting the orthodontic treatment.Ni. ICP-MS.Healthy patients:
Ni: 182.8 ng/mL,
Cr: 6.35 ng/mL.
Periodontal patients:
Ni: 338.2 ng/mL,
Cr: 7.45 ng/mL.
Statistically significant differences in salivary Ni concentrations between healthy and periodontal patients. Inconclusive for Cr.[258] Amini et al. (2019)
60 patients divided into 2 groups according to the oral health products used (fluorinated vs. nonfluorinated)SS brackets and NiTi archwires.Gingival crevicular fluid. Sampling before and after 1 week, 1 month and 6 months after placement.Ni, Cr, Ti, and Mn. Nonfluorinated:
Before—Ni: 0.49 µg/L; Ti: 0.49 µg/L.
7 days—Ni: 0.52 µg/L; Ti: 0.51 µg/L.
30 days—Ni: 13.42 µg/L; Ti: 40.09 µg/L; Cr: 0.50 µg/L; Mn: 0.50 µg/L.
6 months—Ni: 0.51 µg/L; Ti: 4.80 µg/L; Cr: 0.49 µg/L; Mn: 0.49 µg/L.
Fluorinated:
Before—Ni: 0.51 µg/L; Ti: 0.51 µg/L.
7 days—Ni: 0.52 µg/L; Ti: 0.50 µg/L.
30 days: Ni: 101.78 µg/L; Ti: 64.69 µg/L; Cr: 12.00 µg/L.
6 months—Ni: 0.51 µg/L; Ti: 0.51 µg/L; Cr: 0.53 µg/L; Mn: 0.48 µg/L.
Statistically significant increase in Ni, Cr, and Ti concentrations at 30 days only. Higher metal content in patients using fluoride-containing oral hygiene products, with statistically significant difference for Ni when compared with patients using nonfluorinated products.[259] Chitra et al. (2019)
50 patients + 30 controlsSS brackets and bands; SS and NiTi archwires.Saliva and serum. Sampling before and during the orthodontic treatment (1 week, 3 months, 1 year, and 1.5 years).Ni, Cr and Zn. AAS.Saliva—before treatment:
Controls—Ni: 4.33 ppb, Cr: 1.13 ppb, Zn: 10.73 ppb;
Patients—Ni: 4.24 ppb, Cr: 1.18 ppb, Zn: 11.8 ppb.
Saliva—1.5 years after:
Controls—Ni: 5.02 ppb, Cr: 1.27 ppb, Zn: 10.24 ppb;
Patients—Ni: 67 ppb, Cr: 30.8 ppb, Zn: 164.7 ppb.
Serum—before treatment:
Controls—Ni: 8.31 ppb, Cr: 6.18 ppb, Zn: 29.1 ppb;
Patients—Ni: 8.46 ppb, Cr: 6.46 ppb, Zn: 28.3 ppb;
Serum—1.5 years after:
Controls—Ni: 8.47 ppb, Cr: 6.02 ppb, Zn: 30.1 ppb;
Patients—Ni: 81.65 ppb, Cr: 35.6 ppb, Zn: 597.16 ppb.
Statistically significant increase in salivary and serum concentrations of Ni, Cr, and Zn between controls and patients. Below toxic levels.[260] Quadras et al. (2019)
100 patients + 40 controlsFixed orthodontic appliances (not specified).Serum. Sampling between 3 weeks and over 18 months.Ni, Cr, Fe, Cu, Mn, and Zn. ICP-MS.Controls:
Ni: 26.95 µg/L,
Cr: 44.45 µg/L,
Fe: 200.72 µg/L,
Cu: 31.43 µg/L,
Mn: 13.75 µg/L,
Zn: 32.90 µg/L.
Patients:
Ni: 61.40 µg/L,
Cr: 44.28 µg/L,
Fe: 454.92 µg/L,
Cu: 55.42 µg/L,
Mn: 18.85 µg/L,
Zn: 143.70 µg/L.
All ions’ concentrations increased in the serum, except for Cr. Ni concentration in serum was dependent on treatment time.[261] Moghadam et al. (2019)
35 patientsFixed orthodontic appliances (not specified) involving NiTi and SS archwires.Saliva and urine. Sampling before and during the orthodontic treatment (3 and 6 months).Ni and Ti. ICP-OES. Statistically significant differences in the Ni concentrations in saliva between 3 and 6 months, as well as Ti in urine in the same periods.[198] Velasco-Ibañez et al. (2020)
20 patients + 20 controlsSS brackets, bands and archwires.Saliva. Sampling before and 6–12 months after placement.Ni. AAS.Controls:
Before: 9.82 ng/mL,
After: 10.21 ng/mL.
Patients:
Before: 9.90 ng/mL,
After: 15.83 ng/mL.
Higher but statistically insignificant increase in Ni concentrations in patients when compared with the control group.[262] Butt et al. (2020)
43 patients divided in groups according to soldering of the lingual arch + 21 controlsLingual arches composed of SS bands and wires and soldered or welded SS wires.Saliva. Sampling before and after placement (7, 15 and 30 days).Ni, Cr, Fe, Cu, Zn, Ag, Cd, and Sn. ICP-MS.Controls:
Ni: 8.0–6.0 µg/L,
Cr: 3.5–3.9 µg/L,
Fe: 227.9–289.9 µg/L,
Cu: 23.1–34.2 µg/L,
Zn: 461.0–499.8 µg/L,
Ag: 10.1–18.3 µg/L,
Cd: 0.6–1.0 µg/L,
Sn: 16.6–25.3 µg/L.
With lingual arches:
Ni: 5.3–34.5 µg/L,
Cr: 3.3–4.2 µg/L,
Fe: 201.0–314.8 µg/L,
Cu: 28.0–40.7 µg/L,
Zn: 384.3–963.4 µg/L,
Ag: 3.7–20.8 µg/L,
Cd: 0.5–1.5 µg/L,
Sn: 11.4–27.7 µg/L.
No statistically significant differences for most metallic ions. Below toxic levels.[263] Schacher et al. (2020)
40 patients divided into 2 groups according to the toothpaste used (nonfluorinated or fluorinated)Brackets (non-specified) and NiTi archwires.Gingival crevicular fluid. Before and after placement (7 and 30 days, and 6 months).Ni and Cr. ICP-MS.Nonfluorinated toothpaste [ng/mL]:
Before: Ni: 0.49, Cr: 0.48;
7 days: Ni: 0.52, Cr: 0.52;
30 days: Ni: 13.4, Cr: 40.6;
60 days: Ni: 0.54, Cr: 4.9;
Fluorinated toothpaste [ng/mL]:
Before: Ni: 0.52, Cr: 0.52;
7 days: Ni: 0.54, Cr: 0.53;
30 days: Ni: 100.2, Cr: 62.4;
60 days: Ni: 0.52, Cr: 0.52.
Statistically significant increase in Ni and Cr release for patients with prescribed fluorinated toothpastes.[264] Pritam et al. (2021)
12 patients with fixed appliances + 15 patients with removal appliancesFixed appliances (NiCr brackets and NiTi archwires); removal aligners (polyurethane).Saliva. Sampling before and 3 months after starting the treatment.From Al (Z = 13) to Y (Z = 39). Total reflection XRF. No significant alterations regarding metals from metal corrosion and inflammatory reactions in patients under dental plaque control.[265] Zeffa et al. (2021)
20 patientsSS brackets (other components were not specified).Saliva. Before and during treatment (2 weeks, and 1, 4, and 6 months after start).Ni, Cr, Fe, Ti, and Cu. ICP-OES.Before treatment:
Ni: 3.94 µg/L, Cr:2.37 µg/L, Fe: 45.13 µg/L, Ti: 48.25 µg/L, Cu: 1.53 µg/L.
During treatment:
Ni: 9.73–34.22 µg/L, Cr: 8.20–17.70 µg/L, Fe: 56.71–99.96 µg/L, Ti: 42.29–62.53 µg/L, Cu: 8.12–25.31 µg/L.
Maximum ion concentrations obtained 1 month after starting the orthodontic treatment. Increase in saliva pH and flow rate. Kinetic model proposed.[266] Hamadamin (2022)
17 patientsSS brackets, bands, and tubes; NiTi archwires.Saliva. Sampling before and during treatment (2 days, and 1, 4, and 12 weeks).Ni, Cr, and Fe. ICP-OES.Ni: 132–175 µg/L,
Cr: 171–192 µg/L,
Fe: 826–1023 µg/L.
No statistically significant variations registered throughout the study time.[116] Fróis et al. (2022)
60 patientsSS brackets; metallic archwires.Saliva. Sampling 1,3 and 5 years after starting the treatment.Ni. ICP-MS.1 week after starting:
Ni: 1.25–1.74 ppb;
2 weeks after starting:
Ni: 5.76–6.07 ppb;
3 weeks after starting:
Ni: 4.32–4.78 ppb.
Significant increase between the 1st and the 2nd sampling time, reportedly linked to the use of hand-held mobile phones.[267] Rajendran et al. (2023)
Friction always opposes movement between two sliding surfaces under load and may be divided into two modes: Static—when the applied force is still insufficient to induce relative motion—and dynamic—when surfaces are in relative motion (Figure 12). Despite the focus on kinetic friction, the motion of an archwire is hardly ever continuous; therefore, the most likely relevant friction type in orthodontic tooth movement is static friction, opposing any applied force [37,215]. Factors affecting both friction modes include: Geometry and type of archwires, brackets, and ligatures; surface chemical composition and roughness; cleanliness; and lubrication conditions [12,37,68,216,219,268,269,270]. Moreover, the manipulation of the components when placing and adjusting the orthodontic appliances during the treatment may lead to plastic deformation, wear tracks, and debris. Figure 13 points out these features in a retrieved orthodontic tube after two years of intraoral use.
Saliva is the natural intraoral lubricant by forming a protective pellicle [68,69,70,71]—a double layer of proteins [73]—on any material surface and, therefore, reducing the dynamic coefficient of friction [74,271]. While biofilms might have a protective role [66,90,272], microbiological activity most likely contributes to surface degradation by inducing corrosive microcells, rupturing the biofilm, and roughening the appliance surfaces [73]. These effects increase friction, wear, and metallic ions released from the bracket/wire contact pair.

3.3. Oral Hygiene with Fluoride-Based Products

Functional and aesthetic success is essential in orthodontics, but patients must comply with proper oral hygiene during treatment to avoid tooth demineralization and white spot lesions [91,108,273]. Fixed appliances make this task difficult, as the number of oral bacteria related to gingivitis increases shortly after their oral placement [273]. In fact, dental plaque accumulates in several regions (see Figure 7, Figure 9 and Figure 13), namely in the gingival areas or behind the archwires (e.g., on the bracket slots) [28,106,116,137,274]. To fight dental plaque, orthodontists prescribe fluoride-containing toothpastes, mouth rinses, gels, and varnishes to further control its accumulation and growth, enhance enamel integrity, and prevent dental and gingival diseases [72,91,275].
The downside of using these fluorides is the increased corrosion susceptibility of metallic alloys [29,42,264,276,277,278,279,280,281,282]. Fluoride ions (F)—combined with mechanical brushing—easily degrade the protective oxide layers of both SS and Ti-based alloys (see Equations (8) and (9)), increasing localized and general corrosion, promoting metallic ions release [276,283], and negatively impacting their mechanical and surface properties [41,42,277,279,282,284,285,286]—especially at low pH [104,118] and under the simultaneous presence of chloride ions [287].
Walker et al. [41] reported reduced unloading mechanical properties of SS and β -Ti archwires when exposed to neutral or acidulated prophylactic fluoride gels, which may prolong the orthodontic treatment time. On the other hand, Sufarnap and colleagues [56] reported an increase in both surface roughness and Ni and Cu release from Cu–NiTi archwires in NaF solution in vitro without a significant change in the deflection force.
Corrosion of SS bands and brackets [117,118,288,289] also increases in the presence of fluoride ions. Chantarawaratit and Yanisarapan [289] argued that acidulated phosphate fluoride gel should not be used in patients wearing fixed metal-based orthodontic appliances.
Since these SS components are the support for NiTi or Cu-NiTi archwires, galvanic coupling risk increased during the 1st orthodontic treatment phase (leveling/aligning) with possible mechanical and/or biocompatibility-related adverse consequences [108,112,284]. Figure 14 evidences the risk of galvanic coupling between dissimilar bracket/archwire pairs (concerning their alloys’ composition), showing different in vitro metallic ion release profiles, according to the three solutions used. Synthetic F-free Fusayama–Meyer saliva—widely used in in vitro studies—showed lower aggressiveness than the commercial mouthwashes containing fluoride [108].

4. Orthodontic Alloys Modification

The main topic of this chapter is to briefly point out the current mitigation strategies to reduce the metal corrosion susceptibility of orthodontic appliances, including existing and emerging approaches for surface and bulk alloy modifications.

4.1. Non-Metallic Components

One way to reduce the release of metallic ions due to intraoral corrosion is to replace metallic-based alloys with non-metallic materials. Several examples of Ni-free substitutes for common orthodontic devices and components are listed in Table 7. However, existing or under development non-metallic alternatives for orthodontic appliances are more focused on satisfying patients’ increasing aesthetic demands rather than on clinical concerns. Polymeric brackets—made of polycarbonate (PC), polyurethane (PU), or polyoxymethylene (POM); ceramic brackets—as those made of mono or polycrystalline aluminum oxide (Al2O3); non-metallic brackets with and without metallic slots [10,11,253,290,291,292,293]; and non-metallic archwires—produced from thermoplastic polymer [294], polyphenylene (PP) [295], poly-ether-ether-ketone (PEEK) resin [296], and fiber-reinforced polymers [34,297,298]—are some examples. A systematic literature review on transparent orthodontic archwires was prepared by Mikulewicz and colleagues [299], while J. Russel [290] reviewed aesthetic orthodontic brackets. Table 8 summarizes some of the current commercially available non-metallic aesthetic brackets and archwires.
Multiple concerns [11,300] still limit clinicians’ acceptance and widespread use of alternative aesthetic solutions [290,301]. For instance, polymeric-based brackets may experience hardness reduction [291], undergo discoloration and staining from food dyes and sterilization procedures [302,303], release highly cytotoxic substances (e.g., bisphenol A [BPA]) [304], or have difficulty sliding against metallic archwires [11]. Similarly, ceramic brackets may also stain and discolor [305] during orthodontic treatment, in addition to their brittleness and high susceptibility to fracture [290,306,307], increased friction [308] and more severe archwire notching [218], higher pain during treatment [309], and enamel damage [11]. In a recent in vitro study concerning bracket genotoxicity [194], the authors even argue that ceramic brackets are more genotoxic than conventional SS or CoCr alloy ones. Non-metallic, aesthetical archwires present limitations as well. The foremost drawbacks—crazing formation, plastic deformation, and lower deliverable forces when bended [297,310] in addition to noticeable color change [311]—have limited their clinical widespread use.
Table 8. Non-metallic available aesthetic brackets and archwires [10,11,253,290,293,297,298,299,310].
Table 8. Non-metallic available aesthetic brackets and archwires [10,11,253,290,293,297,298,299,310].
BracketsArchwires
PolymericPolyurethane (PU).
Polycarbonate (PC).
Polyoxymethylene (POM).
Fiberglass-reinforced PC.
Polyethylene terephthalate (PET) polytetrafluoroethylene (PTFE).
Polyetheretherketone (PEEK).
Fiber-reinforced polymer composites, and 3-layered wire (SiO2/silicone resin/nylon).
CeramicAlumina.
Zirconia.

4.2. Nominal Composition Optimization

Another approach to decreasing intraoral corrosion susceptibility is to adjust the bulk nominal composition of the alloys by adding or subtracting certain elements [95,312]. As already mentioned in Section 2, Ni stabilizes the austenitic phase of SS and increases the alloy’s resistance to corrosion and oxidation [10], namely against non-oxidizing acids [127]. However, the abovementioned clinical concerns make these alloys undesirable. Ni-“free” SS alloys (maximum of 2% Ni) were pointed out as viable alternatives to allergic orthodontic patients to this transition element [313]. The main development involves, at least, the presence of another element, such as Mn [10], in the SS nominal composition. Nevertheless, the final microstructural, mechanical, and corrosion resistance properties of the resulting steel can be compromised [36]. Additionally, Ni-“free” brackets/tubes are more cytotoxic than Ti-based ones [312] and may still provoke oral allergic reactions [314], even though conventional SS alloys are less biocompatible [304]. Another alternative approach is appropriately described by Eliades et al. [28] concerning the replacement of the most commonly used bracket SS alloy—316L AISI—by the duplex SS 2205 grade, which contains less Ni nominal content (Table 3).
To mitigate the nickel impact on oral health, the high Ni-containing alloys (TiNi) have been chemically optimized by adding Ag or Cu. The ternary TiNiAg alloys show similar mechanical properties and cytotoxicity effects as the binary TiNi ones and are more effective in reducing bacterial adhesion [315], but up to a certain Ag content limit [316] (reportedly 0.5–1.5 wt.% by Oh et al. [317]). Newer Cu–NiTi archwires [34] provide thermal activation and increase their resistance to permanent deformation. However, such alloys cannot be easily welded, are expensive, have low formability and high friction against SS brackets, and may fracture when bent over sharp edges [10,42,53]. New composite archwires (CAW), produced by NiTi and SS segments laser-welded with Cu, also seem to be promising for orthodontic treatments due to their excellent mechanical properties. Nevertheless, the resulting Cu interlayer could corrode in in vivo conditions, leading to biocompatibility and performance complications [318].

4.3. Manufacturing Processes

Different manufacturing processes lead to different final microstructural and physico-mechanical properties that can influence the overall corrosion resistance of metallic appliances. Brackets, tubes, and bands, for instance, can be made by casting and/or machining (milling) separate metallic pieces (e.g., the base and the wings, as shown in Figure 15). These parts are joined by soldering or welding [319]. Ag-based brazing alloys were popular and are still widely used [320], but these solders may dissolve or induce galvanic coupling inside the oral cavity, releasing cytotoxic ions [111,319,321,322]. Alternatives include the use of Au-brazing alloys, which induce galvanic corrosion of the less-noble SS following the Ag case, or Ni-based brazing alloys, with no available data [98,319] but undesirable due to their toxicity. Laser welding, a more recent and preferable advanced manufacturing technique, allows joining without brazing alloys [95,192,322], with lower corrosion susceptibility [323], even though large gaps in the joints may be a disadvantage for Ti-based brackets [319].
Metal injection molding (MIM) is nowadays a popular manufacturing technique [11,319] to provide single-piece metallic orthodontic components with superior surface finishing in a cost-effective manner. Advantages over conventional casting/welding brackets include the absence of galvanic coupling between the base and the wings, the reduction of oral metallic ion release, and lower wear when used with NiTi archwires contacts [98,242,319]. However, internal porosity and possibly different biocompatibility and electrochemical behaviors of well-known alloys are some concerns that motivate further optimization [98,242,319].
Important developments in orthodontic appliance manufacturing include the production of functionally graded archwires (that apply different triggering forces depending on the location, accomplished by heat treatments) [324] and customized pieces by additive manufacturing [325,326,327], which may lead to less prolonged and more effective orthodontic treatments.

4.4. Surface Modification and Coatings

Another important optimization route for orthodontic alloys is surface modification and/or coatings deposition, preserving their best features, that is, their bulk properties. Corrosion most likely develops on bearing surfaces (Section 2) rather than in the bulk. Hence, different surface finishes can greatly impact the alloys’ corrosion performance [95,328]. Mechanical polishing, for example, can reduce surface defects, leaving a smooth topography [329]. According to Hunt et al. [96]—who polished metallic archwires without impacting surface hardness—this process seemed detrimental or indifferent to CoCr, SS, and β -Ti archwires exposed to a corrosive medium (0.9% NaCl solution), but polished NiTi archwires presented higher corrosion resistance than its pristine condition [96]. Electropolishing is a common treatment to enhance both the appearance and corrosion resistance of the components, but galvanic corrosion cells may be produced between polished and unpolished regions [95]. Ion implantation is also a commercially used surface treatment of NiTi archwires [329,330] to reduce friction when compared with uncoated ones. However, Wichelhaus et al. [329] showed that this effect only occurred at the beginning of the orthodontic treatment, alerting them to the risks of recycling these components.
Current progress is also devoted to thin film/coating deposition with new and unique properties—a versatile approach to improving the final surface properties [39]. Within this regard, the flowchart presented in Figure 16 condenses the biomaterial surface properties that can be adjusted by plasma synthesis, as proposed by Chu et al. [331]. Without error, it can be generalized to a variety of methods and technologies available for biomedical applications, including orthodontics.
Due to the high importance of coatings’ deposition in tailoring the surface of orthodontic bioalloys, the authors opted to present and develop this topic separately in the following chapter (Section 5).

5. Protective Coatings in Orthodontics

This chapter aims to present the multitude of coatings available in the literature that have been deposited onto real fixed orthodontic appliances (the widely used archwires, brackets, and bands) to improve features such as corrosion, friction, fretting, biocompatibility, and antibacterial activity. For that, the results of the current review will be presented considering the three main groups: metal-, polymeric- and ceramic-based coatings (Table 9, Table 10 and Table 11). Several industrial deposition techniques can be used, as reviewed by Arango et al. [39] and presented in Figure 17, some of them which are applied in orthodontics.

5.1. Metal-Based Coatings

Due to their higher corrosion resistance in addition to their aesthetic and pleasant color, metallic-based coatings have been deposited onto fixed orthodontic appliances, as presented in Table 9. Most pure metal coatings are devoted to rhodium (Rh), gold (Au), and silver (Ag). Titanium (Ti) and zinc (Zn) films were also included in this summary due to their excellent biocompatibility.

5.1.1. Transition Metal-Based Coatings

  • Rhodium (Rh):
Rhodium-coated archwires—usually as bimetallic Rh–Au coatings [332,333,334]—are available on the market due to their pleasant color and superior corrosion resistance, without compromising the mechanical properties of the bulk alloys [253,310,335,336,337,338].
Katić and coauthors [332,339,340,341] compared the corrosion resistance of the uncoated NiTi archwires with those nitrified and coated with Rh. Surprisingly, electrochemical tests showed that corrosion susceptibility increased in artificial saliva, being highest for Rh-coated archwires: The heterogeneous layers induced galvanic coupling between the noble metals (Au and Rh) and the NiTi alloy, causing pitting corrosion and pronounced Ni release, as well as aesthetical appearance degradation [339,340]. Other researchers compared different types of coated and uncoated archwires, finding the worst corrosion resistance for the Rh-coated ones [61,342]. However, when fluoride-based prophylactic agents were weakly added during a 4-week immersion test in artificial saliva, ionic release from Rh-coated archwires decreased with increasing F ionsconcentration [332].
Nsaif et al. [343] found a protective effect of Rh-based coatings against an acidulated phosphate fluoride agent without compromising the overall mechanical properties of the archwires. Mlinaric et al. [279] demonstrated that the alteration of the properties of Rh-coated NiTi archwires depends on the type of oral antiseptics used. In turn, Osmani and colleagues [338] advised that such archwires should be avoided if oral hygiene is compromised due to the higher release of metallic ions.
The use of probiotic supplements—recommended to prevent caries and gingivitis during the fixed orthodontic treatment—also negatively impacts the corrosion resistance of Rh-coated NiTi archwires in artificial saliva: Both general and localized (pitting) corrosion increased, with precipitation of corrosion products on the surfaces [344,345]. However, a simple 1-month immersion test of Rh-coated SS and NiTi archwires in probiotic-containing artificial saliva suggested a decrease in corrosion susceptibility, possibly due to a benign anticorrosion effect of the early-formed biofilms [346]. Since mature biofilms will more likely promote MIC, the in situ higher accumulation of microorganisms for Rh-coated NiTi archwires reported by Lima et al. [347] may be a demerit result. Other researchers disagree [63], reporting that: (i) commercial Rh-coated NiTi archwires can significantly decrease S. mutans bacteria adhesion in vitro by altering the apparent surface free energy; and (ii) Au coating on SS archwires did not significantly influence the adhesion of S. mutans due to similar apparent surface free energy as uncoated SS [63].
Usui et al. [348] compared the kinetic and static frictional properties of both uncoated and Rh-coated SS archwires, under dry and wet conditions, against a ceramic bracket. In all tests, coated wires generated significantly higher kinetic and static frictional forces than the uncoated ones—more than 46% higher. Still, three-point bending test results indicated equivalent flexural properties. The bending effect can, however, increase the surface roughness of NiTi archwires coated with Rh-based materials [333]. Albawardi and colleagues [349] concluded that Rh-coated β-Ti archwires are not ideal to decrease friction against brackets for angulations up to 10° when compared to uncoated β-Ti or SS wires.
The color stability of commercial Rh-coated NiTi archwires has also been investigated by immersion in a coffee staining solution. While Alsanea and Shehri [335] found just a slight discoloration after 4 weeks, Ramasamy and colleagues [350] reported an extremely marked color change after 3 weeks of wetting.
  • Gold (Au) and Platinum (Pt):
Gold and platinum are two noble metals that are widely used in many medical applications due to their bioinertness and notable corrosion resistance [351,352]. Despite their high cost, Au coatings are commercially available in orthodontics [62,64]. Surprisingly, Toy et al. [62] found that Au-plated SS brackets “remarkably decreased cell proliferation of HGFs” (human gingival fibroblasts), possibly related to an increase in in vitro corrosion during the test caused by surface defects of the substrate alloy. Conversely, other tested brackets (uncoated, made of metallic, ceramic, and composite materials) increased cell proliferation [62]. In other research work, Ito et al. [61] evaluated the coating loss of Au-coated SS wires after acid immersion (35% hydrochloric acid, pH = 1.1), wire ending, and mechanical brushing. While the Au coating resisted mechanical brushing and effectively prevented the release of metallic ions during immersion, wire bending showed signs of cracks and delamination of the coated wire [61]. Krishnan et al. [64] studied several coated and uncoated NiTi archwires and found that the breakdown potential of Au-coated ones is only surpassed by PTFE-coated wires. Regarding Pt-based coatings, Khonsari et al. [353] deposited Pt nanoparticles (NPs) on NiTi archwires by an electrochemical method, achieving adherent and homogeneous layers with a suitable tooth-like color.
  • Silver (Ag):
Silver is recognized as one of the oldest antibacterial agents known, dating back to ancient times, at least by ~1500 B.C. [354,355]. In orthodontics, Ag-based coatings have been used to take advantage of this feature, that is, to reduce both bacterial adhesion and biofilm growth. Hence, periodontal diseases, tooth decay, and demineralization [356] have been prevented during orthodontic treatments.
Mhaske et al. [357] deposited pure Ag onto two types of archwires (NiTi and SS) by thermal vacuum evaporation. In addition to its antimicrobial effect, the 10 nm thick film effectively reduced the in vitro adhesion of L. acidophilus on both wire types, and, therefore, the total weight increase due to biofilm accumulation was lower: 35.5 and 20.5% for uncoated SS and NiTi wires, respectively, and less than 4.5% for Ag-coated wires [357]. Ghasemi et al. [358] reported that 60 and 100 µm thick Ag films deposited by physical vapor deposition (PVD) onto SS brackets were able to reduce the S. mutans counts right after 3 h of incubation.
Recently, an in situ study analyzed the initial biofilm formation (up to 48 h) onto an austenitic Ni-free bracket coated with Ag by using 3 distinct production processes: PVD, galvanic, and plasma immersion ion implantation and deposition (PIIID) processes [359]. The results showed statistically significant lower biofilm volume and surface coverage for all thin-film types in comparison to uncoated samples, but not between modified samples. All coatings showed antibacterial properties, with the PIIID-modified bracket displaying the highest increase in the dead/alive cell proportion.
An in vivo study lasting up to 75 days was conducted by Metin-Gürsoy et al. [360] to evaluate the antibacterial properties and the ion release from nano-Ag-coated SS brackets. The PVD-coated brackets placed on the mandibular incisors of Wistar Albino rats effectively inhibited S. mutans growth, which reduced caries development on the smooth surfaces. Concerns regarding the biocompatibility of Ag-based coatings for orthodontic applications previously motivated the same authors [356] to perform a 60-day animal histological experiment. Encouraging findings indicated that uncoated and PVD-nano-Ag-coated standard brackets implanted in healthy rats have similar biocompatibility behaviors. Nevertheless, the authors found brown-black silver granules in the adjacent tissues. Further studies are required to assess an eventual negative biological impact.
The research work conducted by Espinosa-Cristóbal et al. [361] described the influence of silver nanoparticles’ (Ag-NPs) size on the activity of S. mutans. The coatings were chemically deposited on both SS brackets and archwires. The obtained results showed that coatings possessed good antimicrobial and anti-adhesion properties against this bacterial strain, especially for smaller-sized Ag-NPs (diameter of ~8.1 nm). Similarly, other authors [362] reported a significant decrease in the activity of S. mutans and L. acidophilus after a 24–48 h incubation with Ag-NPs-coated segments of SS orthodontic bands. Ag-NP coatings provide antimicrobial properties to SS brackets [48], bands [363,364] and archwires [365], namely against S. mutans, S. sobrinus, C. albicans, and L. acidophilus [48,364,365]. Coated NiTi archwires induced a more than 90% decrease in the adhesion of S. sanguinis and L. salivaris [366]. Zeidan et al. [367] reported an Ag-NPs coating (deposited on SS brackets by the thermal evaporation method) with an antibacterial effect against S. mutans and L. acidophilus over a period of 3 months of incubation. According to an interesting work by Anand et al. [368]—who synthesized Ag-NPs from plant extracts—the well-known antibacterial effect of Ag-NP coatings can be further improved by simultaneously adding TiO2-NPs to functionalize orthodontic archwires. These coatings “showed an enhanced antibacterial effect against Multi-Drug Resistant (MDR) bacteria” [368].
Since corrosion and abrasion impact the intraoral performance of coatings, Ryu et al. [369] added Pt (up to 7%) to Ag coatings by PVD. The goal was to increase chemical stability. From the potentiodynamic polarization tests, the resulting AgPt-coated SS substrates showed higher corrosion resistance than the single Ag-coated or uncoated samples. In addition, all coatings diminished S. mutans and A. actinomycetemcomitans growth by up to 60%. Simultaneously, the cytotoxicity due to silver ion release was statistically insignificant when tested with human gingival fibroblasts. The authors [369] claim that brackets and archwires can be coated with these binary Ag–Pt materials, even though only flat substrates were used in their studies.
Whereas the (micro)biological impact is very important in the oral cavity, the Ag-based coatings must not compromise the surface tribological properties, especially the frictional behavior. Thermal vacuum evaporation-synthesized Ag coatings (~10 nm thick) onto two types of SS archwires did not affect their frictional resistance against SS brackets [370]. Reportedly, friction in the SS bracket/archwire combination decreases when at least the wire is Ag-coated [371]. Nevertheless, thick Ag-electroplated SS brackets (8–10 µm) significantly increased both surface roughness and friction forces against SS and NiTi archwires [372].
Shirakawa, Usui and colleagues [373,374] deposited tooth-colored Ag coatings on SS wires by eletropolishing. After 1 week of in vivo use with metallic brackets, no significant peeling was observed; however, the part of the wire in contact with the bracket slot turned black after 2 days [373].
  • Titanium (Ti):
Titanium is scarcely used as a surface material in fixed orthodontics. To the authors’ knowledge, only two research works (see Table 9) were found concerning fixed orthodontic appliances [375,376]. Ozeki and colleagues [375,377] coated Nitinol archwires with a 1 µm thick Ti layer by magnetron sputtering. This surface modification effectively reduced the Ni release in physiologically saline immersion tests up to 8 weeks (~5 times lower than uncoated wires). On the other hand, the superelasticity of the NiTi alloys decreased by almost 10.4% in comparison to uncoated wires, and the three-point bending tests (for 2 mm deflection) revealed a load decrease of 39% (from 4.18 to 2.55 N). Despite the decline in mechanical behavior, Ozeki et al. suggested that the performance of the sputter-coated archwires was not clinically compromised. In fact, when used in vivo for 4 weeks, both coated and uncoated archwires showed no inflammation signs, and little coating exfoliation—attributed to the use of pliers—was documented. Yet, the total sample size was very small: Only 5 patients [375]. Anuradha et al. [376] immersed NiTi archwires coated with sputter-deposited Ti (3–5 µm thick) in artificial saliva for 30 days. The results showed a well-adherent and smooth coating that remained mechanically stable over the test duration, without Ni or Ti detection in the artificial saliva solution.
  • Zinc (Zn):
Zinc is an essential trace element [378,379,380], and has been used in medicine and dentistry due to its antibacterial effects, bioactivity/biodegradability, and good mechanical properties [381,382,383,384]. Karandish et al. [381] first reported the study of PVD-deposited Zn coatings on SS archwires. According to the obtained results, these coatings can improve both tensile and load-bending strength. In addition, the deposited Zn layers successfully reduced the friction resistance for the bracket/wire angulation of 10° by a factor of ~50%.

5.1.2. Metal-Based Oxides, Nitrides and Carbides Coatings

The addition of interstitial elements—such as oxygen (O), nitrogen (N), carbon (C), or their combinations—to transition metals to form binary, ternary, or quaternary systems is a practical and common method to enhance the hardness performance of protective coatings, mainly those used in mechanical/tribological applications [385]. In orthodontics, Ti-based coatings have a special position. Although the authors did not account for any research work involving the Ti–C system in fixed orthodontic appliances, the same cannot be said concerning those of Ti–O and Ti–N, as can be seen in Table 9. In addition, some research on binary Zr–O, Al–O, Zn–O, W–C/N, and Cr–C systems will also be considered.
  • Ti–O System:
The Ti–O system, particularly the titanium oxide phase TiO2, is nowadays commercially available for coated brackets and archwires [64]. This surface material is well recognized as inert and more stable than the native oxide layer formed onto Ti-based alloys, preventing the degradation of the alloys and the subsequent oral metal release [386]. Electropolishing, followed by sterilization and thermal passivation, not only smooths the surface of bulk metals but also induces the formation of a desirable TiO2 top layer on the NiTi alloys [386].
Espinar et al. [387] heat-treated NiTi archwires to promote oxygen diffusion, forming a highly-adherent TiO2 passive film whose thickness was proportional to both treatment temperature and time. The resulting Ni-depleted surface (up to approximately 20 nm depth) exhibited an inferior friction coefficient against 316 SS disks in comparison to untreated archwires: CoF = 0.40 and 1.53, respectively. This oxidation treatment can, therefore, make the bracket/archwire sliding easier and prevent Ni-induced allergy and cytotoxicity since the release of metallic ions decreased.
Horiushi et al. [388] heat treated electrolytically grown TiO2 on NiTi alloys. They intended to develop a rutile crystalline phase and achieve photolytic activity when exposed to ultraviolet radiation (UV-A light, 315–400 nm), which could provide antibacterial activity. Other researchers explored the photocatalytic effect of TiO2-based coatings deposited on metallic substrates by sol–gel [389,390,391,392] and by magnetron sputtering [393,394,395,396,397,398] methods. In particular, surface functionalization of SS and NiTi archwires [389,390] and SS brackets [395,397] showed effective decays in bacterial adhesion and survival rates of representative oral microorganisms, namely for P. gingivalis, S. mutans, L. acidophilus, and Candida albicans, when exposed to UV-A radiation. Reactive oxygen species (including hydroxyl radicals) decompose surface organic molecules of those microorganisms and damage cell walls, allowing the control of dental plaque growth [395]. On the contrary, Awata et al. [391] found no statistically significant difference in the amount of bacterial adhesion after coating 316L SS disks via sol–gel. Zhang et al. [399] reported that nano-TiO2 coatings deposited on metallic brackets failed to show antibacterial activity up to 240 min in the dark. However, a nano-Ag–TiO2 system demonstrated antibacterial activities of ~80–94% after 20 min in the dark, depending on the bacteria strain.
Fatani et al. [398] added silver to sputter-deposited photocatalytic TiO2 coatings. The results confirmed the beneficial combination effect of this system: When compared with uncoated and Ag or TiO2-coated metallic SS brackets, Ag–TiO2 coatings showed lower S. mutans and P. Gingivalis adhesion and growth and biofilm development without cytotoxicity effects on human gingival fibroblast (HGF) cells. Kielan-Grabowska et al. [400] produced TiO2 layers containing Ag-NPs via the sol–gel dip-coating method on SS 316L orthodontic wires. The results revealed a significant decrease in the overall corrosion resistance of the coated samples, even though they might possess antimicrobial properties [400,401].
While TiO2 photocatalytic activity requires UV-A light, magnetron-sputtering-deposited and annealed N-doped Ti oxide films on SS brackets showed similar anti-adherent and antibacterial activity when exposed to visible light [393,394]. A 60-minute irradiation period was enough to prevent in vitro S. mutans growth for up to 3 months [393]. Nitrogen addition also seems beneficial to diminish the bracket/archwire friction [402].
Ghasemi et al. [358] deposited 60 and 100 µm thick TiO2 layers on SS brackets by PVD. Results showed that such oxide coatings effectively improved surface roughness and bacterial growth but failed to reduce friction against archwires. Other authors also reported an inhibition of bacterial adhesion on PVD-deposited TiO2 layers on SS orthodontic components [403].
Jung and colleagues treated both TMA and NiTi archwires with a plasma electrolytic oxidation (PEO, or micro-arc oxidation—MAO) method [404]. The 20–30 µm thick ceramic coating had a pleasant appearance and a higher biocompatibility and could improve the wear resistance of the NiTi archwire [405]. However, wire mechanical bending tests revealed the formation of defects in the coatings, cracking and chipping, and releasing sharp-edge particles that may harm teeth enamel or soft tissues [404].
Campeol et al. [406] reported that delamination of modified plasma-oxidized NiTi wires only occurs at high processing temperatures (>200 °C). The researchers prevented in vitro Ni release and delamination from substrates with an optimal 75–100 nm thick TiO2 layer at ~135 °C by using a simplified plasma-assisted method. Moreover, the superelasticity of plasma-treated Nitinol wires increased by 8.6% (at 13% strain) when compared with untreated metallic samples [406].
Supriadi et al. [407] suggested that a previous electropolishing step smooths and cleans the surface of 17–4 PH SS orthodontic brackets and consequently increases the adherence of magnetron-sputtered TiO2 coatings to the metallic substrates. Sol–gel-synthesized TiO2 (300–400 nm) directly formed on 316L SS plates by Fu et al. [408] showed long and thin cracks and just slightly increased surface hardness. The solution involved the deposition of a thick TiN interlayer (~8 µm) that prevented the formation of large cracks and significantly increased the hardness (from 280 HV to 810 HV). The main disadvantage reported by the researchers was the decrease in toughness due to surface oxidation, with variable corrosion behavior in a 0.9% saline solution depending on the annealing temperature.
Regarding new composite archwires, Liu et al. [409] used magnetron sputtering to attain TiO2 and N-doped TiO2 coatings. They reported an 87% reduction in the antibacterial activity against S. mutans of N-doped TiO2 films (~1.34% of N), in contrast to the undoped films (6% decrease only). The N addition was also beneficial to the in vitro biocompatibility and to the corrosion resistance of bulk CAW, enhancing both performances [409].
For further reading, an interesting meta-analysis was recently published by Solanki et al. [410] concerning in vitro studies using TiO2 coatings on orthodontic brackets. The notable antibacterial effect with low cytotoxicity on eukaryotic cells was highlighted.
Table 9. Overview of metal-based coatings deposited on different orthodontic appliances and substrates. PVD: physical vapor deposition; PIIID: plasma immersion ion implantation deposition; PEO: plasma electrolytic oxidation; CVD: chemical vapor deposition.
Table 9. Overview of metal-based coatings deposited on different orthodontic appliances and substrates. PVD: physical vapor deposition; PIIID: plasma immersion ion implantation deposition; PEO: plasma electrolytic oxidation; CVD: chemical vapor deposition.
CoatingsSubstrate MaterialsDeposition MethodsLiterature References
Transition Metal-based Coatings
RhNiTi archwiresCommercial[339] Katić et al. (2014)
RhNiTi archwiresCommercial[340] Katić et al. (2014)
RhNiTi archwiresCommercial[63] Kim et al. (2014)
RhNiTi archwiresCommercial[344] Trolic (2017)
RhNiTi archwiresCommercial[332] Katić et al. (2017)
RhNiTi archwiresCommercial[253] Lages et al. (2017)
RhNiTi archwiresCommercial[310] Matias et al. (2018)
RhNiTi archwiresCommercial[411] Asiry et al. (2018)
RhSS archwiresCommercial[348] Usui et al. (2018)
RhNiTi archwiresCommercial[341] Katić et al. (2018)
RhNiTi and SS archwiresCommercial[343] Nsaif et al. (2019)
RhNiTi archwiresCommercial[335] Alsanea and Shehri (2019)
RhNiTi archwiresCommercial[346] Trolic et al. (2019)
RhNiTi archwiresCommercial[347] Costa Lima (2019)
RhNiTi archwiresCommercial[345] Trolic et al. (2019)
RhNiTi archwiresCommercial[279] Mlinaric et al. (2019)
RhNiTi archwiresCommercial[337] Batista et al. (2020)
RhNiTi archwiresCommercial[336] Pinzan-Vercelino et al. (2020)
RhNiTi archwiresCommercial[350] Ramasamy et al. (2020)
RhNiTi archwiresCommercial[412] Madasamy et al. (2021)
RhNiTi archwiresCommercial[338] Osmani et al. (2022)
RhNiTi archwiresCommercial[342] Amorim et al. (2022)
RhSS archwiresCommercial[49] Ito et al. (2022)
Rhβ-Ti archwiresCommercial[349] Albawardi et al. (2022)
Rh-AuNiTi archwiresCommercial[334] Iijima et al. (2012)
Rh-AuNiTi archwiresCommercial[333] Albuquerque et al. (2017)
AuSS archwiresCommercial[63] Kim et al. (2014)
AuNiTi archwiresCommercial[64] Krishnan et al. (2014)
AuSS bracketsCommercial[62] Toy et al. (2014)
AuSS archwiresCommercial[49] Ito et al. (2022)
Pt-NPsNiTi archwiresElectrochemical deposition[353] Khonsari et al. (2011)
AgSS and NiTi archwiresThermal evaporation[357] Mhaske et al. (2015)
AgSS bracketsElectroplating[372] Arash et al. (2015)
AgSS bracketsMagnetron sputtering[398] Fatani et al. (2017)
AgSS bracketsPVD[358] Ghasemi et al. (2017)
AgSS wiresElectroplating[374] Usui et al. (2017)
AgSS wiresElectroplating[373] Shirakawa et al. (2017)
AgSS archwiresThermal vacuum evaporation[370] Shah et al. (2018)
AgSS bracketsGalvanic, PVD, and PIIID[359] Meyer-Kobbe et al. (2019)
AgSS brackets and archwiresThermal vacuum evaporation[371] Shah et al. (2023)
Nano-AgSS bracketsPVD[356] Metin-Gürsoy et al. (2016)
Nano-AgSS bracketsPVD[360] Metin-Gürsoy et al. (2017)
Nano-AgSS brackets and archwiresChemical deposition[361] Espinosa-Cristóbal et al. (2018)
Ag-NPsSS bandsThermal evaporation[363] Prabha et al. (2016)
Ag-NPsSS bandsThermal evaporation[362] Bindu et al. (2019)
Ag-NPsSS archwiresHydrothermal synthesis[365] Gonçalves et al. (2020)
Ag-NPsSS and Co–Cr bracketsCommercial[48] Jasso-Ruiz et al. (2020)
Ag-NPsNiTi archwiresElectrodeposition[366] Gil et al. (2020)
Ag-NPsSS bracketsThermal evaporation[367] Zeidan et al. (2022)
Ag-NPsMetallic bracketsLaser ablation[413] Tawakal et al. (2023)
Ag-NPsSS bandsElectrostatic spray-assisted vapor deposition[364] Bahrami et al. (2023)
Ag-NPsSS archwiresElectrochemical deposition[368] Anand et al. (2023)
Ag-NPs + TiO2-NPsSS archwiresElectrochemical deposition[368] Anand et al. (2023)
TiNiTi archwiresMagnetron sputtering[375] Ozeki et al. (2003)
TiNiTi archwiresSputtering[376] Anuradha et al. (2015)
ZnSS archwiresThermal evaporation[381] Karandish et al. (2021)
Metallic Oxides, Nitrides and Carbides Coatings
TiO2SS archwiresSol–gel dip-coating[389] Chun et al. (2007)
TiO2NiTi archwiresOxidation treatment[387] Espinar et al. (2011)
TiO2SS bracketsMagnetron sputtering[395] Shah et al. (2011)
TiO2NiTi archwiresOxidation process[414] Satiyorini and Pintowantoro (2013)
TiO2NiTi archwiresOxidation process[415] Pintowantoro and Setiyorini (2013)
TiO2SS and NiTi archwiresSol–gel dip-coating[390] Chhattani et al. (2014)
TiO2NiTi archwiresCommercial[64] Krishnan et al. (2014)
TiO2SS archwiresSol–gel dip-coating[392] Özyildiz et al. (2014)
TiO2metallic bracketsSpin-on deposition[399] Zhang et al. (2015)
TiO2SS bracketsPVD[358] Ghasemi et al. (2017)
TiO2SS bracketsMagnetron sputtering[397] Baby et al. (2017)
TiO2SS bracketsMagnetron sputtering[398] Fatani et al. (2017)
TiO2Composite archwiresMagnetron sputtering[409] Liu et al. (2017)
TiO2SS bracketsMagnetron sputtering[407] Supriadi et al. (2019)
TiO2SS bracketsMagnetron sputtering[416] Supriadi et al. (2019)
TiO2β-Ti and NiTi archwiresPEO[404] Jung et al. (2019)
TiO2SS archwiresPVD[403] Mollabasci et al. (2020)
TiO2NiTi wiresPlasma oxidation[406] Campeol et al. (2020)
TiO2NiTi wiresNH3 treatments[417] Kurtoğlu et al. (2020)
TiO2SS bracketsMagnetron sputtering[418] Math et al. (2021)
TiO2SS archwiresSol–gel dip-coating[400] Kielan-Grabowska et al. (2021)
TiO2SS wiresSol–gel dip-coating[401] Bącela et al. (2022)
N-doped TiO2SS bracketsMagnetron sputtering[394] Cao et al (2013)
N-doped TiO2composite archwiresMagnetron sputtering[409] Liu et al. (2017)
N-doped TiO2SS bracketsMagnetron sputtering[393] Salehi et al. (2018)
TiO2-xNyMetallic bracketsMagnetron sputtering[402] Li et al. (2014)
TiOxNyNiTi wiresNH3 treatments[417] Kurtoğlu et al. (2020)
Nano-Ag/TiO2metallic bracketsSpin-on deposition[399] Zhang et al. (2015)
Ag-TiO2SS bracketsMagnetron sputtering[398] Fatani et al. (2017)
TiO2 + Ag-NPsSS archwiresSol–gel dip-coating[400] Kielan-Grabowska et al. (2021)
TiO2 + Ag-NPsSS wiresSol–gel dip-coating[401] Bącela et al. (2022)
TiNNiTi archwiresCommercial[419] Kim and Johnson (1999)
TiNSS bracketsIon plating[420] Kao et al. (2002)
TiNNiTi and NiTiCu archwiresNitrogen gas difffusion[421] Gil et al. (2004)
TiNNiTi wiresCommercial[422] Iijima et al. (2010)
TiNSS bracketsIon plating[423] Huang et al. (2010)
TiNSS bracketsIon plating[424] Kao et al. (2011)
TiNSS bracketsCommercial[115] Saporeti et al. (2012)
TiNNiTi archwiresChemical deposition[414] Setiyorini and Pintowantoro (2013)
TiNNiTi archwiresCommercial[339] Katić et al. (2014)
TiNNiTi wiresCommercial[332] Katić et al. (2017)
TiNNiTi archwiresCommercial[340] Katić et al. (2014)
TiNNiTi archwiresCommercial[64] Krishnan et al. (2014)
TiNNiTi archwiresCommercial[425] Rongo et al. (2014)
TiNSS bracketsIon plating[426] Zuo et al. (2015)
TiNNiTi and β-Ti archwiresCommercial[427] Rongo et al. (2016)
TiNNiTi archwiresCommercial[344] Musa Trolić (2017)
TiNNiTi archwiresCommercial[341] Katić et al. (2018)
TiNSS and NiTi archwiresIon plating[428] Sugisawa et al. (2018)
TiNNiTi archwiresCommercial[345] Trolic et al. (2019)
TiNNiTi archwiresCommercial[279] Mlinaric et al. (2019)
TiNNiTi archwiresCommercial[346] Musa Trolic et al. (2019)
TiNNiTi archwiresNH3 treatments[417] Kurtoğlu et al. (2020)
TiNSS brackets and wires; NiTi archwiresMagnetron sputtering[429] Arici et al. (2021)
TiNSS bracketsCathodic cage[430] Teixeira et al. (2021)
TiNNiTi archwiresCommercial[338] Osmani et al. (2022)
TiNSS archwiresIon plating[49] Ito et al. (2022)
Ti/TiNNiTi archwiresMagnetron sputtering[431] Liu et al. (2014)
TiAlNβ-Ti archwiresCathodic arc PVD[432] Krishnan et al. (2011)
TiAlNβ-Ti archwiresCathodic arc PVD[433] Krishnan et al. (2012)
TiN doped with CaPSS bracketsCathodic cage[430] Teixeira et al. (2021)
ZrO2SS, NiTi and β-Ti archwiresSol–gel[434] Golshah and Feyli (2022)
Al2O3SS brackets and wires; NiTi archwiresMagnetron sputtering[429] Arici et al. (2021)
Al-SiO2NiTi archwiresMagnetron sputtering[435] Wu et al. (2022)
Black oxideNiTi archwiresCommercial[436] Krishnan et al. (2012)
ZnOSS bracketsMagnetron sputtering[418] Math et al. (2021)
CrNSS brackets and wires; NiTi archwiresMagnetron sputtering[429] Arici et al. (2021)
CrCSS archwiresElectroplating[348] Usui et al. (2018)
WC/Cβ-Ti archwiresMagnetron sputtering[432] Krishnan et al. (2011)
WC/Cβ-Ti archwiresMagnetron sputtering[433] Krishnan et al. (2012)
Metal Oxide-Based NPs coatings
ZnOSS bracketsSpray pyrolysis[437] Ramazanzadeh et al. (2015)
ZnO-NPsSS wiresChemical solution method[438] Kachoei et al. (2015)
ZnOSS archwiresBath immersion[439] Behroozian et al. (2016)
ZnONiTi archwiresChemical deposition[440] Kachoei et al. (2016)
ZnONiTi archwiresElectrochemical deposition[441] Hammad et al. (2020)
ZnONiTi archwiresChemical precipitationCVDSol–gel method[442] Gholami et al. (2021)
ZnOSS brackets and archwiresSol–gel method[443] Elhelbawy and Ellaithy (2021)
ZnO-NPsSS bracktesThermal evaporation[367] Zeidan et al. (2022)
ZnO-NPsNiTi archwiresHydrothermal method[444] Palanivel et al. (2022)
ZnO-NPsSS archwiresChemical precipitationHydrothermal method[445] Tanbakuchi et al. (2022)
AlO-NPsNiTi archwiresHydrothermal method[444] Palanivel et al. (2022)
CuOSS bracketsSpray pyrolysis[437] Ramazanzadeh et al. (2015)
CuO-NPsSS bracketsDip-coating[446] Ameli et al. (2022)
CuO-NPsSS bracketsDip-coating[447] Ameli et al. (2022)
ZnO-CuOSS bracketsSpray pyrolysis[437] Ramazanzadeh et al. (2015)
Ag-NPs+ZnO-NPsSS bracketsThermal evaporation[367] Zeidan et al. (2022)
Ag-HA-NPsSS bracketsDip-coating[446] Ameli et al. (2022)
Ag-HA-NPsSS bracketsDip-coating[447] Ameli et al. (2022)
TiO2-NPsNiTi archwiresMagnetron sputtering[448] Venkatesan et al. (2020)
TiO2-NPsSS bracketsDip-coating[446] Ameli et al. (2022)
TiO2-NPsSS bracketsDip-coating[447] Ameli et al. (2022)
TiO2-NPsSS archwiresDip-coating[449] Silveira et al. (2022)
TiO2-NPsSS archwiresMagnetron sputtering[368] Anand et al. (2023)
TiO2-NPsβ-Ti and NiTi archwiresDip-coating[450] Chaturvedi et al. (2023)
IF-NPs-reinforced Metal-based Coatings
WS2-reinforced NiSS archwiresElectrochemical co-deposition[451] Redlich et al. (2008)
WS2-reinforced NiSS archwiresElectrochemical co-deposition[452] Samorodnitzky-Naveh et al. (2010)
WS2-reinforced Ni-PSS archwiresElectroless deposition[453] Katz et al. (2006)
WS2-reinforced Ni-PSS archwiresElectroless deposition[454] Redlich et al. (2008)
MoS2-reinforced NiSS archwiresElectrochemical co-deposition[455] Gracco et al. (2019)
WS2-reinforced CoNiTi archwiresCo-electrodeposition[456] Samorodnitzky-Naveh et al. (2009)
WS2-reinforced CoNiTi archwiresCo-electrodeposition[452] Samorodnitzky-Naveh et al. (2010)
  • Ti–N System:
The titanium nitride system is commercially accessible as TiN-coated brackets and archwires [64,115,279,332,338,340,341,344,345,346]. The typically gold-colored TiN coatings may be produced onto orthodontic-compatible metals by several methods, including pulsed high-energy density plasma treatment with Ti electrodes [457], nitrogen gas (N2) diffusion [421] and ammonia gas (NH3) treatments [417], plasma immersion ion implantation [422,458], ion beam assisted deposition [459], ion plating [420,423,424,426,428,460], and other PVD methods [396,429,431,461]. As presented below, there is no academic consensus on the corrosion susceptibility of these nitride coatings in orthodontics.
Kao et al. [420] attempted to reduce the corrosion of SS brackets in artificial saliva by TiN ion plating, but no statistically significant difference between coated and uncoated brackets was observed on both metallic ion release and biocompatibility. Later attempts also failed: TiN ion-plated metallic brackets were unable to reduce friction against archwires in several media [423,424]. Opposite results by Zuo et al. [426] showed super low coefficients of friction (CoF < 0.03) under humid wear conditions in artificial saliva, high hardness (14.62 GPa), and improved adhesion of coated SS316L disks and brackets. Such a combination of mechanical and tribological properties is expected to decrease wear and release of toxic elements during orthodontic treatments. In addition, the coated samples presented an appealing golden color.
Jin et al. [460] were also optimistic about their research work: TiN ion-plated NiTi substrates may lead to improved biocompatibility, as fibroblasts’ adhesion and proliferation increased, whereas no cracks were reported during the three-point bending test after a 4099 N load. Deposition of the coatings onto SS and NiTi archwires diminished corrosion susceptibility in artificial saliva and friction force against metallic brackets up to a bracket/archwire angulation of 10°, as reported by Saporeti and coauthors [428]. The resulting higher tensile strength and stiffness of the SS wires are particularly useful properties in orthodontics. In turn, Mlinaric et al. [279] argued that, depending on the oral antiseptic used, the mechanical behavior of nitride NiTi (elastic properties and force delivery) can be affected, and Ni release may have been promoted during immersion in both the artificial saliva and the antiseptic.
A single TiN layer and a TiN/TiO2 multilayer deposited onto AISI 316L substrates by reactive magnetron sputtering effectively reduced the adhesion of calcium-precipitating bacteria, which are responsible for the formation of dental calculus [396]. Wang et al. [462] previously nitrided the surface of commercially pure Ti—a duplex treatment—decreasing ion release by almost 5 times compared to untreated samples when exposed to F-containing artificial saliva. When single Ti was used as a 200 nm thick interlayer for a 1 µm TiN coating on NiTi orthodontic archwires, potentiodynamic tests showed that localized corrosion was prevented in acidulated artificial saliva, both under loading and unloading conditions [431]. Other researchers reported that 1.6 and 3.2 µm thick Ti/TiN multilayer systems (16 bilayers) were more effective in improving the corrosion resistance of 316L SS samples than a simple TiN monolayer. This was demonstrated by using NaCl and Hank’s solutions [463]. However, the hardness decreased from ~13 to 10 GPa when using the multilayer system due to the presence of the softer Ti adhesion layers [463].
In opposition to the Ti–O system, no Ag addition to TiN coatings was found in the literature (Table 9). Instead, research work concerning the aluminum (Al) effect on TiN coatings by cathodic arc PVD was carried out [432,433]. The presence of Al (content not specified) decreased through-coating defects and prevented corrosion of the β -Ti archwires but maintained the mechanical properties (namely under unloading process). The results indicated that the TiAlN films effectively protected the wire during potentiodynamic tests in F-containing artificial saliva (no ionic release or coating damage was detected), while offering good biocompatibility without inducing apoptosis of oral epithelial cells [432].
Teixeira et al. [430] attempted to improve the surface properties of SS brackets by depositing single and calcium phosphate (CaPa)-containing TiN coatings via the cathodic cage method. Unfortunately, these coatings were unable to prevent the formation of S. mutans biofilms on the surface [430].
Recently, Kurtoğlu et al. [417] presented a simple and low-cost NH3 surface treatment in He atmosphere at 700 °C and assessed the in vitro corrosion resistance against artificial saliva. Thin TiN and TiOxNy films (<1 µm thick) significantly reduced the Ni release from shape-memory NiTi substrates. Nonetheless, the high treatment temperature used may have negatively impacted the mechanical properties of the bulk orthodontic alloys [357,464].
  • Zr–O and Al–O Systems:
Both Zr–O and Zr–Al systems—zirconia (ZrO2) and alumina (Al2O3)—are well recognized as bulk biomaterials due to their attractive combination of high corrosion resistance, low friction, high strength, and high wear resistance [465,466]. As surface material, ZrO2 coatings (1 µm thick) have been deposited onto AISI 316L substrates by electron-beam PVD [467]. These coatings showed excellent corrosion resistance in NaF-containing Fusayama–Meyer artificial saliva (2% NaF, pH = 3.0), while both S. mutans bacterial adhesion and fibroblast cytotoxicity decreased in vitro [467]. More recently, Golshah and Feyli [434] attempted to deposit ZrO2 on SS, TMA, and NiTi archwires using the sol–gel technique. Unfortunately, the researchers only successfully synthesized the coating on the TMA archwires, without a significant decrease in the static or kinetic friction against SS brackets [434].
Arici et al. [429] studied the binary Al–O system, deposited by magnetron sputtering, on SS brackets and SS and NiTi archwires. For this investigation, the researchers [429] designed a special experimental procedure to simulate the intraoral temperature variation (involving 250 thermal cycles, from −5 to +55 °C) and mechanical toothbrushing with fluoride toothpaste (12 h). The Al2O3 coatings presented good thermal, frictional, and brushing behavior. The referred work [429] also intended to study other sputtered coatings, TiN and CrN, which presented worse results in similar test conditions. While the TiN coatings showed small peeling areas, the CrN ones exhibited large detachment areas from the substrate. In fact, the CrN coatings failed the designed thermal, frictional, and brushing tests. The researchers therefore concluded that these Cr-based coatings were unsuitable for decreasing the coefficient of friction (CoF) in orthodontics. The best wear behavior was found for Al2O3 (or TiN)-coated archwires against metal brackets (CoF = 0.207 and 0.372, respectively). Wu et al. [435] deposited Al–SiO2 coatings by magnetron sputtering that smoothed the surface of NiTi and SS archwires and enhanced their overall corrosion resistance with no apparent cytotoxicity.
  • Zn–O System:
The Zn–O system, zinc oxide (ZnO), was also deposited as a coating by magnetron sputtering on SS brackets by Math et al. [418]. The coatings underwent thermal oxidation at 500 °C in an open-air furnace for 5 h. The researchers studied both anti-adhesion and antibacterial properties against S. mutans. The reported results were disappointing because they showed that ZnO coatings were unable to inhibit adhesion and growth of that bacterial strain.
  • W–C, W–N, Cr–C, and Cr–N Systems:
The binary W–C, W–N, Cr–C, and Cr–N systems—included in the well-known protective hard coatings for high-speed cutting operations [385]—have also merited attention in orthodontics. β -Ti archwires were functionalized with sputtered tungsten carbide/carbon (WC/C) coatings by Krishnan et al. [432,433]. The idea was to achieve protection against fluoride-induced corrosion. It was concluded that these coatings were mechanically stable and biocompatible; however, to the identified surface defects could act as initiators of pitting corrosion. In fact, deeper and accentuated cracks were observed in the WC/C coating after fluoride immersion tests [432]. The approach by single W and W–N [468] sputtered coatings was also disappointing on AISI 316L flat samples immersed in artificial saliva immersion. The external passive oxide phase revealed unprotective behavior, accompanied by severe and deep crack formation, as shown in Figure 18.
Usui et al. [348] compared the mechanical, frictional, and aesthetic properties of hard Cr–C-plated SS wires with uncoated, Rh-, and polymeric-coated SS orthodontic archwires. While the three-point bending test revealed similar values of flexural strength and modulus among all wire types, frictional forces significantly decreased against ceramic brackets. Compared with uncoated, Rh-, and polymer-coated samples, both maximum static and kinetic, dry, and wet frictional forces decreased by approximately 15–18, 42–45, and 17–22%, respectively.

5.1.3. Metal Oxide-Based Nanoparticles Coatings

Metal oxide-based NPs have been used to improve the performance of biocomponent surfaces, particularly as antimicrobial agents for multiple biomedical applications [469]. According to the literature, the following types of metallic oxide NPs have been considered to cover metallic fixed appliances: ZnO- [367,437,438,439,440,441,442,443,445] and/or CuO-NPs [437,446,447], AlO-NPs [444], TiO2-NPs [446,447,448,449], and Ag–HA-NPs [446,447], as presented in Table 9.
ZnO-NPs are novel dental materials with several potential applications in dentistry, including prosthodontic, endodontic, restorative, implantology, periodontal, and orthodontic fields [470]. As protective coatings, the research works concerning ZnO-NPs intend to reduce not only the bacterial adhesion and activity [367,437,440,441,442], but also the bracket/wire contact wear [439,440,441,443,444,445].
Excellent antibacterial results of 93, 96, and 98% against S. mutans have been achieved by Gholami et al. [442] with ZnO-NPs deposited by CVD, chemical precipitation and sol–gel methods, respectively, on the surface of NiTi archwires. Hammad et al. [441] also described excellent antibacterial activity for electrochemically deposited ZnO-NPs onto NiTi archwires against different bacterial strains (S. aureus, S. pyogens and E. coli). Other researchers [437] support the antibacterial effect of ZnO-NPs deposited by spray pyrolysis on SS brackets and by a chemical method on NiTi archwires [439]. Nonetheless, a mixture of CuO–ZnO–NPs seems to be more effective than a single ZnO-NPs for SS brackets, by using the spray pyrolysis method. Yet, the color of the surface unsuitably changed to copper (orange-red) [437].
Zeidan et al. [367] compared the antibacterial effect of ZnO- and/or Ag-NPs deposited on SS brackets by thermal evaporation. The researchers found the highest antibacterial effect against S. mutans and L. acidophilus for the ZnO/Ag-NPs coating. Interestingly, all coatings showed antibacterial effects that persisted for up to 3 months [367].
Concerning the mechanical properties, Elhelbawy and Ellaithy [443] deposited ZnO-NPs coatings via sol–gel on SS brackets and/or archwires, allowing a 64% decrease in the mean friction forces.Other researchers reported more modest reductions for coated NiTi archwires against uncoated SS brackets: ~34% [441,444], ~21% [440] (bracket/wire angulation up to 10°), and ~11% [445]. Palanivel et al. [444] also compared the friction forces of SS brackets against uncoated, ZnO-NPs-, or AlO-NPs-coated NiTi archwires, finding that the latter coating type showed an intermediate performance. ZnO-NPs-coated SS bands showed in vitro biocompatibility with HGF cells [364].
A different approach for reducing bacterial adhesion and subsequent enamel damage consisted of magnetron-sputtered TiO2 nanoparticles (TiO2-NPs) [448]. A thin layer (~81 nm) successfully prevented in vivo bacterial adhesion to NiTi archwires for up to 1 month. Coating physical stability was the main downside: Around 60% of the coating showed adhesion failures [448]. Ameli et al. [446,447] showed that TiO2-NPs and Cu-NPs coatings on SS brackets deposited by a sol–gel method possessed antibacterial properties and could significantly reduce the friction against metallic archwires. Chaturvedi and colleagues [450] dip-coated β-Ti and NiTi archwires with TiO2-NPs using different concentrations and dipping durations. The authors found an optimum NP concentration range of 1:2–1:6 for 48 h and a decrease in frictional forces for almost all coated wires [450].

5.1.4. Inorganic Fullerene-like Nanoparticles (IF-NPs)-Reinforced Metal-Based Coatings

Metal-based coatings containing inorganic fullerene-like nanoparticles (IF-NPs) are also listed in this review (Table 9). Round-shaped IF-NPS have been embedded into a metallic matrix to achieve composite coatings with self-lubricating properties. The wear mechanism is particularly relevant in bracket/archwires contacts in the salivary environment. Tungsten disulfide (WS2) was the IF-NPs of choice for nickel (Ni) [451,452,455], nickel–phosphorus (Ni–P) [453,454] and cobalt (Co) [452,456] matrixes, probably due to its excellent dry bulk lubricity.
Samorodnitzky-Naveh et al. [456] reinforced Co coatings with WS2 IF-NPs by using co-electrodeposition onto NiTi alloys. Successfully, the friction coefficient of the NiTi plates decreased by 66%—from 0.26 to 0.09 (as evaluated by pin-on-disk). Moreover, friction tests of the coated NiTi wires against SS brackets revealed minimum static and kinetic CoF of 0.080 and 0.061, respectively, for a maximum bracket/wire angulation of 5° [456].
Redlich and colleagues [451,453,454] selected Ni and Ni–P matrixes to study the WS2 IF-NPs effect over SS archwires. Reduced friction forces were claimed, both under dry and wet conditions: 17 and 54% decrease for 0 and 10° bracket/wire angulation, respectively. The 3–5 µm thick Ni-P/[Ni-P + IF-NPs] coatings reduced CoF by half (to ~0.05) after a 50-cycle ball-on-flat test [454].
Recently, Gracco et al. [455] developed and compared molybdenum disulfide (MoS2)—another solid lubricant—with WS2 IF-NPs-containing Ni coatings (~20 and 15 µm thick, respectively) on SS orthodontic wires. The coatings were synthesized by electrochemical co-deposition. Overall, both coating types always decreased the in vitro frictional forces against two types of metallic brackets, under dry and wet conditions, for 0 and 5° bracket/wire angulation. The researchers [455] observed no significant damage for minimum bending of the composite Ni/IF-NPs-coated wires, but a complete failure was registered for a maximum bending of 5° angulation.
The use of IF-NPs-containing coatings is, therefore, promising, but further research on their behavior under simulated oral conditions is still required, mainly regarding the effect of sulphur (S) on surfaces’ biocompatibility. Based on the reasons mentioned in the previous chapter, the selection of Ni-based matrices is quite undesirable in orthodontics.
Despite the large number of scientific works, further studies on durability, corrosion and wear resistance, and possible cytotoxicity to oral epithelial cells are still required for metals and their oxide, nitride, or carbide phase coatings and thin films.

5.2. Polymer-Based Coatings

Several polymeric coatings are clinically available for fixed orthodontics (Table 10). Polytetrafluoroethylene (PTFE) is, undoubtedly, the most studied thermoplastic polymer for orthodontic applications, followed by epoxy resins. Recently, research on polymeric composite materials has also gained importance, particularly those reinforced by ceramic NPs.

5.2.1. Thermoplastic and Thermoset Coatings

  • Polytetrafluoroethylene (PTFE):
Polytetrafluoroethylene, also known as TeflonTM, is widely used in multiple fields, including many medical applications [471,472] for its outstanding biocompatibility and inertness. In orthodontics, PTFE aesthetic coatings for archwires are commercially available due to their suitable white color, similar to the tooth’s tone.
A recent study [473] reported that PTFE coatings onto SS, NiTi, and β -Ti wires “exhibited low coloration, low microbial adhesion, low friction against metallic brackets, and high tolerance of detachment and wear against toothbrushing”, especially if produced at a low temperature (200 °C) using a two-stage spraying technique. Other researchers agree that PTFE coatings onto metallic archwires are highly stable; significantly improve the corrosion resistance of the substrates in acidulated artificial saliva [330,474,475], in Ringer’s solution [64], and in contact with certain food products [476,477]; maintain the surface morphology after cyclic mechanical loading of coated wires [330]; and decrease frictional forces against brackets [478,479]. PTFE coatings can also reduce biofilm adhesion to SS brackets. An in situ study [480] revealed that biofilm accumulation covered more than 22% of the uncoated brackets, in contrast to scarcely 4% of the PTFE-coated brackets. Still, the coated areas under high shear forces (e.g., bracket wings) partially degraded, exposing the metallic substrate.
Other reports, however, concluded that PTFE coatings directly affect load-deflection properties [333,481,482], surface roughness [333,481,483,484] and hardness [335] of NiTi archwires; undergo slight discoloration after 3 to 4 weeks of immersion in a staining solution [335,485]; and suffer from significant to severe coating detachment, both in vitro [486,487] and in vivo [373,488,489,490] (Figure 19). For instance, Rongo et al. [427] concluded that TeflonTM-coated NiTi archwires had similar biocompatibility to uncoated ones. Nevertheless, the mechanical response was different: Friction forces against metallic and aesthetic brackets increased, besides suffering delamination during 1 month of clinical use [425]. Scratch tests performed by Silva et al. [491] with four commercially available archwires coated with a polymer mixture (polytetrafluoroethylene and polyester) revealed high elasticity recoveries of more than 60%, but this mechanical behavior was accompanied by “delamination, crack propagation, and debris generation along the coatings scratches”. Lin et al. [492] ascertained that prolonged water immersion (up to 4 weeks) of PTFE-coated archwires had negatively impacted the resistance to sliding against ceramic brackets. Thus, the functionality and durability of such coatings in the oral environment were questioned.
  • Epoxy Resins:
Epoxy resins (containing an epoxide group) are another widely used polymer in medicine [493], existing as commercially available aesthetic coatings for archwires with an appealing white color.
Some researchers concluded that epoxy-coated archwires reduced bacterial adhesion [63], prevented dental plaque accumulation [494], and possessed similar mechanical properties (e.g., unloading force) [336] and surface roughness [495] to uncoated wires. Back in 1999, Kim and Johnson [419] recommended the use of epoxy-coated NiTi archwires due to their superior corrosion resistance when compared with uncoated ones. Similarly, Amorim et al. [342] concluded that epoxy resin coatings effectively enhanced the corrosion resistance of NiTi archwires in artificial saliva, for instance, by decreasing Ni release. Nevertheless, several in vitro studies reported major flaws, namely: low coating stability, durability, and color stability [138,348,474,485,486,496,497], lower hardness [335], and loading/unloading forces [497,498], or increased roughness [333,412,481] and friction forces against brackets [499], which may limit the clinical performance.
Abdulkader et al. [474] performed a simple durability test: Coated and uncoated archwires were immersed in artificial saliva at 3.5 and 6.75 pH values for 28 days at 37 °C. After the immersion period, samples were washed with normal saline solution, fixed to brackets with elastomeric ligatures, and subjected to mechanical stress by toothbrushing (210 s). The results revealed “rupture, roughness, and coatings damage in multiple locations”. Under acidic conditions, 48% coating loss was reported (Figure 20), while for more neutral pH (6.75), this value decreased to 31%. Alavi and Hosseini [496] immersed coated and uncoated samples in a commercial artificial saliva (at pH 6.7 and 37 °C), refreshed every day. After 3 weeks, they performed 500 thermocycles—10 min at 5 °C, 10 s at room temperature, and 10 min at 55 °C—and then a three-aesthetic bracket bending test. Very recently, Aboalnaga and colleagues [500] tested three commercially available epoxy-coated NiTi archwires and found undesirable surface changes after a one-month immersion test in artificial saliva, including lower surface hardness and higher roughness values.
Elyayyan et al. [498] obtained lower loading and unloading forces of new coated wires when compared with uncoated wires in a three-point bending test. The same researchers [501] also argued that epoxy-coated NiTi wires “had low aesthetic value”, as “25 % of the coatings was lost within 33 days in vivo and surface morphology showed severe deterioration”. Abdulkhabeer et al. [489] found mean coating losses between 18.3 and 28.6% of epoxy-coated NiTi archwires after 8 weeks of orthodontic treatment. The color stability of these coatings has also been questioned, since an extremely marked change in color was noticed in a 21-day immersion test in coffee staining solutions [350,485].
Shao et al. [502] used a mixture of epoxy resin and PTFE suspension, in different proportions and with white and yellow dies, to dip-coat NiTi archwires. Coatings with the color of human teeth were obtained. The results from the acute toxicity and mucous membrane irritation tests showed no negative signs regarding in vivo biocompatibility [502].
  • Polyether-ether-ketone (PEEK):
Polyether-ether-ketone resin is a cheap and non-cytotoxic advanced polymer that meets several chemical, tribological, and mechanical properties for orthodontics. To overcome the low PEEK adhesion to NiTi archwires, Sheiko et al. [503] proposed a new method to increase the adhesion of the polymeric coating. After cleaning the Nitinol archwires, they performed electrochemical polishing to create a thin and smooth TiO2 interlayer, above which a homogeneous PEEK top layer (~12 µm) grew via dip-coating. The coatings strongly adhered to the metallic substrate without delamination after immersion in Hank’s solution for up to 31 days, or 7 million compressing/stretching cycles at 20% strain. Regarding biocompatibility and corrosion properties, the coated archwires showed negligible Ni release during the immersion test and no cytotoxic effect on murine fibroblasts. However, local pressures above 2 GPa may disrupt the PEEK coating and unprotect the NiTi alloy surfaces against corrosion reactions [503].
More recently, Shirakawa et al. [504] followed a different method to cover NiTi archwires with PEEK resin: A simple outer polymeric tube around the arches. Friction tests revealed lower friction forces against brackets when compared to uncoated SS-, CoCr, and NiTi archwires. Since coated wires were able to slide more easily in contact, bracket slots maintained the as-received surface conditions [504].
Table 10. Overview of polymer-based coatings deposited on different orthodontic appliances and substrates. PECVD: plasma-enhanced chemical vapor deposition; n.s.—not specified.
Table 10. Overview of polymer-based coatings deposited on different orthodontic appliances and substrates. PECVD: plasma-enhanced chemical vapor deposition; n.s.—not specified.
CoatingsSubstratesDeposition MethodLiterature References
Thermoplastic and Thermoset Coatings
PTFENiTi and SS archwiresCommercial[330] Neumann et al. (2002)
PTFENiTi and SS archwiresCommercial[479] Husman et al. (2002)
PTFESS bracketsn.s.[480] Demling et al. (2010)
PTFESS and NiTi archwiresCommercial[478] Farronato et al. (2012)
PTFENiTi archwireCommercial[490] Zegan et al. (2012)
PTFENiTi archwireCommercial[64] Krishnan et al. (2014)
PTFENiTi archwiresCommercial[425] Rongo et al. (2014)
PTFENiTi archwiresCommercial[475] Mareci et al. (2015)
PTFENiTi archwiresCommercial[481] Ryu et al. (2015)
PTFENiTi archwiresCommercial[484] Choi et al. (2015)
PTFENiTi archwiresCommercial[476] Earar et al. (2016)
PTFENiTi archwiresCommercial[427] Rongo et al. (2016)
PTFENiTi archwiresCommercial[333] Albuquerque et al. (2017)
PTFENiTi archwiresCommercial[505] Rego et al. (2017)
PTFENiTi archwiresCommercial[477] Matei et al. (2016)
PTFENiTi archwiresCommercial[373] Shirakawa et al. (2017)
PTFENiTi archwiresCommercial[485] Rego et al. (2017)
PTFENiTi archwiresCommercial[310] Matias et al. (2018)
PTFENiTi archwiresCommercial[411] Asiry et al. (2018)
PTFENiTi archwiresCommercial[483] Dokku et al. (2018)
PTFESS archwiresCommercial[487] Shahabi et al. (2018)
PTFENiTi archwiresCommercial[335] Alsanea and Shehri (2019)
PTFENiTi archwiresCommercial[347] Costa Lima (2019)
PTFESS, NiTi and β-Ti archwiresThermal spraying [473] Kameda et al. (2020)
PTFENiTi archwiresCommercial[474] Abdulkader et al. (2020)
PTFENiTi archwiresCommercial[489] Abdulkhabeer et al. (2020)
PTFENiTi archwiresCommercial[486] Jejurikar et al. (2020)
PTFENiTi archwiresCommercial[337] Batista et al. (2020)
PTFENiTi archwiresCommercial[482] Elsaka et al. (2021)
PTFESS archwiresCommercial[492] Lin et al. (2021)
PTFEβ-Ti and SS archwiresSpray treatment[506] Zhou et al. (2023)
EpoxyNiTi arcwhiresCommercial[419] Kim and Johnson (1999)
EpoxyNiTi archwiresCommercial[501] Elayyan et al. (2008)
EpoxyNiTi archwiresCommercial[498] Elayyan et al. (2010)
EpoxyNiTi archwiresElectrostatic powder deposition[138] Bandeira et al. (2011)
EpoxyNiTi archwiresCommercial[496] Alavi et al. (2012)
EpoxyNiTi archwiresCommercial[494] Raji et al. (2014)
EpoxyNiTi archwiresCommercial[64] Krishnan et al. (2014)
EpoxyNiTi archwiresCommercial[63] Kim et al. (2014)
EpoxyNiTi archwiresCommercial[497] Pop et al. (2015)
EpoxyNiTi archwiresCommercial[484] Choi et al. (2015)
EpoxyNiTi archwiresCommercial[333] Albuquerque et al. (2017)
EpoxyNiTi archwiresCommercial[505] Rego et al. (2017)
EpoxySS archwiresCommercial[374] Usui et al. (2017)
EpoxyNiTi archwiresCommercial[485] Rego et al. (2017)
EpoxySS archwiresCommercial[348] Usui et al. (2018)
EpoxyNiTi archwiresCommercial[310] Matias et al. (2018)
EpoxyNiTi archwiresCommercial[411] Asiry et al. (2018)
EpoxyNiTi archwiresCommercial[483] Dokku et al. (2018)
EpoxyNiTi archwiresCommercial[335] Alsanea and Shehri (2019)
EpoxyNiTi archwiresCommercial[499] Dragomirescu et al. (2019)
EpoxyNiTi archwiresCommercial[495] Shamohammadi et al. (2019)
EpoxyNiTi archwiresCommercial[474] Abdulkader et al. (2020)
EpoxyNiTi archwiresCommercial[336] Pinzan-Vercelino et al. (2020)
EpoxyNiTi archwiresCommercial[489] Abdulkhabeer et al. (2020)
EpoxyNiTi archwiresCommercial[486] Jejurikar et al. (2020)
EpoxyNiTi archwiresCommercial[350] Ramasamy et al. (2020)
EpoxyNiTi archwiresCommercial[412] Madasamy et al. (2021)
EpoxySS archwiresCommercial[492] Lin et al. (2021)
EpoxyNiTi archwiresCommercial[342] Amorim et al. (2022)
Epoxyβ-Ti and SS archwiresSpray treatment[506] Zhou et al. (2023)
EpoxyNiTi archwiresCommercial[500] Aboalnaga et al. (2023)
PEEKNitinol wiresDip-coating deposition[503] Sheiko et al. (2016)
PEEK tubesSS, NiTi, and CoCr archwiresTube coverage[504] Shirakawa et al. (2018)
PENiTi archwiresCommercial[330] Neumann et al. (2002)
PENiTi archwiresCommercial[479] Husman et al. (2002)
PEN (PE naphthalate)SS archwiresCommercial[61] Ito et al. (2022)
PolyamideNiTi archwiresDipping treatment[507] Bravo et al. (2014)
Epoxy + PTFENiTi archwiresDip-coating deposition[502] Shao et al. (2009)
PTFE + polyesterSS and NiTi archwiresCommercial[491] da Silva et al. (2015)
Other Polymer Coatings
LysozymeComposite archwiresCoating protein deposition[318] He et al. (2020)
Hexamethyldisiloxane (HMDSO)SS bracketsPECVD[508] Tupinambá et al. (2017)
OrganosilaneSS bracketsSol–gel method[509] Oliveria et al. (2015)
2-methacryloyloxyethyl phosphorylcholineSS archwiresChemical deposition[510] Kunimatsu et al. (2022)
ChitosanBrackets (n.s.)Freeze-drying[511] Want et al. (2023)
Chitosan nanoparticlesSS brackets and archwiresSol–gel method[443] Elhelbawy and Ellaithy (2021)
Ag–chitosan nanoparticlesMetallic bracketsLaser ablation[413] Tawakal et al. (2023)
Polyoxazoline + tryptophanSS bracketsPlasma polymerization/Immobilization[512] Kumarasinghe et al. (2021)
1H,1H,2H,2H-perfluorodecyltrimethoxysilane(FAS) + bovine serum albumin (BSA)SS brackets and archwiresChemical deposition[513] Liu et al. (2018)
PVA hydrogelSS archwiresChemical deposition[514] MingWen et al. (2023)
Polydopamine + honokiol C-dotsSS bracketsHydrothermal method/chemical deposition[515] Wang et al. (2023)
Butyl-3-methylimidazolium chlorideNiTi archwiresIonic liquid coating[516] Ahmed et al. (2021)
ParyleneSS archwiresCommercial[492] Lin et al. (2021)
Parylene with Ag–Pt layerNiTi archwiresCommercial[334] Iijima et al. (2012)
Parylene with Ag–Pt layerNiTi archwiresCommercial[484] Choi et al. (2015)
Ag/biopolymer bilayerNiTi archwiresCommercial[63] Kim et al. (2014)
Ag/polymer bilayerNiTi archwiresCommercial[482] Elsaka et al. (2021)
Siliconeβ-Ti and SS archwiresDip coating[506] Zhou et al. (2023)
Ceramic-reinforced epoxy compositeβ-Ti and SS archwiresSpray deposition[506] Zhou et al. (2023)
Polymer (?)NiTi archwiresCommercial[496] Alavi et al. (2012)
Polymer (?)NiTi archwiresCommercial[425] Rongo et al. (2014)
Polymer (?)NiTi archwiresCommercial[427] Rongo et al. (2016)
Polymer (?)NiTi archwiresCommercial[488] Argalji et al. (2017)
Polymer (?)NiTi archwiresCommercial[499] Dragomirescu et al. (2019)
Polymer (?)NiTi archwiresCommercial[495] Shamohammadi et al. (2019)
Polymer (?)NiTi archwiresCommercial[482] Elsaka et al. (2021)
NPs reinforced Polymer Coatings
TiO2 reinforced EpoxyNiTi archwiresElectrophoretic deposition[517] Xu et al. (2019)
ZnO reinforced PVPNiTi archwiresElectrospinning[442] Gholami et al. (2021)
ZnO reinforced PVANiTi archwiresPolymer composite coating[442] Gholami et al. (2021)
  • Polyethylene (PE) and Polyethylene Naphthalate (PEN):
Polyethylene is a commercially available coating for NiTi archwires [330,479]. Husman et al. [479] indicated that the plasma-based synthesis process starts by removing microcontaminants on the wire surface through high-energy oxygen radicals, followed by the formation of titanium oxide, and finally the deposition of polyethylene. The surface is smooth due to the filling of its depressions [330,479]. Commercial PE-coated archwires showed an apparently superior in vitro corrosion resistance than the uncoated ones, even though the rupture potential has been lowered. This may indicate different corrosion processes. Moreover, PE-coated NiTi archwires showed no changes after cyclic mechanical loading [330] and reduced the friction forces against SS brackets [479]. Ito et al. [61] recommended the use of polyethylene naphthalate—also a commercial coating—for patients with metal allergies and when wire bending is required due to the in vitro mechanical stability of coated SS archwires. Even so, only toothbrushes with soft filaments should be used for oral hygiene procedures to prevent damaging the coating and, consequently, leading to corrosion of the metallic substrate [61].
  • Polyamides:
Polyamides are additional potential polymeric coatings for NiTi superelastic archwires. According to Bravo et al. [507], the dip-coated wires showed 85% less Ni release than uncoated samples after the 30-day electrochemical corrosion tests. During self-designed friction tests against SS- and Ti6Al4V-brackets, wear rates and both static and dynamic CoF were always inferior in comparison to other alloy archwires types (NiTi, TiMo, commercially pure Ti, Cu–NiTi or SS 304 AISI). In addition, both coated and uncoated wires had very similar superelastic behavior [507].
  • Other Polymer Coatings:
Lysozyme coatings were used by He et al. [318] to avoid Cu corrosion from laser-welded NiTi + SS CAW. These coatings, synthesized via liquid-phase deposition, presented antimicrobial properties. S. aureus activity decreased from 97 down to 59%, with increased electrochemical corrosion resistance, while cytocompatibility with fibroblast cells was higher for coated samples (79 vs. 83–90%).
Hexamethyldisilane (HMDSO) films were also deposited on conventional and self-ligating SS brackets by PECVD [508]. The plasma-polymerized resulting coatings were only effective in reducing surface roughness and bacterial adhesion in conventional brackets. Their uncoated, smoother surface and more favorable external geometry led to a better film deposition than in self-ligating brackets.
Organosilane-based coatings with super-hydrophobic properties were deposited via sol–gel on SS and ceramic brackets by Oliveira and colleagues [509]. A water contact angle of ~123° was reported, in addition to an exponential decrease in in vitro biofilm accumulation with increasing water contact angle of the surface up to 24 h of incubation (Figure 21).
Chitosan—an antibacterial carbohydrate polymer—was studied as a coating for orthodontic brackets. Non-crosslinked chitosan was deposited by freeze-drying from previously dissolved chitosan in 2% (w/v) acetic acid at different concentrations (1.0, 2.5, 5.0, and 10 mg/mL). Promising results were obtained, as S. mutans bacterial adhesion decreased by up to 99.6% for coated brackets [511].
Chitosan-NP materials were also studied as antibacterial surfaces for orthodontics. The deposition process was performed by sol–gel thin film dip coating. A decrease in friction of SS brackets and archwires against coated and/or uncoated counterparts [443] was accessed by the researchers. A significant reduction in friction forces and S. aureus activity was reported as well. Ag–chitosan NPs deposited on metallic brackets showed outstanding antibacterial activity against that bacterial strain [413].
MingWen et al. [514] quite recently deposited a toughly-adherent polyvinyl alcohol (PVA) hydrogel onto SS archwires. The authors reported a remarkable decrease in friction in water (CoF as low as 0.005) and a significant decrease in E. coli and S. aureus adhesion on coated samples.
Further attempts regarding polymer-based coatings deposited onto fixed orthodontic metallic components include: (i) 2-methacryloyloxyethyl phosphorylcholine: an hydrophilic and antibacterial phospholipid polymer with low friction property, which reduced the CoF of coated SS wires against brackets, and inhibited the in vitro adhesion of S. mutans and P. aeruginosa [510]; (ii) butyl-3-methylimidazolium chloride, which reduced the in vitro corrosion of NiTi archwires in fluoride- and/or bovine albumin-containing artificial saliva; (iii) 1H,1H,2H,2H-perfluorodecyltrimethoxysilane with bovine serum albumin, showing a non-bactericidal but highly antibacterial effect by preventing adhesion through a protein-mediated mechanism [513]; (iv) polydopamine and carbon dots: a fluorescent and antibacterial coating, with in vitro cytocompatibility, deposited on SS brackets [515]; (v) polyoxazoline with tryptophan coating, which decreased metal release and showed good in vitro biocompatibility to primary human dermal fibroblasts [512]; and (vi) silicone coatings on β-Ti and SS archwires [506].

5.2.2. Polymeric-Based Composite Coatings

Parylene polymer matrix reinforced with Ag and Pt is a commercially available coating on SS archwires [334,484,492]. The surface of the coated wires is typically rougher and softer, even though the bulk maintained its mechanical properties when compared with the uncoated ones [334].
Epoxy matrix reinforced with nanoscale TiO2 particles (~200 nm) was studied by Xu et al. [517] on NiTi sheets and round wires. The researchers concluded that the tooth-like-colored coatings were thinner (~22 µm) and presented better surface quality (smooth, defect-free, and uniform) than commercially available aesthetic epoxy coatings (~50 µm). Stress–strain curves of both coated and uncoated archwires, from loading-unloading tensile tests at 8% pre-strain, almost overlapped, indicating that superelasticity and shape-memory properties remained unchanged. After a 30-day immersion test in Fusayama-Meyer artificial saliva, both coating microhardness and adhesion slightly decreased (yet statistically non-significant), while Ni release was effectively blocked without corrosion signs. Moreover, no cytotoxicity in MG-63 cells was detected. Results are, therefore, optimistic, even though further variables must be added to the experiment to better simulate the oral environment; for instance, a more corrosive medium, bracket/wire contact, and microbiological attack [517]. Zhou et al. spray-coated β-Ti and SS archwires with commercial ceramic-reinforced epoxy coatings [488].
Recently, Gholami et al. [442] successfully reinforced polymeric coatings with ZnO-NPs: (i) ZnO-NPs-containing polyvinyl pyrrolidone (PVP) by electrospinning; and (ii) ZnO-NPs-containing PVA by a polymer composite coating method. However, such coatings displayed less antibacterial efficacy than simple ZnO-NPs coatings (non-composites) even though all surfaces were antibacterial: the reduction percentages regarding S. mutans viability with non-composite ZnO-NPs coatings ranged between ~93 and 99%, whereas ZnO-NPs-containing PVP and PVA composite films showed values of ~72 and 90%, respectively [442].
A more complex polymeric-based composite coating system, which consisted of a 3-layered structure covered with Ag nanoparticles onto a flat SS substrate, was developed by Lee et al. [518] and is worth mentioning. The studied architecture can be seen in Figure 22. While the two poly(3,4-ethylenedioxythiophene)-based inner layers enhanced adhesion to the substrates, the outmost dopamine layer decreased bacterial adhesion for both S. mutans and E. coli. Adsorbed silver (Ag+ and Ag-NPs) conferred antifouling and antibacterial properties without releasing toxic amounts of Ag species or compromising the in vitro cell viability of human gingival fibroblasts [518].

5.3. Ceramic-Based Coatings

The authors chose to include within this section the studies covering calcium (Ca) and phosphorus (P)-based coatings, comprehending hydroxyapatite (HA) materials, and bioactive glass (Table 11). An overview of covalent C-based materials is also presented, including hydrogen-free amorphous carbon nitride (a-CNx) and diamond-like carbon (DLC) coatings, as well as silicon-based coatings (SiN, SiC, and SiO2-NPs).

5.3.1. Hydroxyapatite (HA)

Hydroxyapatite (HA), that is, pentacalcium hydroxyl apatite [Ca10(PO4)6(OH)2], is a mineral frequently used in many biomedical applications, namely bone-related conditions [519], due to its excellent biocompatibility and osteoconductive properties. Moreover, HA is now commercially available in oral care products [520], and some attempts were made to use it as a surface material for orthodontic components.
Jiang et al. [521] coated porous SMA–Nitinol alloy with HA by the SHS method, doing a 5-day immersion in Hank’s solution after NH3 and NaOH aqueous solution treatments. The resulting bone-like HA layer reduced, by one order of magnitude, the Ni release after a 50-day body fluid immersion test. Kocijan et al. [522] confirmed the barrier property of the HA coatings: The electrodeposited layer effectively enhanced corrosion resistance in Hank’s solution of NiTi samples. Even a thin amorphous calcium phosphate layer (~600 nm thick) deposited by r.f. magnetron sputtering reduced nickel release from NiTi plates [523].
Unfortunately, only two studies were found using real orthodontic components. Setiyorini and Pintowantoro [414] synthesized biomimetic HA on NiTi wires by electrodeposition, revealing the highest cell viability percentage compared to other studied coatings. Dimasruhin et al. [524] deposited a SiO2-containing HA coating via electrodeposition. The researchers successfully coated NiTi archwires but only performed a morphological characterization.
The evident lack of studies dealing with metallic orthodontic components (brackets, archwires) is probably related to the ceramic-brittle nature of HA.

5.3.2. Bioactive Glasses

Bioactive glasses have also been studied by Kawaguchi et al. by deposition on both SS disks [525] and archwires [526] through an electrophoretic process under 10 or 15 V (direct or alternating current: d.c. or a.c., respectively). Coatings (1–4 µm thick) onto SS disks showed promising aesthetics, non-cytotoxicity, and remineralization ability up to 3 months of acid etching [525]. Archwires coated at 10 V demonstrated significantly higher friction against SS brackets up to 10° of bracket/archwire angulation [526]. This can be due to the lower coating hardness when compared with the metallic surface (<0.5 vs. 6.11 GPa, respectively). On the other hand, coatings deposited at 15 V, with a hardness value of ~2 GPa, provided similar friction as unmodified wires against SS brackets. The main downside reported was the disruption/detachment behavior in the bracket/archwire interface during the friction tests [526].

5.3.3. Silicon-Based Coatings

SiN and SiC-based coatings were deposited on SS brackets and archwires by Rapiejko et al. [527] in an attempt to reduce both friction and wear during the sliding of those orthodontic components. Those authors designed an ex-situ fretting test to allow micro-sliding (simulating the occlusion movements) to superimpose on macro-sliding (which in turn simulates the global displacement of the teeth), both in dry and wet (artificial saliva) environments. The lowest friction and the highest wear were found for the uncoated SS bracket and archwire contact, whereas the opposite was true for the coated pairs. The researchers concluded that due to such a higher CoF (0.50 vs. 0.20 for coated and uncoated pairs, respectively, in artificial saliva), and despite the lower wear, the coatings are unsuitable for medical applications [527]. Silveira et al. [449] also reported a decrease in surface roughness and friction forces in both dry and artificial saliva environments when SS orthodontic wires were coated with an SiO2-NPs-based film, when compared with uncoated ones, against brackets.

5.3.4. Carbon-Based Coatings

  • C and C–H Systems:
Single C and binary C–H systems are very broad groups of materials, encompassing the non-hydrogenated and hydrogenated amorphous carbon—a-C and a-C:H, respectively. Usually, these materials are ambiguously recognized as diamond-like carbon (DLC) in the literature [528,529], despite their variable C=C sp2 (similar to graphite) and C-C sp3 (similar to diamond) hybridization bondsand H content. DLC coatings are widely used in the medical field [530,531,532,533]. The major advantages of DLC coatings are their outstanding overall biocompatibility, corrosion resistance, achievable high hardness, and low CoF. For further knowledge on the DLC materials, it is recommended to read the well-known review by J. Robertson [534]. To the authors’ best knowledge, there is yet no commercial application in fixed orthodontics, and the first attempts to use DLC coatings on real orthodontic components date back to 2005 (Table 11). Studies focused on coating brackets and archwires with DLC are hereafter described.
Kobayashi et al. [535,536] were able to reduce the in vitro corrosion of NiTi archwires. The researchers successfully deposited a DLC coating (presumably a-C:H) with an SiC interlayer by ion beam plating with benzene as the reactive gas. The coating was stable for a 24 h continuous mechanical brushing and significantly decreased the Ni release in physiological saline solution up to 14 days (short-term, 80 °C) and 6 months (long-term, 37 °C) when compared with uncoated NiTi archwires. In all, the DLC coating showed excellent adhesion and mechanical properties to mechanical brushing and prevented corrosion [535]. In two quite similar reports [537,538], the ~1 µm thick a-C:H coatings—deposited by an arc discharge ion plating method with benzene—reduced Ni release from NiTi archwires by 80% in a 5-day static immersion test (physiological saline solution, 85 °C).
Simulating basic intraoral conditions is essential, but the corrosion susceptibility of orthodontic alloys increases in fluoride-acidic environments [29,42,95,104,276,277]—and daily-used mouthwashes and toothpastes include fluoride prophylactic agents in their compositions. Accordingly, the corrosion protection ability of a 100 nm thick DLC (likely a-C) film deposited by mirror-confinement-type electron cyclotron resonance (MCECR) plasma sputtering on NiTi archwires was studied against a NaF-based anticavity dental rinse (pH 4) [539]. All samples were statically immersed in an artificial saliva for 12 weeks at 37 °C, and additionally dipped for 5 min in a mouthwash (0.044% sodium fluoride, in acidulated phosphate solution) three times a day. While the surface roughness of uncoated samples increased by ~50% due to F-induced dimples and cracks, the coatings protected the substrate against F-induced corrosion, with only a slight increase in surface roughness (~8%) and no significant morphological alterations [539].
To reduce resistance to sliding, some researchers deposited a-C:H coatings onto orthodontic brackets [423,539,540,541,542] and archwires [535,537,538,539,542,543,544,545,546,547], or even both components [548]. Deposition techniques included Plasma-based ion implantation/deposition (PBIID) [540,542,544,547], MCECR plasma sputtering [539,548], arch-discharge ion plating [535,537,538,545], and PECVD [541,543,546]. To promote and evaluate bracket/wire sliding properties, such as frictional forces and friction coefficient, custom-made devices are usually selected.
Tantiwinyupong et al. [547] found significantly different static frictional forces between conventional and DLC-coated NiTi archwires against brackets (dry, room temperature 25 °C), representing a decrease of ~33%. The PBIID-deposited a-C:H coating also improved surface hardness from 1.06 to 11.44 GPa. Similar results were reported by Muguruma et al. [542] for NiTi and SS wires coated by PBIID against uncoated conventional or self-ligating brackets, up to a 10° angulation at room temperature and in a dry environment. The 500 nm thick a-C:H layer increased surface hardness from 11.6 to 17.6, and from 4.7 to 9.1 in the case of SS and NiTi, respectively. Thicker a-C:H coatings (~5–7 µm) produced by the same technique on SS brackets can also reduce both friction forces against SS brackets under dry and wet conditions [540].
Zhang et al. [546] produced a-C:H coatings (~1 µm) on SS wires by PECVD, obtaining a DLC surface smoother and 1.46 times harder than the uncoated wire. However, no statistically significant differences between the static friction coefficient or maximum static friction force were observed under dry conditions—only the kinetic friction coefficient decreased by ~40%. Kang and colleagues [548] also reported a decrease in the kinetic friction between SS wires and brackets. They tested MCECR plasma sputtering-produced, 200 nm thick a-C:H coatings, obtaining a reduction in the kinetic friction coefficient by ~80 and 70% under ambient air and artificial saliva, respectively, when at least the wire was coated. The lowest value of CoF = 0.11 was reported for both components under ambient air. Although kinetic friction seems irrelevant in orthodontics [215], the authors argued that their test involved oscillations with small displacements (±150 µm, 0.5 Hz), which recreated the discontinuous clinical motion (5 mm of distance, 10 mm/min) [548], in opposition to others [423,540,542].
Danisman et al. [549] deposited an 83 nm thick DLC top-coating above a Ti/TiN coating (36/687 nm) film on SS brackets by closed-field magnetron sputtering and measured the static friction force against SS archwires in air (5 mm of distance, 10 mm/min). The researchers reported a 28–39% decrease in the friction forces when coatings were used in comparison with uncoated brackets [549].
S. Huang et al. [539] coated NiTi archwires with a thin 100 nm a-C:H layer by PECVD. Surface roughness did not significantly change, but bracket/wire friction coefficients diminished up to 79.7% and 70.0% in ambient air and artificial saliva, respectively, when compared with uncoated tribo-couples. In a static immersion test simulating fluoride mouth rinse, surface roughness variations—caused by fluoride-induced corrosion—decreased by 91.3%, demonstrating the anti-corrosion behavior of amorphous carbon films [539].
The general tribological improvement by using a-C:H coatings may be related to the expected increase in surface hardness, which leads to a lower wear rate and, therefore, lower frictional forces [542]. Higher surface hardness is advantageous for orthodontics, as it may minimize binding or notching effects [542,547]. Muguruma et al. [544] pointed out the role of H on the a-C:H surface, suggesting that DLC coatings with high H content (>30 at.%) are potential candidates for this application.
Surface wettability may also play another important role in reducing static friction [541,543]. To further explore this idea, Akaike et al. [541] deposited unmodified and F- and Si-modified a-C:H coatings on SS bracket slots by PECVD. Under wet (PBS-sprayed) conditions, all coatings decreased static friction when compared with uncoated bracket slots, up to a bracket/wire angulation of 10°. However, a-C:H:F coatings showed the lowest static friction. This may be due to the higher hydrophobicity of such coating, reflected by the highest PBS and distilled water contact angles (~80 and 90°, respectively) when compared to the SS surface (~60°) and even to the a-C:H film (~70 and 75°). The a-C:H:Si films were more hydrophilic than the other two DLC-based coatings (contact angles of ~35°), and therefore the decrease in static friction was lower [541].
Mechanical brushing with fluorine-containing toothpaste is an indispensable daily hygiene procedure, especially for patients undergoing fixed orthodontic treatment [29,91,108]. Therefore, Oghoe, Kobayashi et al. studied the stability of 1 µm thick a-C:H coatings deposited on orthodontic NiTi archwires by arch-discharge ion plating with [535] and without [537,545] a SiC interlayer. The researchers simulated a 6-month daily procedure by mechanically brushing the coated wires with equipment carrying a toothbrush for 250 min, 350 min, and 24 h at a constant load of 35 g. The coatings protected the substrate that corroded when uncoated; no peeling, cracks, or even microstructural changes were detected, indicating good coatings’ adherence and stability against brushing hygiene procedures [535,537,545].
Designing coatings with antibacterial properties—for instance, by reducing bacterial adhesion—can minimize the formation of biofilm layers [550]. Currently, the eventual antibacterial properties of DLC coatings are non-consensual, even though the introduction of antibacterial agents is possible [551]. Furthermore, and to the authors’ best knowledge, little investigation has been conducted using amorphous carbon coatings specifically for orthodontics [550,552], and no study was found assessing the antibacterial properties of DLC-coated fixed orthodontic components. Further research is required to assess and enhance the antibacterial properties for orthodontic applications.
  • C–N System:
Carbon nitride (CNx) coatings are well-recognized biocompatible and chemically stable surface materials with attractive mechanical and tribological properties for biomedical applications. Following previous research [553], Wei et al. [554] coated SS orthodontic archwires with CNx by ion beam assisted deposition (IBAD). Published results show lower and more stable frictional forces and CoF under both ambient air and artificial saliva, especially at higher bracket/wire angulations (up to 15°). Compared with bare or TiN-coated SS304L disks, those IBAD-produced CNx thin films (<500 nm) presented the best surface properties, that is, a lower and more stable coefficient of friction (CoF < 0.2), biocompatibility with human fibroblasts and effective antibacterial properties against S. mutans [459].
Table 11. Overview of ceramic-based coatings deposited on different orthodontic appliances and substrates. CVD: chemical vapor deposition; PECVD: plasma-enhanced CVD; PBIID: plasma-based ion implantation and deposition; IBAD: ion beam-assisted deposition; MCECR: mirror-confinement-type electron cyclotron resonance.
Table 11. Overview of ceramic-based coatings deposited on different orthodontic appliances and substrates. CVD: chemical vapor deposition; PECVD: plasma-enhanced CVD; PBIID: plasma-based ion implantation and deposition; IBAD: ion beam-assisted deposition; MCECR: mirror-confinement-type electron cyclotron resonance.
CoatingsSubstrate MaterialsDeposition MethodsLiterature References
Hydroxyapatite and Bioactive Glass Coatings
Bioactive glassSS archwiresElectrophoretic deposition[526] Kawaguchi et al. (2020)
SiO2-reinforced HA NiTi archwiresElectrodeposition[524] Dimasruhin et al. (2014)
HANiTi archwiresElectrodeposition[414] Satiyorini and Pintowantoro (2013)
Silicon-based Coatings
SiNSS brackets and archwiresPECVD[527] Rapiejko et al. (2009)
SiCSS bracketsPECVD[527] Rapiejko et al. (2009)
SiO2-NPsSS archwiresCommercial ceramic paint[449] Silveira et al. (2022)
Carbon-based Coatings
DLC (a-C:H)NiTi archwiresIon beam plating[535] Kobayashi et al. (2005)
DLC (a-C:H)NiTi archwiresCVD[536] Kobayashi et al. (2005)
DLC (a-C:H)NiTi archwiresArch discharge ion plating[538] Ohgoe et al. (2006)
DLC (a-C:H)NiTi archwiresArch discharge ion plating[537] Ohgoe et al. (2007)
DLC (a-C:H)NiTi archwiresArch discharge ion plating[545] Kobayashi et al. (2007)
DLC (a-C:H)SS bracketsPECVD[423] Huang et a. (2010)
DLC (a-C:H)NiTi and SS archwiresPBIID[542] Muguruma et al.(2011)
DLC (a-C:H)SS bracketsPBIID[540] Muguruma et al. (2013)
DLC (a-C:H)NiTi archwiresPECVD[539] Huag et al. (2013)
DLC (a-C)SS archwiresMCECR plasma sputtering[548] Kang et al. (2015)
DLC (a-C:H)SS backet slotsPECVD[543] Akaike et al. (2015)
DLC (a-C:H)SS bracket slotsPECVD[541] Akaike et al. (2016)
DLC-F (a-C:H:F)SS bracket slotsPECVD[541] Akaike et al. (2016)
DLC-Si (a-C:H:Si)SS bracket slotsPECVD[541] Akaike et al. (2016)
DLC (a-C:H)SS archwiresPECVD[546] Zhang et al. (2016)
DLC (a-C:H)SS archwiresPBIID[544] Muguruma et al. (2018)
DLC (a-C:H)NiTi archwiresPBIID[547] Tantiwinyupong et al. (2019)
DLC (a-C)SS bracketsMagnetron sputtering[549] Danisman et al. (2021)
CNxSS archwiresIBAD[554] Wei et al. (2011)
Carbon-based Nanocomposite Coatings
Graphene sheets embedded in carbonSS archwiresMCECR plasma sputtering[555] Pan et al. (2022)
Graphene sheets embedded in carbonSS archwiresMCECR plasma sputtering[556] Wang et al. (2022)
  • Carbon-based Nanocomposite Coatings:
Graphene sheets embedded in carbon (GSEC) matrixes are a recently deposited material on SS archwires by Pan and coauthors [555,556] via MCECR plasma sputtering. The aim was to decrease the CoF against SS brackets in wet conditions (Fusayama–Meyer artificial saliva). The researchers reported the lowest CoF and wear rate of 0.05 and 0.11 × 10 6 mm3/Nm when sliding against three-row micro-groove-textured SS brackets (10,000 times, 150 μ m displacement stroke, 0.5 Hz, and 0.5–2.0 N). They proposed that the excellent friction and wear performances are due to the formation of a tribofilm on the contact interface formed by an adsorbed saliva layer and graphene sheets, coupled with the removal of debris with artificial saliva through the micro-grove (Figure 23) [555,556].

6. Conclusions

Contemporary orthodontics heavily relies on using fixed orthodontic appliances to treat dental malocclusions. Those devices are commonly manufactured using metallic bioalloys due to their suitable mechanical properties that induce the necessary tooth movement through the treatment time—usually around 2 years. Stainless steels and nickel–titanium stand out due to their widespread use, followed by CoCr- and other Ti-based alloys.
The main downside of using metallic biomaterials in vivo is the inevitable corrosion process in biological environments. In fact, the human mouth is an extreme corrosion-promoting scenario, encompassing frequent and complex variations in chemical composition, temperature, and pH, promoted by many factors such as diet, oral microorganisms, and the use of F--containing hygiene products. Consequently, several types of corrosion could occur, including pitting, crevice, fretting, and so-called microbiologically induced corrosion.
The main consequences of intraoral corrosion are presented in this overview. The release of metallics is the most frequently studied feature; 67 studies measured the metal content in various matrices. While concentrations of released metal ions are below toxic levels, reported allergic reactions to metals such as Ni during orthodontic treatments with fixed appliances raised flags. This topic is not new, but further clarification is still required. Moreover, corrosion can increase the bracket/archwire pair friction, which may make it difficult to determine the correct treatment progression. The use of fluoride-containing products—unquestionably important to maintain oral health—was also focused, as they can potentially aggravate metallic corrosion susceptibility.
The literature shows several paths that researchers have been following to improve the overall corrosion resistance and, in consequence, the biocompatibility of orthodontic alloys. Those include the replacement of metallic components with non-metallic or composite substitutes; the modification of the alloys’ chemical composition; the development of different manufacturing processes; and the application of surface modification techniques by depositing protective coatings. The versatile use of thin films and coatings stands out in this biomedical field and is particularly focused on this review, as it can maintain the crucial properties of the alloys’ bulk. Many different deposition technologies and materials have been used in multiple in vivo and in vitro efforts to protect against oral aging, from monolithic to composite architectures and micro- to nano-scale materials, to meet the best and safest oral practice demands. Unfortunately, the challenging oral environment still undermines the development of a fully effective coating, whereas the high variability of research methodologies challenges the direct comparison between the studies.
Data suggest that even existing commercially available alternatives have important drawbacks and are fallible. Further multidisciplinary research should, therefore, persist, namely by applying new coating materials from the surface engineering field to fixed orthodontics and conducting broader studies to mimic the intraoral cavity complexity.

Author Contributions

A.F. performed the literature review and drafted the work; C.S.L. designed the work, participated in drafting, and substantially revised it; A.C.S. substantially revised the work. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by national funds through FCT-Fundação para a Ciência e a Tecnologia, under the projects UIDB/00285/2020 and LA/P/0112/2020, and through the PhD Grant SFRH/BD/143905/2019 attributed to A. Fróis; and under the projects COMPETE/FEDER/POCI programs CENTRO-01-0247-FEDER-039880 and PTDC/EMD-EMD/21402020.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AASAtomic absorption spectrometry
a-CNxAmorphous carbon nitride
AISIAmerican Iron and Steel Institute
CAWComposite archwires
CVDChemical vapor deposition
CoFCoefficient of friction
d.h.m.Dry hair mass
DLDetection limit
DLCDiamond-like carbon
EDSEnergy-dispersive spectroscopy
FCCFace centered cubic
GSECGraphene sheets embedded carbon
HAHydroxyapatite
HGFHuman gingival fibroblasts
IARCInternational Agency for Research on Cancer
IBADIon beam-assisted deposition
ICP-AES/ICP-OESInductively coupled plasma-atomic/optical emission spectrometry
ICP-MSInductively coupled plasma mass spectroscopy
IF-NPsInorganic fullerene-like nanoparticles
ISOInternational organization for standardization
MAOMicro-arc oxidation
MCECRMirror-confinement-type electron cyclotron resonance
MICMicrobiologically induced corrosion
NPsNanoparticles
PCPolycarbonate
PECVDPlasma-enhanced chemical vapor deposition
PVDPhysical vapor deposition
PEPolyethylene
PEEKPoly-ether-ether-ketone
PENPolyethylene naphthalate
PETPolyethylene terephthalate
PEOPlasma electrolytic oxidation
PHPrecipitation hardening
PIIIDPlasma immersion ion implantation and deposition
POMPolyoxymethylene
PPPolyphenylene
PTFEPolytetrafluoroethylene
PUPolyurethane
PVAPolyvinyl alcohol
PVPPolyvinylpyrrolidone
SCCStress corrosion cracking
SEMScanning electron microscopy
SMAShape memory alloy
SRBSulphur reducing bacteria
SSStainless steel
XRFX-ray fluorescence

References

  1. Dental Board of Australia. Dental List of Recognised Specialties, Related Specialist Titles and Definitions. Available online: https://www.dentalboard.gov.au/Registration-Standards.aspx (accessed on 10 December 2022).
  2. Guo, L.; Feng, Y.; Guo, H.-G.; Liu, B.-W.; Zhang, Y. Consequences of Orthodontic Treatment in Malocclusion Patients: Clinical and Microbial Effects in Adults and Children. BMC Oral Health 2016, 16, 112. [Google Scholar] [CrossRef] [PubMed]
  3. Zou, J.; Meng, M.; Law, C.S.; Rao, Y.; Zhou, X. Common Dental Diseases in Children and Malocclusion. Int. J. Oral Sci. 2018, 10, 7. [Google Scholar] [CrossRef] [PubMed]
  4. Lombardo, G.; Vena, F.; Negri, P.; Pagano, S.; Barilotti, C.; Paglia, L.; Colombo, S.; Orso, M.; Cianetti, S. Worldwide Prevalence of Malocclusion in the Different Stages of Dentition: A Systematic Review and Meta-Analysis. Eur. J. Paediatr. Dent. 2020, 21, 115–122. [Google Scholar] [CrossRef]
  5. Cenzato, N.; Nobili, A.; Maspero, C. Prevalence of Dental Malocclusions in Different Geographical Areas: Scoping Review. Dent. J. 2021, 9, 117. [Google Scholar] [CrossRef] [PubMed]
  6. Redzepagic Vrazalica, L.; Ilic, Z.; Laganin, S.; Dzemidzic, V.; Tiro, A. An Epidemiological Study of Malocclusion and Occlusal Traits Related to Different Stages of Dental Development. S. Eur. J. Orthod. Dentofac. Res. 2017, 4, 9–13. [Google Scholar] [CrossRef]
  7. Mtaya, M.; Brudvik, P.; Astrom, A.N. Prevalence of Malocclusion and Its Relationship with Socio-Demographic Factors, Dental Caries, and Oral Hygiene in 12- to 14-Year-Old Tanzanian Schoolchildren. Eur. J. Orthod. 2009, 31, 467–476. [Google Scholar] [CrossRef]
  8. Jamilian, A.; Kiaee, B.; Sanayei, S.; Khosravi, S.; Perillo, L. Orthodontic Treatment of Malocclusion and Its Impact on Oral Health-Related Quality of Life. Open Dent. J. 2016, 10, 236–241. [Google Scholar] [CrossRef]
  9. Proffit, W.R.; Fields, H.W.; Sarver, D.M.; Ackerman, J.L. Malocclusion and Dentofacial Deformity in Contemporary Society. In Contemporary Orthodontics; Mosby: St. Louis, MO, USA; Elsevier: Amsterdam, The Netherlands, 2012; pp. 2–18. ISBN 978032308317. [Google Scholar]
  10. Abdallah, M.-N.; Lou, T.; Retrouvey, J.-M.; Suri, S. Biomaterials Used in Orthodontics: Brackets, Archwires, and Clear Aligners. In Advanced Dental Biomaterials; Khurshid, Z., Najeeb, S., Zafar, M.S., Sefat, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 541–579. ISBN 978-0-08-102476-8. [Google Scholar]
  11. Proffit, W.R.; Fields, H.W.; Sarver, D.M.; Ackerman, J.L. Contemporary Orthodontic Appliances. In Contemporary Orthodontics; Mosby: St. Louis, MO, USA; Elsevier: Amsterdam, The Netherlands, 2012; pp. 347–389. ISBN 978032308317. [Google Scholar]
  12. Proffit, W.R.; Fields, H.W.; Sarver, D.M.; Ackerman, J.L. Mechanical Principles in Orthodontic Force Control. In Contemporary Orthodontics; Mosby: St. Louis, MO, USA; Elsevier: Amsterdam, The Netherlands, 2012; pp. 312–346. ISBN 978032308317. [Google Scholar]
  13. Wen, Y.; Niu, Q.; Wang, A.; Yang, H.; Wang, H.; Jin, Z. Clear Aligner Treatment for an Adult with Severe Anterior Open Bite Malocclusion. AJO-DO Clin. Companion 2022, 2, 409–417. [Google Scholar] [CrossRef]
  14. Puppin Filho, A. Má Oclusão de Classe I Com Biprotrusão e Ausência Dos Primeiros Molares Inferiores. Dent. Press J. Orthod. 2011, 16, 119–129. [Google Scholar] [CrossRef]
  15. Aljhani, A.S.; Zawawi, K.H. Nonextraction Treatment of Severe Crowding with the Aid of Corticotomy-Assisted Orthodontics. Case Rep. Dent. 2012, 2012, 694527. [Google Scholar] [CrossRef]
  16. Kaya, D.; Taner, T.U. Management of an Adult with Spaced Dentition, Class III Malocclusion and Open-Bite Tendency. Eur. J. Dent. 2011, 5, 121–129. [Google Scholar] [CrossRef]
  17. Schroeder, M.A.; Schroeder, D.K.; Júnior, J.C.; da Silva Santos, D.J. Orthodontic Traction of Impacted Maxillary Canines Using Segmented Arch Mechanics. Dent. Press J. Orthod. 2019, 24, 79–89. [Google Scholar] [CrossRef] [PubMed]
  18. Cobourne, M.T.; DiBiase, A.T.; Seehra, J.; Papageorgiou, S.N. Should We Recommend Early Overjet Reduction to Prevent Dental Trauma? Br. Dent. J. 2022, 233, 387–390. [Google Scholar] [CrossRef] [PubMed]
  19. Fooladi, B.; MacCarthy, T.; Maloney, T.; Suri, L. Category 4: Class II Division 2 Malocclusion with Deep Overbite. Am. J. Orthod. Dentofac. Orthop. 2007, 132, 252–259. [Google Scholar] [CrossRef] [PubMed]
  20. Carriére, L. Nonsurgical Correction of Severe Skeletal Class III Malocclusion. J. Clin. Orthod. 2016, 50, 216–230. [Google Scholar] [PubMed]
  21. de Oliveira, S.R. Class III Malocclusion with Unilateral Posterior Crossbite and Facial Asymmetry. Dent. Press J. Orthod. 2010, 15, 182–191. [Google Scholar] [CrossRef]
  22. Moro, A.; Borges, S.W.; Spada, P.P.; Morais, N.D.; Correr, G.M.; Chaves, C.M.; Cevidanes, L.H.S. Twenty-Year Clinical Experience with Fixed Functional Appliances. Dent. Press J. Orthod. 2018, 23, 87–109. [Google Scholar] [CrossRef]
  23. Tsichlaki, A.; Chin, S.Y.; Pandis, N.; Fleming, P.S. How Long Does Treatment with Fixed Orthodontic Appliances Last? A Systematic Review. Am. J. Orthod. Dentofac. Orthop. 2016, 149, 308–318. [Google Scholar] [CrossRef]
  24. Proffit, W.R.; Fields, H.W.; Sarver, D.M.; Ackerman, J.L. The First Stage of Comprehensive Treatment: Alignment and Leveling. In Contemporary Orthodontics; Mosby: St. Louis, MO, USA; Elsevier: Amsterdam, The Netherlands, 2012; pp. 530–555. ISBN 978032308317. [Google Scholar]
  25. Business Wire Inc. Business Wire. Available online: https://www.businesswire.com/news/home/20200728005527/en/Global-Orthodontic-Supplies-Market-Worth-4.6-Billion (accessed on 10 September 2020).
  26. Eliaz, N. Corrosion of Metallic Biomaterials: A Review. Materials 2019, 12, 407. [Google Scholar] [CrossRef]
  27. Proffit, W.R.; Fields, H.W.; Sarver, D.M.; Ackerman, J.L. The Biology Basis of Orthodontic Therapy. In Contemporary Orthodontics; Mosby: St. Louis, MO, USA; Elsevier: Amsterdam, The Netherlands, 2012; pp. 278–311. ISBN 978032308317. [Google Scholar]
  28. Eliades, T.; Athanasiou, A.E. In Vivo Aging of Orthodontic Alloys: Implications for Corrosion Potential, Nickel Release, and Biocompatibility. Angle Orthod. 2002, 72, 222–237. [Google Scholar]
  29. Chaturvedi, T.P.; Upadhayay, S.N. An Overview of Orthodontic Material Degradation in Oral Cavity. Indian J. Dent. Res. 2010, 21, 275–284. [Google Scholar] [CrossRef] [PubMed]
  30. Sifakakis, I.; Eliades, T. Adverse Reactions to Orthodontic Materials. Aust. Dent. J. 2017, 62, 20–28. [Google Scholar] [CrossRef] [PubMed]
  31. Shukoor, K.M.; Shaj, F.; Shabeer, N.N.; Jayarajan, J. Nickel Allergies in Orthodontic Treatment. Int. J. Prev. Clin. Dent. Res. 2016, 3, 143–146. [Google Scholar] [CrossRef]
  32. Agarwal, P.; Upadhyay, U.; Tandon, R.; Kumar, S. Nickel Allergy and Orthodontics. Asian J. Oral Health Allied Sci. 2011, 1, 61–63. [Google Scholar]
  33. EN 10088-1; Stainless Steels—Part 1: List of Stainless Steels. European Committee for Standardization (CEN): Brussels, Belgium, 2014.
  34. Malik, N.; Dubey, R.; Kallury, A.; Chauksye, A.; Shrivastav, T.; Kapse, B.R. A Review of Orthodontic Archwires. J. Orofac. Res. 2015, 5, 6–11. [Google Scholar] [CrossRef]
  35. Olszewska, A.; Hanć, A.; Barałkiewicz, D.; Rzymski, P. Metals and Metalloids Release from Orthodontic Elastomeric and Stainless Steel Ligatures: In Vitro Risk Assessment of Human Exposure. Biol. Trace Elem. Res. 2020, 196, 646–653. [Google Scholar] [CrossRef]
  36. Brüngger, D.; Koutsoukis, T.; Al Jabbari, Y.S.; Hersberger-Zurfluh, M.; Zinelis, S.; Eliades, T. A Comparison of the Compositional, Microstructural, and Mechanical Characteristics of Ni-Free and Conventional Stainless Steel Orthodontic Wires. Materials 2019, 12, 3424. [Google Scholar] [CrossRef]
  37. Brantley, W.; Berzins, D.; Iijima, M.; Tufekçi, E.; Cai, Z. Structure/Property Relationships in Orthodontic Alloys. In Orthodontic Applications of Biomaterials; Eliades, T., Brantley, W.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 3–38. ISBN 9780081003831. [Google Scholar]
  38. Wendl, B.; Wiltsche, H.; Lankmayr, E.; Winsauer, H.; Walter, A.; Muchitsch, A.; Jakse, N.; Wendl, M.; Wendl, T. Metal Release Profiles of Orthodontic Bands, Brackets, and Wires: An in Vitro Study. J. Orofac. Orthop. Fortschritte Kieferorthopädie 2017, 78, 494–503. [Google Scholar] [CrossRef]
  39. Arango, S.; Peláez-Vargas, A.; García, C. Coating and Surface Treatments on Orthodontic Metallic Materials. Coatings 2013, 3, 1–15. [Google Scholar] [CrossRef]
  40. Arango Santander, S.; Luna Ossa, C.M. Stainless Steel: Material Facts for the Orthodontic Practitioner. Rev. Nac. Odontol. 2015, 11. [Google Scholar] [CrossRef]
  41. Walker, M.P.; Ries, D.; Kula, K.; Ellis, M.; Fricke, B. Mechanical Properties and Surface Characterization of Beta Titanium and Stainless Steel Orthodontic Wire Following Topical Fluoride Treatment. Angle Orthod. 2007, 77, 342–348. [Google Scholar] [CrossRef]
  42. Castro, S.M.; Ponces, M.J.; Lopes, J.D.; Vasconcelos, M.; Pollmann, M.C.F. Orthodontic Wires and Its Corrosion—The Specific Case of Stainless Steel and Beta-Titanium. J. Dent. Sci. 2015, 10, 1–7. [Google Scholar] [CrossRef]
  43. Cuy, J.L.; Mann, A.B.; Livi, K.J.; Teaford, M.F.; Weihs, T.P. Nanoindentation Mapping of the Mechanical Properties of Human Molar Tooth Enamel. Arch. Oral Biol. 2002, 47, 281–291. [Google Scholar] [CrossRef] [PubMed]
  44. Niinomi, M. Mechanical Properties of Biomedical Titanium Alloys. Mater. Sci. Eng. A 1998, 243, 231–236. [Google Scholar] [CrossRef]
  45. Bauer, S.; Schmuki, P.; Von Der Mark, K.; Park, J. Progress in Materials Science Engineering Biocompatible Implant Surfaces Part I: Materials and Surfaces. Prog. Mater. Sci. 2013, 58, 261–326. [Google Scholar] [CrossRef]
  46. Tian, K.; Darvell, B.W. Determination of the Flexural Modulus of Elasticity of Orthodontic Archwires. Dent. Mater. 2010, 26, 821–829. [Google Scholar] [CrossRef] [PubMed]
  47. Anusavice, K.J.; Phillips, R.W. Phillips’ Science of Dental Materials, 10th ed.; W.B. Saunders: Philadelphia, PA, USA, 1996; ISBN 9780721657417. [Google Scholar]
  48. Jasso-Ruiz, I.; Velazquez-Enriquez, U.; Scougall-Vilchis, R.J.; Morales-Luckie, R.A.; Sawada, T.; Yamaguchi, R. Silver Nanoparticles in Orthodontics, a New Alternative in Bacterial Inhibition: In Vitro Study. Prog. Orthod. 2020, 21, 24. [Google Scholar] [CrossRef]
  49. Arango-Santander, S.; Ramírez-Vega, C. Titanio: Aspectos Del Material Para Uso En Ortodoncia. Rev. Nac. Odontol. 2016, 12, 63–71. [Google Scholar] [CrossRef]
  50. Gioka, C.; Bourauel, C.; Zinelis, S.; Eliades, T.; Silikas, N.; Eliades, G. Titanium Orthodontic Brackets: Structure, Composition, Hardness and Ionic Release. Dent. Mater. 2004, 20, 693–700. [Google Scholar] [CrossRef]
  51. Uysal, I.; Yilmaz, B.; Atilla, A.O.; Evis, Z. Nickel Titanium Alloys as Orthodontic Archwires: A Narrative Review. Eng. Sci. Technol. Int. J. 2022, 36, 101277. [Google Scholar] [CrossRef]
  52. Wadood, A. Brief Overview on Nitinol as Biomaterial. Adv. Mater. Sci. Eng. 2016, 2016, 4173138. [Google Scholar] [CrossRef]
  53. Sifakakis, I.; Bourauel, C. Nickel–Titanium Products in Daily Orthodontic Practice. In Orthodontic Applications of Biomaterials; Elsevier: Amsterdam, The Netherlands, 2017; pp. 107–127. ISBN 9780081003831. [Google Scholar]
  54. Gravina, M.A.; Canavarro, C.; Elias, C.N.; Chaves, M.D.G.A.M.; Brunharo, I.H.V.P.; Quintão, C.C.A. Mechanical Properties of NiTi and CuNiTi Wires Used in Orthodontic Treatment. Part 2: Microscopic Surface Appraisal and Metallurgical Characteristics. Dent. Press J. Orthod. 2014, 19, 69–76. [Google Scholar] [CrossRef] [PubMed]
  55. Parvizi, F. The Load/Deflection Characteristics of Thermally Activated Orthodontic Archwires. Eur. J. Orthod. 2003, 25, 417–421. [Google Scholar] [CrossRef] [PubMed]
  56. Sufarnap, E.; Harahap, K.; Cynthiana, S.; Reza, M. Nickel and Copper Ion Release, Deflection and the Surface Roughness of Copper-Nickel-Titanium Orthodontic Archwire in Sodium Fluoride Solution. J. Orthod. Sci. 2023, 12, 44. [Google Scholar] [CrossRef] [PubMed]
  57. Seyyed Aghamiri, S.M.; Ahmadabadi, M.N.; Raygan, S. Combined Effects of Different Heat Treatments and Cu Element on Transformation Behavior of NiTi Orthodontic Wires. J. Mech. Behav. Biomed. Mater. 2011, 4, 298–302. [Google Scholar] [CrossRef] [PubMed]
  58. Sarul, M.; Kawala, B.; Kawala, M.; Antoszewska-Smith, J. Do the NiTi Low and Constant Force Levels Remain Stable in Vivo? Eur. J. Orthod. 2015, 37, 656–664. [Google Scholar] [CrossRef] [PubMed]
  59. Farzin-Nia, F.; Yoneyama, T. Orthodontic Devices Using Ti-Ni Shape Memory Alloys. In Shape Memory Alloys for Biomedical Applications; Elsevier: Amsterdam, The Netherlands, 2008; pp. 257–296. ISBN 9781845693442. [Google Scholar]
  60. Polychronis, G.; Al Jabbari, Y.S.; Eliades, T.; Zinelis, S. Galvanic Coupling of Steel and Gold Alloy Lingual Brackets with Orthodontic Wires: Is Corrosion a Concern? Angle Orthod. 2018, 88, 450–457. [Google Scholar] [CrossRef]
  61. Ito, A.; Kitaura, H.; Noguchi, T.; Ohori, F.; Mizoguchi, I. Analysis of Coating Loss from Coated Stainless Steel Orthodontic Wire. Appl. Sci. 2022, 12, 9497. [Google Scholar] [CrossRef]
  62. Toy, E.; Malkoc, S.; Corekci, B.; Bozkurt, B.S.; Hakki, S.S. Real-Time Cell Analysis of the Cytotoxicity of Orthodontic Brackets on Gingival Fibroblasts. J. Appl. Biomater. Funct. Mater. 2014, 12, 248–255. [Google Scholar] [CrossRef]
  63. Kim, I.-H.; Park, H.-S.; Kim, Y.K.; Kim, K.-H.; Kwon, T.-Y. Comparative Short-Term in Vitro Analysis of Mutans Streptococci Adhesion on Esthetic, Nickel-Titanium, and Stainless-Steel Arch Wires. Angle Orthod. 2014, 84, 680–686. [Google Scholar] [CrossRef]
  64. Krishnan, M.; Seema, S.; Kumar, A.V.; Varthini, N.P.; Sukumaran, K.; Pawar, V.R.; Arora, V. Corrosion Resistance of Surface Modified Nickel Titanium Archwires. Angle Orthod. 2014, 84, 358–367. [Google Scholar] [CrossRef]
  65. Hansen, D.C. Metal Corrosion in the Human Body: The Ultimate Bio-Corrosion Scenario. Electrochem. Soc. Interface 2008, 17, 31–34. [Google Scholar] [CrossRef]
  66. Maruthamuthu, S.; Rajasekar, A.; Sathiyanarayanan, S.; Muthukumar, N.; Palaniswamy, N. Electrochemical Behaviour of Microbes on Orthodontic Wires. Curr. Sci. 2005, 89, 988–996. [Google Scholar]
  67. Wolf, H.F.; Hassell, T.M. Biofilm—Plaque Formation on Tooth and Root Surfaces. In Color Atlas of Dental Hygene-Periodontology; Thieme: Stuttgart, Germany; New York, NY, USA, 2006; p. 24. ISBN 9783131417619. [Google Scholar]
  68. Zhou, Z.R.; Zheng, J. Tribology of Dental Materials: A Review. J. Phys. D Appl. Phys. 2008, 41, 113001. [Google Scholar] [CrossRef]
  69. Dawes, C.; Pedersen, A.M.L.; Villa, A.; Ekström, J.; Proctor, G.B.; Vissink, A.; Aframian, D.; McGowan, R.; Aliko, A.; Narayana, N.; et al. The Functions of Human Saliva: A Review Sponsored by the World Workshop on Oral Medicine VI. Arch. Oral Biol. 2015, 60, 863–874. [Google Scholar] [CrossRef] [PubMed]
  70. Mosca, A.C.; Chen, J. Food-Saliva Interactions: Mechanisms and Implications. Trends Food Sci. Technol. 2017, 66, 125–134. [Google Scholar] [CrossRef]
  71. de Almeida, P.D.V.; Grégio, A.M.T.; Machado, M.A.N.; de Lima, A.A.S.; Azevedo, L.R. Saliva Composition and Functions: A Comprehensive Review. J. Contemp. Dent. Pract. 2008, 9, 72–80. [Google Scholar]
  72. Humphrey, S.P.; Williamson, R.T. A Review of Saliva: Normal Composition, Flow, and Function. J. Prosthet. Dent. 2001, 85, 162–169. [Google Scholar] [CrossRef]
  73. Mystkowska, J.; Niemirowicz-Laskowska, K.; Łysik, D.; Tokajuk, G.; Dąbrowski, J.R.; Bucki, R. The Role of Oral Cavity Biofilm on Metallic Biomaterial Surface Destruction–Corrosion and Friction Aspects. Int. J. Mol. Sci. 2018, 19, 743. [Google Scholar] [CrossRef]
  74. Yakubov, G.E.; Macakova, L.; Wilson, S.; Windust, J.H.C.; Stokes, J.R. Aqueous Lubrication by Fractionated Salivary Proteins: Synergistic Interaction of Mucin Polymer Brush with Low Molecular Weight Macromolecules. Tribol. Int. 2015, 89, 34–45. [Google Scholar] [CrossRef]
  75. Castagnola, M.; Cabras, T.; Vitali, A.; Sanna, M.T.; Messana, I. Biotechnological Implications of the Salivary Proteome. Trends Biotechnol. 2011, 29, 409–418. [Google Scholar] [CrossRef] [PubMed]
  76. Choi, J.E.; Lyons, K.M.; Kieser, J.A.; Waddell, N.J. Diurnal Variation of Intraoral PH and Temperature. BDJ Open 2017, 3, 17015. [Google Scholar] [CrossRef] [PubMed]
  77. Moore, R.J.; Watts, J.T.F.; Hood, J.A.A.; Burritt, D.J. Intra-Oral Temperature Variation over 24 Hours. Eur. J. Orthod. 1999, 21, 249–261. [Google Scholar] [CrossRef] [PubMed]
  78. Kwak, D.Y.; Kim, N.Y.; Kim, H.J.; Yang, S.Y.; Yoon, J.E.; Hyun, I.A.; Nam, S.H. Changes in the Oral Environment after Tooth Brushing and Oral Gargling. Biomed. Res. 2017, 28, 7093–7097. [Google Scholar]
  79. Neyraud, E.; Palicki, O.; Schwartz, C.; Nicklaus, S.; Feron, G. Variability of Human Saliva Composition: Possible Relationships with Fat Perception and Liking. Arch. Oral Biol. 2012, 57, 556–566. [Google Scholar] [CrossRef] [PubMed]
  80. Poles, A.A.; Balcão, V.M.; Chaud, M.V.; Vila, M.M.D.C.; Aranha, N.; Yoshida, V.M.H.; Oliveira, J.M. Study of the Elemental Composition of Saliva of Smokers and Nonsmokers by X-Ray Fluorescence. Appl. Radiat. Isot. 2016, 118, 221–227. [Google Scholar] [CrossRef]
  81. Upadhyay, D.; Panchal, M.A.; Dubey, R.S.; Srivastava, V.K. Corrosion of Alloys Used in Dentistry: A Review. Mater. Sci. Eng. A 2006, 432, 1–11. [Google Scholar] [CrossRef]
  82. Hans, R.; Thomas, S.; Garla, B.; Dagli, R.J.; Hans, M.K. Effect of Various Sugary Beverages on Salivary PH, Flow Rate, and Oral Clearance Rate amongst Adults. Scientifica 2016, 2016, 5027283. [Google Scholar] [CrossRef]
  83. Dental Caries. Essentials of Oral Pathology and Oral Medicine; Cawson, R., Odell, E., Eds.; Churchill Livingstone: London, UK, 2008; pp. 40–59. ISBN 978-0443-10125-0. [Google Scholar]
  84. Walsh, L.J. Dental Plaque Fermentation and Its Role in Caries Risk Assessment. Int. Dent. S. Afr. 2006, 8, 34–40. [Google Scholar]
  85. Goel, I.; Navit, S.; Mayall, S.S.; Rallan, M.; Navit, P.; Chandra, S. Effects of Carbonated Drink & Fruit Juice on Salivary PH of Children: An in Vivo Study. Int. J. Sci. Study 2013, 1, 60. [Google Scholar]
  86. Lubis, H.F.; Simamora, G.H. Release of Nickel Ions and Changes in Surface Microstructure of Stainless Steel Archwire after Immersion in Tomato and Orange Juice. IOP Conf. Ser. Earth Environ. Sci. 2021, 912, 012018. [Google Scholar] [CrossRef]
  87. He, J.; Li, Y.; Cao, Y.; Xue, J.; Zhou, X. The Oral Microbiome Diversity and Its Relation to Human Diseases. Folia Microbiol. 2015, 60, 69–80. [Google Scholar] [CrossRef]
  88. Kilian, M.; Chapple, I.L.C.; Hannig, M.; Marsh, P.D.; Meuric, V.; Pedersen, A.M.L.; Tonetti, M.S.; Wade, W.G.; Zaura, E. The Oral Microbiome—An Update for Oral Healthcare Professionals. Br. Dent. J. 2016, 221, 657–666. [Google Scholar] [CrossRef]
  89. Marsh, P.D.; Head, D.A.; Devine, D.A. Dental Plaque as a Biofilm and a Microbial Community—Implications for Treatment. J. Oral Biosci. 2015, 57, 185–191. [Google Scholar] [CrossRef]
  90. Alasvand Zarasvand, K.; Rai, V.R. Microorganisms: Induction and Inhibition of Corrosion in Metals. Int. Biodeterior. Biodegrad. 2014, 87, 66–74. [Google Scholar] [CrossRef]
  91. Øgaard, B. White Spot Lesions During Orthodontic Treatment: Mechanisms and Fluoride Preventive Aspects. Semin. Orthod. 2008, 14, 183–193. [Google Scholar] [CrossRef]
  92. Weyant, R.J.; Tracy, S.L.; Anselmo, T.T.; Beltrán-Aguilar, E.D.; Donly, K.J.; Frese, W.A.; Hujoel, P.P.; Iafolla, T.; Kohn, W.; Kumar, J.; et al. Topical Fluoride for Caries Prevention. J. Am. Dent. Assoc. 2013, 144, 1279–1291. [Google Scholar] [CrossRef]
  93. ISO 8044:2015; Corrosion of Metals and Alloys—Basic Terms and Definitions. ISO (International Organization for Standardization): Geneva, Switzerland, 2015. Available online: https://www.iso.org/obp/ui/#iso:std:iso:8044:ed-4:v1:en (accessed on 18 May 2018).
  94. Sato, N. Basics of Corrosion Chemistry. In Green Corrosion Chemistry and Engineering: Opportunities and Challenges; Wiley: Hoboken, NJ, USA, 2011; pp. 1–32. ISBN 9783527329304. [Google Scholar]
  95. House, K.; Sernetz, F.; Dymock, D.; Sandy, J.R.; Ireland, A.J. Corrosion of Orthodontic Appliances-Should We Care? Am. J. Orthod. Dentofac. Orthop. 2008, 133, 584–592. [Google Scholar] [CrossRef]
  96. Hunt, N.; Cunningham, S.; Golden, C.; Sheriff, M. An Investigation into the Effects of Polishing on Surface Hardness and Corrosion of Orthodontic Archwires. Angle Orthod. 1999, 69, 433–440. [Google Scholar] [CrossRef]
  97. Strehblow, H.-H. Phenomenological and Electrochemical Fundamentals of Corrosion. In Materials Science and Technology, Corrosion and Environmental Degradation, Vol. I.; Schütze, M., Cahn, R., Haasen, P., Kramer, E., Eds.; Wiley-VCH: Weinheim, Germany, 2000; pp. 1–66. ISBN 3-527-29505-4. [Google Scholar]
  98. Eliades, T.; Zinelis, S.; Bourauel, C.; Eliades, G. Manufacturing of Orthodontic Brackets: A Review of Metallurgical Perspectives and Applications. Recent Pat. Mater. Sci. 2008, 1, 135–139. [Google Scholar] [CrossRef]
  99. Hanawa, T. Metal Ion Release from Metal Implants. Mater. Sci. Eng. C 2004, 24, 745–752. [Google Scholar] [CrossRef]
  100. Marcus, P.; Maurice, V. Passivity of Metals and Alloys. In Materials Science and Technology, Corrosion and Environmental Degradation, Vol. I.; Schütze, M., Cahn, R., Haasen, P., Kramer, E., Eds.; Wiley-VCH: Weinheim, Germany, 2000; pp. 131–169. ISBN 3-527-29505-4. [Google Scholar]
  101. Hedberg, Y.S.; Odnevall Wallinder, I. Metal Release from Stainless Steel in Biological Environments: A Review. Biointerphases 2016, 11, 018901. [Google Scholar] [CrossRef]
  102. Prasad, S.; Ehrensberger, M.; Gibson, M.P.; Kim, H.; Monaco, E.A. Biomaterial Properties of Titanium in Dentistry. J. Oral Biosci. 2015, 57, 192–199. [Google Scholar] [CrossRef]
  103. Mazinanian, N.; Hedberg, Y.S. Metal Release Mechanisms for Passive Stainless Steel in Citric Acid at Weakly Acidic PH. J. Electrochem. Soc. 2016, 163, C686–C693. [Google Scholar] [CrossRef]
  104. Walker, M.P.; White, R.J.; Kula, K.S. Effect of Fluoride Prophylactic Agents on the Mechanical Properties of Nickel-Titanium-Based Orthodontic Wires. Am. J. Orthod. Dentofac. Orthop. 2005, 127, 662–669. [Google Scholar] [CrossRef]
  105. Mystkowska, J. Biocorrosion of Dental Alloys Due to Desulfotomaculum Nigrificans Bacteria. Acta Bioeng. Biomech. 2016, 18, 87–96. [Google Scholar] [CrossRef]
  106. Daems, J.; Celis, J.-P.; Willems, G. Morphological Characterization of As-Received and in Vivo Orthodontic Stainless Steel Archwires. Eur. J. Orthod. 2009, 31, 260–265. [Google Scholar] [CrossRef]
  107. Porcayo-Calderon, J.; Casales-Diaz, M.; Salinas-Bravo, V.M.; Martinez-Gomez, L. Corrosion Performance of Fe-Cr-Ni Alloys in Artificial Saliva and Mouthwash Solution. Bioinorg. Chem. Appl. 2015, 2015, 930802. [Google Scholar] [CrossRef]
  108. Schiff, N. Galvanic Corrosion between Orthodontic Wires and Brackets in Fluoride Mouthwashes. Eur. J. Orthod. 2006, 28, 298–304. [Google Scholar] [CrossRef]
  109. Oh, K.T.; Choo, S.U.; Kim, K.M.; Kim, K.N. A Stainless Steel Bracket for Orthodontic Application. Eur. J. Orthod. 2005, 27, 237–244. [Google Scholar] [CrossRef]
  110. Mendes, B.D.A.B.; Ferreira, R.A.N.; Pithon, M.M.; Horta, M.C.R.; Oliveira, D.D. Physical and Chemical Properties of Orthodontic Brackets after 12 and 24 Months: In Situ Study. J. Appl. Oral Sci. 2014, 22, 194–203. [Google Scholar] [CrossRef]
  111. Jacoby, L.S.; Junior, V.D.S.R.; Campos, M.M.; de Menezes, L.M. Cytotoxic Outcomes of Orthodontic Bands with and without Silver Solder in Different Cell Lineages. Am. J. Orthod. Dentofac. Orthop. 2017, 151, 957–963. [Google Scholar] [CrossRef]
  112. Tahmasbi, S.; Ghorbani, M.; Masudrad, M. Galvanic Corrosion of and Ion Release from Various Orthodontic Brackets and Wires in a Fluoride-Containing Mouthwash. J. Dent. Res. Dent. Clin. Dent. Prospects 2015, 9, 159–165. [Google Scholar] [CrossRef]
  113. Petković Didović, M.; Jelovica Badovinac, I.; Fiket, Ž.; Žigon, J.; Rinčić Mlinarić, M.; Čanadi Jurešić, G. Cytotoxicity of Metal Ions Released from NiTi and Stainless Steel Orthodontic Appliances, Part 1: Surface Morphology and Ion Release Variations. Materials 2023, 16, 4156. [Google Scholar] [CrossRef]
  114. Schweitzer, P.A. Fundamentals of Metallic Corrosion, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2006; ISBN 9780429127137. [Google Scholar]
  115. Saporeti, M.P.; Mazzieiro, E.T.; Sales, W.F. In Vitro Corrosion of Metallic Orthodontic Brackets: Influence of Artificial Saliva with and without Fluorides. Dent. Press J. Orthod. 2012, 17, 24e1–24e7. [Google Scholar] [CrossRef]
  116. Fróis, A.; Mendes, A.R.; Pereira, S.A.; Louro, C.S. Metal Release and Surface Degradation of Fixed Orthodontic Appliances during the Dental Levelling and Aligning Phase: A 12-Week Study. Coatings 2022, 12, 554. [Google Scholar] [CrossRef]
  117. Zhang, Y. Corrosion Resistance of Passive Films on Orthodontic Bands in Fluoride-Containing Artificial Saliva. Int. J. Electrochem. Sci. 2017, 12, 292–304. [Google Scholar] [CrossRef]
  118. Kao, C.-T.; Huang, T.-H. Variations in Surface Characteristics and Corrosion Behaviour of Metal Brackets and Wires in Different Electrolyte Solutions. Eur. J. Orthod. 2010, 32, 555–560. [Google Scholar] [CrossRef]
  119. Mahato, N.; Sharma, M.R.; Chaturvedi, T.P.; Singh, M.M. Effect of Dietary Spices on the Pitting Behavior of Stainless Steel Orthodontic Bands. Mater. Lett. 2011, 65, 2241–2244. [Google Scholar] [CrossRef]
  120. Chaturvedi, T. Corrosion of Orthodontic Brackets in Different Spices:In Vitro Study. Indian J. Dent. Res. 2014, 25, 630. [Google Scholar] [CrossRef]
  121. Chiba, A.; Muto, I.; Sugawara, Y.; Hara, N. Pit Initiation Mechanism at MnS Inclusions in Stainless Steel: Synergistic Effect of Elemental Sulfur and Chloride Ions. J. Electrochem. Soc. 2013, 160, C511–C520. [Google Scholar] [CrossRef]
  122. Yang, S.; Zhao, M.; Feng, J.; Li, J.; Liu, C. Induced-Pitting Behaviors of MnS Inclusions in Steel. High Temp. Mater. Process. 2018, 37, 1007–1016. [Google Scholar] [CrossRef]
  123. Alnajjar, M.; Christien, F.; Barnier, V.; Bosch, C.; Wolski, K.; Fortes, A.D.; Telling, M. Influence of Microstructure and Manganese Sulfides on Corrosion Resistance of Selective Laser Melted 17-4 PH Stainless Steel in Acidic Chloride Medium. Corros. Sci. 2020, 168, 108585. [Google Scholar] [CrossRef]
  124. Hodgson, A.W.E.; Kurz, S.; Virtanen, S.; Fervel, V.; Olsson, C.-O.A.; Mischler, S. Passive and Transpassive Behaviour of CoCrMo in Simulated Biological Solutions. Electrochim. Acta 2004, 49, 2167–2178. [Google Scholar] [CrossRef]
  125. Bagatin, C.R.; Ito, I.Y.; Andrucioli, M.C.D.; Nelson-Filho, P.; Ferreira, J.T.L. Corrosion in Haas Expanders with and without Use of an Antimicrobial Agent: An in Situ Study. J. Appl. Oral Sci. 2011, 19, 662–667. [Google Scholar] [CrossRef]
  126. Wang, J.; Li, N.; Rao, G.; Han, E.; Ke, W. Stress Corrosion Cracking of NiTi in Artificial Saliva. Dent. Mater. 2007, 23, 133–137. [Google Scholar] [CrossRef]
  127. Kameda, T.; Oda, H.; Ohkuma, K.; Sano, N.; Batbayar, N.; Terashima, Y.; Sato, S.; TeradaA, K. Microbiologically Influenced Corrosion of Orthodontic Metallic Appliances. Dent. Mater. J. 2014, 33, 187–195. [Google Scholar] [CrossRef]
  128. Muyzer, G.; Stams, A.J.M. The Ecology and Biotechnology of Sulphate-Reducing Bacteria. Nat. Rev. Microbiol. 2008, 6, 441–454. [Google Scholar] [CrossRef]
  129. Jiang, J.; Chan, A.; Ali, S.; Saha, A.; Haushalter, K.J.; Lam, W.-L.M.; Glasheen, M.; Parker, J.; Brenner, M.; Mahon, S.B.; et al. Hydrogen Sulfide—Mechanisms of Toxicity and Development of an Antidote. Sci. Rep. 2016, 6, 20831. [Google Scholar] [CrossRef]
  130. Dhandapani, P.; MuraliKannan, M.; Anandkumar, B.; Maruthamuthu, S.; Manoharan, S.P. Electrochemistry of Calcium Precipitating Bacteria in Orthodontic Wire. Oral Sci. Int. 2014, 11, 22–29. [Google Scholar] [CrossRef]
  131. Petoumeno, E.; Kislyuk, M.; Hoederath, H.; Keilig, L.; Bourauel, C.; Jäger, A. Corrosion Susceptibility and Nickel Release of Nickel Titanium Wires during Clinical Application. J. Orofac. Orthop. 2008, 69, 411–423. [Google Scholar] [CrossRef]
  132. Harada, R.; Kokubu, E.; Kinoshita, H.; Yoshinari, M.; Ishihara, K.; Kawada, E.; Takemoto, S. Corrosion Behavior of Titanium in Response to Sulfides Produced by Porphyromonas Gingivalis. Dent. Mater. 2018, 34, 183–191. [Google Scholar] [CrossRef] [PubMed]
  133. Fukushima, A.; Mayanagi, G.; Nakajo, K.; Sasaki, K.; Takahashi, N. Microbiologically Induced Corrosive Properties of the Titanium Surface. J. Dent. Res. 2014, 93, 525–529. [Google Scholar] [CrossRef] [PubMed]
  134. Kumar, A.; Khanam, A.; Ghafoor, H. Effects of Intraoral Aging of Arch-Wires on Frictional Forces: An Ex Vivo Study. J. Orthod. Sci. 2016, 5, 109. [Google Scholar] [CrossRef] [PubMed]
  135. Cury, S.; Aliaga-Del Castillo, A.; Pinzan, A.; Sakoda, K.; Bellini-Pereira, S.; Janson, G. Orthodontic Brackets Friction Changes after Clinical Use: A Systematic Review. J. Clin. Exp. Dent. 2019, 11, e482–e490. [Google Scholar] [CrossRef] [PubMed]
  136. Eliades, T.; Bourauel, C. Intraoral Aging of Orthodontic Materials: The Picture We Miss and Its Clinical Relevance. Am. J. Orthod. Dentofac. Orthop. 2005, 127, 403–412. [Google Scholar] [CrossRef]
  137. Regis, S.; Soares, P.; Camargo, E.S.; Guariza Filho, O.; Tanaka, O.; Maruo, H. Biodegradation of Orthodontic Metallic Brackets and Associated Implications for Friction. Am. J. Orthod. Dentofac. Orthop. 2011, 140, 501–509. [Google Scholar] [CrossRef]
  138. Bandeira, A.M.B.; dos Santos, M.P.A.; Pulitini, G.; Elias, C.N.; da Costa, M.F. Influence of Thermal or Chemical Degradation on the Frictional Force of an Experimental Coated NiTi Wire. Angle Orthod. 2011, 81, 484–489. [Google Scholar] [CrossRef]
  139. Lima, A.A.S.; de Grégio, A.M.T.; Tanaka, O.; Machado, M.Â.N.; França, B.H.S. Tratamento Das Ulcerações Traumáticas Bucais Causadas Por Aparelhos Ortodônticos. Rev. Dent. Press Ortod. Ortop. Facial 2005, 10, 30–36. [Google Scholar] [CrossRef]
  140. Pires, L.P.B.; de Oliveira, A.H.A.; da Silva, H.F.; de Oliveira, P.T.; dos Santos, P.B.D.; Pinheiro, F.H. de S.L. Can Shielded Brackets Reduce Mucosa Alteration and Increase Comfort Perception in Orthodontic Patients in the First 3 Days of Treatment? A Single-Blind Randomized Controlled Trial. Am. J. Orthod. Dentofac. Orthop. 2015, 148, 956–966. [Google Scholar] [CrossRef]
  141. Kluemper, G.T.; Hiser, D.G.; Rayens, M.K.; Jay, M.J. Efficacy of a Wax Containing Benzocaine in the Relief of Oral Mucosal Pain Caused by Orthodontic Appliances. Am. J. Orthod. Dentofac. Orthop. 2002, 122, 359–365. [Google Scholar] [CrossRef] [PubMed]
  142. Bourauel, C.; Scharold, W.; Jäger, A.; Eliades, T. Fatigue Failure of As-Received and Retrieved NiTi Orthodontic Archwires. Dent. Mater. 2008, 24, 1095–1101. [Google Scholar] [CrossRef]
  143. Petoumeno, E.; Arndt, M.; Keilig, L.; Reimann, S.; Hoederath, H.; Eliades, T.; Jäger, A.; Bourauel, C. Nickel Concentration in the Saliva of Patients with Nickel-Titanium Orthodontic Appliances. Am. J. Orthod. Dentofac. Orthop. 2009, 135, 59–65. [Google Scholar] [CrossRef] [PubMed]
  144. Martín-Cameán, A.; Jos, Á.; Mellado-García, P.; Iglesias-Linares, A.; Solano, E.; Cameán, A.M. In Vitro and in Vivo Evidence of the Cytotoxic and Genotoxic Effects of Metal Ions Released by Orthodontic Appliances: A Review. Environ. Toxicol. Pharmacol. 2015, 40, 86–113. [Google Scholar] [CrossRef] [PubMed]
  145. Wang, J.J.; Sanderson, B.J.S.; Wang, H. Cyto- and Genotoxicity of Ultrafine TiO2 Particles in Cultured Human Lymphoblastoid Cells. Mutat. Res. Toxicol. Environ. Mutagen. 2007, 628, 99–106. [Google Scholar] [CrossRef] [PubMed]
  146. Messer, R.L.W.; Bishop, S.; Lucas, L.C. Effects of Metallic Ion Toxicity on Human Gingival Fibroblasts Morphology. Biomaterials 1999, 20, 1647–1657. [Google Scholar] [CrossRef] [PubMed]
  147. Chakravarthi, S.; Chitharanjan, A.; Padmanabhan, S. Allergy and Orthodontics. J. Orthod. Sci. 2012, 1, 83. [Google Scholar] [CrossRef] [PubMed]
  148. 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]
  149. Samitz, M.H.; Katz, S.A. Nickel Dermatitis Hazards from Prostheses: In Vivo and in Vitro Stabilization Studies. Br. J. Dermatol. 1975, 92, 287–290. [Google Scholar] [CrossRef]
  150. Mikulewicz, M.; Chojnacka, K. Release of Metal Ions from Orthodontic Appliances by In Vitro Studies: A Systematic Literature Review. Biol. Trace Elem. Res. 2011, 139, 241–256. [Google Scholar] [CrossRef]
  151. Macedo de Menezes, L.; Cardoso Abdo Quintão, C. The Release of Ions from Metallic Orthodontic Appliances. Semin. Orthod. 2010, 16, 282–292. [Google Scholar] [CrossRef]
  152. Urbutytė, K.; Barčiūtė, A.; Lopatienė, K. The Changes in Nickel and Chromium Ion Levels in Saliva with Fixed Orthodontic Appliances: A Systematic Review. Appl. Sci. 2023, 13, 4739. [Google Scholar] [CrossRef]
  153. Gjerdet, N.R.; Erichsen, E.S.; Remlo, H.E.; Evjen, G. Nickel and Iron in Saliva of Patients with Fixed Orthodontic Appliances. Acta Odontol. Scand. 1991, 49, 73–78. [Google Scholar] [CrossRef] [PubMed]
  154. Cempel, G.N.M. Nickel: A Review of Its Sources and Environmental Toxicolog. Pol. J. Environ. Stud. 2006, 15, 372–382. [Google Scholar]
  155. Smart, G.A.; Sherlock, J.C. Nickel in Foods and the Diet. Food Addit. Contam. 1987, 4, 61–71. [Google Scholar] [CrossRef] [PubMed]
  156. WHO (World Health Organization). Nickel in Drinking Water. In Background Document for Development of WHO Guidelines for Drinking-Water Quality; (WHO/SDE/WSH/04.08/55); World Health Organization: Geneva, Switzerland, 2005. [Google Scholar]
  157. Becker, W.; Kumpulainen, J. Contents of Essential and Toxic Mineral Elements in Swedish Market-Basket Diets in 1987. Br. J. Nutr. 1991, 66, 151–160. [Google Scholar] [CrossRef] [PubMed]
  158. CCME (Canadian Council of Ministers of the Environment). Scientific Criteria Document for Canadian Soil Quality Guidelines for the Protection of Environmental and Human Health: Nickel; Canadian Council of Ministers of the Environment: Winnipeg, MB, Canada, 2015. [Google Scholar]
  159. PHE (Public Health England). Nickel. In Toxicological Overview; Version 1; Toxicology Department, CRCE, Public Health England: London, UK, 2009. [Google Scholar]
  160. WHO (World Health Organization). Nickel. In Air Quality Guidelines for Europe; Regional Office for Europe, Ed.; World Health Organization: Copenhagen, Denmark, 2000; ISBN 9789289013581. [Google Scholar]
  161. Haber, L.T.; Bates, H.K.; Allen, B.C.; Vincent, M.J.; Oller, A.R. Derivation of an Oral Toxicity Reference Value for Nickel. Regul. Toxicol. Pharmacol. 2017, 87, S1–S18. [Google Scholar] [CrossRef]
  162. WHO (World Health Organization). Chromium. In Air Quality Guidelines for Europe; Regional Office for Europe, Ed.; World Health Organization: Copenhagen, Denmark, 2000; ISBN 9789289013581. [Google Scholar]
  163. Duda-Chodak, A.; Blaszczyk, U. The Impact of Nickel on Human Health. J. Elem. 2008, 13, 685–696. [Google Scholar]
  164. Forgacs, Z.; Massányi, P.; Lukac, N.; Somosy, Z. Reproductive Toxicology of Nickel—Review. J. Environ. Sci. Health Part A 2012, 47, 1249–1260. [Google Scholar] [CrossRef]
  165. IARC (International Agency for Research on Cancer). Nickel and Nickel Compounds; Academic Press: New York, NY, USA, 2011; Volume 100C, pp. 169–218. [Google Scholar] [CrossRef]
  166. World Health Organization. Guidelines for Drinking-Water Quality, 4th ed.; World Health Organization: Valletta, Malta, 2011. [Google Scholar]
  167. Genchi, G.; Carocci, A.; Lauria, G.; Sinicropi, M.S.; Catalano, A. Nickel: Human Health and Environmental Toxicology. Int. J. Environ. Res. Public Health 2020, 17, 679. [Google Scholar] [CrossRef]
  168. Silverberg, N.B.; Pelletier, J.L.; Jacob, S.E.; Schneider, L.C.; Cohen, B.; Horii, K.A.; Kristal, C.L.; Maguiness, S.M.; Tollefson, M.M.; Weinstein, M.G.; et al. Nickel Allergic Contact Dermatitis: Identification, Treatment, and Prevention. Pediatrics 2020, 145, e20200628. [Google Scholar] [CrossRef] [PubMed]
  169. Das, K.K.; Das, S.N.; Dhundasi, S.A. Nickel, Its Adverse Health Effects & Oxidative Stress. Indian J. Med. Res. 2008, 128, 412–425. [Google Scholar] [PubMed]
  170. Sahoo, N.; Kailasam, V.; Padmanabhan, S.; Chitharanjan, A.B. In-Vivo Evaluation of Salivary Nickel and Chromium Levels in Conventional and Self-Ligating Brackets. Am. J. Orthod. Dentofac. Orthop. 2011, 140, 340–345. [Google Scholar] [CrossRef] [PubMed]
  171. De Souza, R.M.; De Menezes, L.M. Nickel, Chromium and Iron Levels in the Saliva of Patients with Simulated Fixed Orthodontic Appliances. Angle Orthod. 2008, 78, 345–350. [Google Scholar] [CrossRef] [PubMed]
  172. Aǧaoǧlu, G.; Arun, T.; Izgü, B.; Yarat, A. Nickel and Chromium Levels in the Saliva and Serum of Patients with Fixed Orthodontic Appliances. Angle Orthod. 2001, 71, 375–379. [Google Scholar] [CrossRef] [PubMed]
  173. Masjedi, M.; Niknam, O.; Haghighat Jahromi, N.; Javidi, P.; Rakhshan, V. Effects of Fixed Orthodontic Treatment Using Conventional, Copper-Included, and Epoxy-Coated Nickel-Titanium Archwires on Salivary Nickel Levels: A Double-Blind Randomized Clinical Trial. Biol. Trace Elem. Res. 2016, 174, 27–31. [Google Scholar] [CrossRef]
  174. Yassaei, S.; Dadfarnia, S.; Ahadian, H.; Moradi, F. Nickel and Chromium Levels in the Saliva of Patients with Fixed Orthodontic Appliances. Orthodontics 2013, 14, e76–e81. [Google Scholar] [CrossRef]
  175. Haleem, R.; Ahmad Shafiai, N.; Mohd Noor, S. Perspective on Metal Leachables from Orthodontic Appliances: A Scoping Review. J. Int. Oral Health 2021, 13, 539–548. [Google Scholar]
  176. Dwivedi, A.; Tikku, T.; Khanna, R.; Maurya, R.P.; Verma, G.; Murthy, R.C. Release of Nickel and Chromium Ions in the Saliva of Patients with Fixed Orthodontic Appliance: An in-Vivo Study. Natl. J. Maxillofac. Surg. 2015, 6, 62–66. [Google Scholar] [CrossRef]
  177. Kocadereli, L.; Ataç, A.; Kale, S.; Özer, D. Salivary Nickel and Chromium in Patients with Fixed Orthodontic Appliances. Angle Orthod. 2000, 70, 431–434. [Google Scholar] [CrossRef]
  178. Fors, R.; Persson, M. Nickel in Dental Plaque and Saliva in Patients with and without Orthodontic Appliances. Eur. J. Orthod. 2006, 28, 292–297. [Google Scholar] [CrossRef] [PubMed]
  179. Olms, C.; Yahiaoui-Doktor, M.; Remmerbach, T.W. Contact Allergies to Dental Materials. Swiss Dent. J. 2019, 129, 571–579. [Google Scholar] [PubMed]
  180. Hafez, H.S.; Selim, E.M.N.; Kamel Eid, F.H.; Tawfik, W.A.; Al-Ashkar, E.A.; Mostafa, Y.A. Cytotoxicity, Genotoxicity, and Metal Release in Patients with Fixed Orthodontic Appliances: A Longitudinal in-Vivo Study. Am. J. Orthod. Dentofac. Orthop. 2011, 140, 298–308. [Google Scholar] [CrossRef]
  181. Singh, D.P.; Sehgal, V.; Pradhan, K.L.; Chandna, A.; Gupta, R. Estimation of Nickel and Chromium in Saliva of Patients with Fixed Orthodontic Appliances. World J. Orthod. 2008, 9, 196–202. [Google Scholar]
  182. Nayak, R.S.; Khanna, B.; Pasha, A.; Vinay, K.; Narayan, A.; Chaitra, K. Evaluation of Nickel and Chromium Ion Release During Fixed Orthodontic Treatment Using Inductively Coupled Plasma-Mass Spectrometer: An In Vivo Study. J. Int. Oral Health 2015, 7, 14–20. [Google Scholar]
  183. European Comission Regulation (EC). 1907/2006 of the European Parliament and of the Council of 18 December 2006—REACH. Available online: http://eur-lex.europa.eu/legal-content/EN/TXT/PDF/?uri=CELEX:32006R1907&from=en (accessed on 5 June 2023).
  184. Mikulewicz, M.; Chojnacka, K. Human Exposure to Trace Elements from Dental Biomaterials. In Recent Advances in Trace Elements; Wiley: Hoboken, NJ, USA, 2018; pp. 469–479. ISBN 9781119133780. [Google Scholar]
  185. Büdinger, L.; Hertl, M. Immunologic Mechanisms in Hypersensitivity Reactions to Metal Ions: An Overview. Allergy Eur. J. Allergy Clin. Immunol. 2000, 55, 108–115. [Google Scholar] [CrossRef]
  186. Saito, M.; Arakaki, R.; Yamada, A.; Tsunematsu, T.; Kudo, Y.; Ishimaru, N. Molecular Mechanisms of Nickel Allergy. Int. J. Mol. Sci. 2016, 17, 202. [Google Scholar] [CrossRef]
  187. Peltonen, L. Nickel Sensitivity. Int. J. Dermatol. 2008, 20, 352–353. [Google Scholar] [CrossRef]
  188. Zambelli, B.; Uversky, V.N.; Ciurli, S. Nickel Impact on Human Health: An Intrinsic Disorder Perspective. Biochim. Biophys. Acta Proteins Proteom. 2016, 1864, 1714–1731. [Google Scholar] [CrossRef]
  189. Buczko, P.; Szarmach, I.; Grycz, M.; Kasacka, I. Caspase-3 as an Important Factor in the Early Cytotoxic Effect of Nickel on Oral Mucosa Cells in Patients Treated Orthodontically. Folia Histochem. Cytobiol. 2017, 55, 37–42. [Google Scholar] [CrossRef] [PubMed]
  190. Luz, M.; Souza, A.; Haddad, A.; Tartomano, A.; Oliveira, P. In Vitro Cr(VI) Speciation in Synthetic Saliva after Releasing from Orthodontic Brackets Using Silica-Aptes Separation and GF AAS Determination. Quim. Nova 2016, 39, 951–955. [Google Scholar] [CrossRef]
  191. Setcos, J.C.; Babaei-Mahani, A.; Di Silvio, L.; Mjör, I.A.; Wilson, N.H.F. The Safety of Nickel Containing Dental Alloys. Dent. Mater. 2006, 22, 1163–1168. [Google Scholar] [CrossRef]
  192. Schuster, G.; Reichle, R.; Bauer, R.R.; Schopf, P.M. Allergies Induced by Orthodontic Alloys: Incidence and Impact on Treatment. J. Orofac. Orthop. 2004, 65, 48–59. [Google Scholar] [CrossRef]
  193. Flores-Bracho, M.G.; Takahashi, C.S.; Castillo, W.O.; Saraiva, M.C.P.; Küchler, E.C.; Matsumoto, M.A.N.; Ferreira, J.T.L.; Nelson-Filho, P.; Romano, F.L. Genotoxic Effects in Oral Mucosal Cells Caused by the Use of Orthodontic Fixed Appliances in Patients after Short and Long Periods of Treatment. Clin. Oral Investig. 2019, 23, 2913–2919. [Google Scholar] [CrossRef]
  194. Loyola-Rodríguez, J.P.; Lastra-Corso, I.; García-Cortés, J.O.; Loyola-Leyva, A.; Domínguez-Pérez, R.A.; Avila-Arizmendi, D.; Contreras-Palma, G.; González-Calixto, C. In Vitro Determination of Genotoxicity Induced by Brackets Alloys in Cultures of Human Gingival Fibroblasts. J. Toxicol. 2020, 2020, 1467456. [Google Scholar] [CrossRef] [PubMed]
  195. Bass, J.K.; Fine, H.; Cisneros, G.J. Nickel Hypersensitivity in the Orthodontic Patient. Am. J. Orthod. Dentofac. Orthop. 1993, 103, 280–285. [Google Scholar] [CrossRef] [PubMed]
  196. Kerosuo, H.; Kullaa, A.; Kerosuo, E.; Kanerva, L.; Hensten-Pettersen, A. Nickel Allergy in Adolescents in Relation to Orthodontic Treatment and Piercing of Ears. Am. J. Orthod. Dentofac. Orthop. 1996, 109, 148–154. [Google Scholar] [CrossRef] [PubMed]
  197. Staerkjaer, L.; Menne, T. Nickel Allergy and Orthodontic Treatment. Eur. J. Orthod. 1990, 12, 284–289. [Google Scholar] [CrossRef] [PubMed]
  198. Velasco-Ibáñez, R.; Lara-Carrillo, E.; Morales-Luckie, R.A.; Romero-Guzmán, E.T.; Toral-Rizo, V.H.; Ramírez-Cardona, M.; García-Hernández, V.; Medina-Solís, C.E. Evaluation of the Release of Nickel and Titanium under Orthodontic Treatment. Sci. Rep. 2020, 10, 22280. [Google Scholar] [CrossRef]
  199. Kochanowska, I.E.; Chojnacka, K.; Pawlak-Adamska, E.; Mikulewicz, M. Metallic Orthodontic Materials Induce Gene Expression and Protein Synthesis of Metallothioneins. Materials 2021, 14, 1922. [Google Scholar] [CrossRef] [PubMed]
  200. Muris, J.; Feilzer, A.J. Micro Analysis of Metals in Dental Restorations as Part of a Diagnostic Approach in Metal Allergies. Neuro Endocrinol. Lett. 2006, 27 (Suppl. 1), 49–52. [Google Scholar]
  201. Dunlap, C.L.; Vincent, S.K.; Barker, B.F. Allergic Reaction to Orthodontic Wire: Report of Case. J. Am. Dent. Assoc. 1989, 118, 449–450. [Google Scholar] [CrossRef] [PubMed]
  202. Ellis, P.E.; Benson, P.E. Potential Hazards of Orthodontic Treatment—What Your Patient Should Know. Dent. Update 2002, 29, 492–496. [Google Scholar] [CrossRef] [PubMed]
  203. Kolokitha, O.E.; Chatzistavrou, E. A Severe Reaction to Ni-Containing Orthodontic Appliances. Angle Orthod. 2009, 79, 186–192. [Google Scholar] [CrossRef] [PubMed]
  204. Noble, J.; Ahing, S.I.; Karaiskos, N.E.; Wiltshire, W.A. Nickel Allergy and Orthodontics, a Review and Report of Two Cases. Br. Dent. J. 2008, 204, 297–300. [Google Scholar] [CrossRef] [PubMed]
  205. Ehrnrooth, M.; Kerosuo, H. Face and Neck Dermatitis from a Stainless Steel Orthodontic Appliance. Angle Orthod. 2009, 79, 1194–1196. [Google Scholar] [CrossRef]
  206. Navarro-Triviño, F.J.; Ruiz-Villaverde, R. Contact Urticaria/Angioedema Caused by Nickel from Metal Dental Braces. Contact Dermat. 2020, 83, 425–427. [Google Scholar] [CrossRef] [PubMed]
  207. Maheshwari, S.; Verma, S.; Dhiman, S. Metal Hypersensitivity in Orthodontic Patients. J. Dent. Mater. Tech. 2015, 4, 111–114. [Google Scholar]
  208. Rahilly, G.; Price, N. Nickel Allergy and Orthodontics. J. Orthod. 2003, 30, 171–174. [Google Scholar] [CrossRef]
  209. Kolokitha, O.E.G.; Chatzistavrou, E. Allergic Reactions to Nickel-Containing Orthodontic Appliances: Clinical Signs and Treatment Alternatives. World J. Orthod. 2008, 9, 399–406. [Google Scholar]
  210. Gursoy, U.K.; Sokucu, O.; Uitto, V.J.; Aydin, A.; Demirer, S.; Toker, H.; Erdem, O.; Sayal, A. The Role of Nickel Accumulation and Epithelial Cell Proliferation in Orthodontic Treatment-Induced Gingival Overgrowth. Eur. J. Orthod. 2007, 29, 555–558. [Google Scholar] [CrossRef] [PubMed]
  211. Gurgel Maia, L.H.E.; de Lima Filho, H.L.; Araújo, M.V.A.; de Oliveira Ruellas, A.C.; de Souza Araújo, M.T. Incorporation of Metal and Color Alteration of Enamel in the Presence of Orthodontic Appliances. Angle Orthod. 2012, 82, 889–893. [Google Scholar] [CrossRef] [PubMed]
  212. Pazzini, C.A.; Pereira, L.J.; Marques, L.S.; Ramos-Jorge, J.; Aparecida da Silva, T.; Paiva, S.M. Nickel-Free vs Conventional Braces for Patients Allergic to Nickel: Gingival and Blood Parameters during and after Treatment. Am. J. Orthod. Dentofac. Orthop. 2016, 150, 1014–1019. [Google Scholar] [CrossRef] [PubMed]
  213. Martín-Cameán, A.; Jos, A.; Cameán, A.M.; Solano, E.; Iglesias-Linares, A. Genotoxic and Cytotoxic Effects and Gene Expression Changes Induced by Fixed Orthodontic Appliances in Oral Mucosa Cells of Patients: A Systematic Review. Toxicol. Mech. Methods 2015, 25, 440–447. [Google Scholar] [CrossRef] [PubMed]
  214. Downarowicz, P.; Mikulewicz, M. Trace Metal Ions Release from Fixed Orthodontic Appliances and DNA Damage in Oral Mucosa Cells by in Vivo Studies: A Literature Review. Adv. Clin. Exp. Med. 2017, 26, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  215. Burrow, S.J. Friction and Resistance to Sliding in Orthodontics: A Critical Review. Am. J. Orthod. Dentofac. Orthop. 2009, 135, 442–447. [Google Scholar] [CrossRef] [PubMed]
  216. Kusy, R.P.; Whitley, J.Q. Friction between Different Wire-Bracket Configurations and Materials. Semin. Orthod. 1997, 3, 166–177. [Google Scholar] [CrossRef] [PubMed]
  217. Kusy, R.P.; Whitley, J.Q. Influence of Archwire and Bracket Dimensions on Sliding Mechanics: Derivations and Determinations of the Critical Contact Angles for Binding. Eur. J. Orthod. 1999, 21, 199–208. [Google Scholar] [CrossRef]
  218. Articolo, L. Influence of Ceramic and Stainless Steel Brackets on the Notching of Archwires during Clinical Treatment. Eur. J. Orthod. 2000, 22, 409–425. [Google Scholar] [CrossRef]
  219. Prashant, P.; Nandan, H.; Gopalakrishnan, M. Friction in Orthodontics. J. Pharm. Bioallied Sci. 2015, 7, 334. [Google Scholar] [CrossRef]
  220. Bishara, S.E.; Barrett, R.D.; Selim, M.I. Biodegradation of Orthodontic Appliances. Part II. Changes in the Blood Level of Nickel. Am. J. Orthod. Dentofac. Orthop. 1993, 103, 115–119. [Google Scholar] [CrossRef]
  221. Kerosuo, H.; Moe, G.; Hensten-Pettersen, A. Salivary Nickel and Chromium in Subjects with Different Types of Fixed Orthodontic Appliances. Am. J. Orthod. Dentofac. Orthop. 1997, 111, 595–598. [Google Scholar] [CrossRef] [PubMed]
  222. Eliades, T.; Trapalis, C.; Eliades, G.; Katsavrias, E. Salivary Metal Levels of Orthodontic Patients: A Novel Methodological and Analytical Approach. Eur. J. Orthod. 2003, 25, 103–106. [Google Scholar] [CrossRef] [PubMed]
  223. Faccioni, F.; Franceschetti, P.; Cerpelloni, M.; Fracasso, M.E. In Vivo Study on Metal Release from Fixed Orthodontic Appliances and DNA Damage in Oral Mucosa Cells. Am. J. Orthod. Dentofac. Orthop. 2003, 124, 687–693. [Google Scholar] [CrossRef] [PubMed]
  224. Levrini, L.; Lusvardi, G.; Gentile, D. Nickel Ions Release in Patients with Fixed Orthodontic Appliances. Minerva Stomatol. 2006, 55, 115–121. [Google Scholar]
  225. Menezes, L.M.; Quintão, C.A.; Bolognese, A.M. Urinary Excretion Levels of Nickel in Orthodontic Patients. Am. J. Orthod. Dentofac. Orthop. 2007, 131, 635–638. [Google Scholar] [CrossRef]
  226. Amini, F.; Borzabadi Farahani, A.; Jafari, A.; Rabbani, M. In Vivo Study of Metal Content of Oral Mucosa Cells in Patients with and without Fixed Orthodontic Appliances. Orthod. Craniofacial Res. 2008, 11, 51–56. [Google Scholar] [CrossRef]
  227. Fernández-Miñano, E.; Ortiz, C.; Vicente, A.; Calvo, J.L.; Ortiz, A.J. Metallic Ion Content and Damage to the DNA in Oral Mucosa Cells of Children with Fixed Orthodontic Appliances. BioMetals 2011, 24, 935–941, Correction in BioMetals 2018, 31, 679–679. [Google Scholar] [CrossRef]
  228. Natarajan, M.; Padmanabhan, S.; Chitharanjan, A.; Narasimhan, M. Evaluation of the Genotoxic Effects of Fixed Appliances on Oral Mucosal Cells and the Relationship to Nickel and Chromium Concentrations: An in-Vivo Study. Am. J. Orthod. Dentofac. Orthop. 2011, 140, 383–388. [Google Scholar] [CrossRef]
  229. Freitas, M.P.M.; Oshima, H.M.S.; Menezes, L.M. Release of Toxic Ions from Silver Solder Used in Orthodontics: An in-Situ Evaluation. Am. J. Orthod. Dentofac. Orthop. 2011, 140, 177–181. [Google Scholar] [CrossRef]
  230. Mikulewicz, M.; Chojnacka, K.; Zielińska, A.; Michalak, I. Exposure to Metals from Orthodontic Appliances by Hair Mineral Analysis. Environ. Toxicol. Pharmacol. 2011, 32, 10–16. [Google Scholar] [CrossRef]
  231. Amini, F.; Jafari, A.; Amini, P.; Sepasi, S. Metal Ion Release from Fixed Orthodontic Appliances—An in Vivo Study. Eur. J. Orthod. 2012, 34, 126–130. [Google Scholar] [CrossRef]
  232. Ousehal, L.; Lazrak, L. Change in Nickel Levels in the Saliva of Patients with Fixed Orthodontic Appliances. Int. Orthod. 2012, 10, 190–197. [Google Scholar] [CrossRef] [PubMed]
  233. Amini, F.; Rakhshan, V.; Mesgarzadeh, N. Effects of Long-Term Fixed Orthodontic Treatment on Salivary Nickel and Chromium Levels: A 1-Year Prospective Cohort Study. Biol. Trace Elem. Res. 2012, 150, 15–20. [Google Scholar] [CrossRef] [PubMed]
  234. Talic, N.F.; Alnahwi, H.H.; Al-Faraj, A.S. Nickel and Chromium Levels in the Saliva of a Saudi Sample Treated with Fixed Orthodontic Appliances. Saudi Dent. J. 2013, 25, 129–133. [Google Scholar] [CrossRef] [PubMed]
  235. Abtahi, M.; Jahanbin, A.; Yaghoubi, M.; Esmaily, H.; Zare, H. Are More Nickel Ions Accumulated in the Hair of Fixed Orthodontic Patients? Indian J. Dent. Res. 2013, 24, 298. [Google Scholar] [CrossRef] [PubMed]
  236. Amini, F.; Rahimi, H.; Morad, G.; Mollaei, M. The Effect of Stress on Salivary Metal Ion Content in Orthodontic Patients. Biol. Trace Elem. Res. 2013, 155, 339–343. [Google Scholar] [CrossRef] [PubMed]
  237. Martín-Cameán, A.; Jos, A.; Calleja, A.; Gil, F.; Iglesias, A.; Solano, E.; Cameán, A.M. Validation of a Method to Quantify Titanium, Vanadium and Zirconium in Oral Mucosa Cells by Inductively Coupled Plasma-Mass Spectrometry (ICP-MS). Talanta 2014, 118, 238–244. [Google Scholar] [CrossRef] [PubMed]
  238. Martín-Cameán, A.; Jos, A.; Calleja, A.; Gil, F.; Iglesias-Linares, A.; Solano, E.; Cameán, A.M. Development and Validation of an Inductively Coupled Plasma Mass Spectrometry (ICP-MS) Method for the Determination of Cobalt, Chromium, Copper and Nickel in Oral Mucosa Cells. Microchem. J. 2014, 114, 73–79. [Google Scholar] [CrossRef]
  239. Martín-Cameán, A.; Molina-Villalba, I.; Jos, A.; Iglesias-Linares, A.; Solano, E.; Cameán, A.M.; Gil, F. Biomonitorization of Chromium, Copper, Iron, Manganese and Nickel in Scalp Hair from Orthodontic Patients by Atomic Absorption Spectrometry. Environ. Toxicol. Pharmacol. 2014, 37, 759–771. [Google Scholar] [CrossRef]
  240. Mikulewicz, M.; Wołowiec, P.; Janeczek, M.; Gedrange, T.; Chojnacka, K. The Release of Metal Ions from Orthodontic Appliances Animal Tests. Angle Orthod. 2014, 84, 673–679. [Google Scholar] [CrossRef] [PubMed]
  241. Mikulewicz, M.; Wołowiec, P.; Loster, B.; Chojnacka, K. Metal Ions Released from Fixed Orthodontic Appliance Affect Hair Mineral Content. Biol. Trace Elem. Res. 2015, 163, 11–18. [Google Scholar] [CrossRef] [PubMed]
  242. Amini, F.; Harandi, S.; Mollaei, M.; Rakhshan, V. Effects of Fixed Orthodontic Treatment Using Conventional versus Metal-Injection Molding Brackets on Salivary Nickel and Chromium Levels: A Double-Blind Randomized Clinical Trial. Eur. J. Orthod. 2015, 37, 522–530. [Google Scholar] [CrossRef] [PubMed]
  243. Amini, F.; Mollaei, M.; Harandi, S.; Rakhshan, V. Effects of Fixed Orthodontic Treatment on Hair Nickel and Chromium Levels: A 6-Month Prospective Preliminary Study. Biol. Trace Elem. Res. 2015, 164, 12–17. [Google Scholar] [CrossRef] [PubMed]
  244. Saghiri, M.A.; Orangi, J.; Asatourian, A.; Mehriar, P.; Sheibani, N. Effect of Mobile Phone Use on Metal Ion Release from Fixed Orthodontic Appliances. Am. J. Orthod. Dentofac. Orthop. 2015, 147, 719–724. [Google Scholar] [CrossRef] [PubMed]
  245. Gómez Arcila, V.; Fang Mercado, L.; Herrera Herrera, A.; Díaz Caballero, A. Bioacumulación de Níquel En Encía, Saliva y Hueso Alveolar de Paciente Con Aparatología Ortodóncica Fija: Reporte de Un Caso. Rev. Clín. Periodoncia Implantol. Rehabil. Oral 2015, 8, 163–166. [Google Scholar] [CrossRef]
  246. Gölz, L.; Knickenberg, A.C.; Keilig, L.; Reimann, S.; Papageorgiou, S.N.; Jäger, A.; Bourauel, C. Nickel Ion Concentrations in the Saliva of Patients Treated with Self-Ligating Fixed Appliances: A Prospective Cohort Study. J. Orofac. Orthop. Fortschritte Kieferorthopädie 2016, 77, 85–93. [Google Scholar] [CrossRef]
  247. Amini, F.; Shariati, M.; Sobouti, F.; Rakhshan, V. Effects of Fixed Orthodontic Treatment on Nickel and Chromium Levels in Gingival Crevicular Fluid as a Novel Systemic Biomarker of Trace Elements: A Longitudinal Study. Am. J. Orthod. Dentofac. Orthop. 2016, 149, 666–672. [Google Scholar] [CrossRef]
  248. Kumar, R.V.; Rajvikram, N.; Rajakumar, P.; Saravanan, R.; Deepak, V.A. An Accurate Methodology to Detect Leaching of Nickel and Chromium Ions in the Initial Phase of Orthodontic Treatment: An in Vivo Study. J. Contemp. Dent. Pract. 2016, 17, 205–210. [Google Scholar] [CrossRef]
  249. Wołowiec, P.; Chojnacka, K.; Loster, B.W.; Mikulewicz, M. Do Dietary Habits Influence Trace Elements Release from Fixed Orthodontic Appliances? Biol. Trace Elem. Res. 2017, 180, 214–222. [Google Scholar] [CrossRef]
  250. Causado-Vitola, V.; Rumbo-Zubiría, M.; Fang, L.; Diaz-Caballero, A. Nickel Variation in Biofilm, Saliva and Buccal Mucosa During Orthodontic Treatment. Asian J. Appl. Sci. 2017, 10, 45–49. [Google Scholar] [CrossRef]
  251. Khaneh Masjedi, M.; Haghighat Jahromi, N.; Niknam, O.; Hormozi, E.; Rakhshan, V. Effects of Fixed Orthodontic Treatment Using Conventional (Two-Piece) versus Metal Injection Moulding Brackets on Hair Nickel and Chromium Levels: A Double-Blind Randomized Clinical Trial. Eur. J. Orthod. 2017, 39, 17–24. [Google Scholar] [CrossRef] [PubMed]
  252. Buczko, P.; Knaś, M.; Grycz, M.; Szarmach, I.; Zalewska, A. Orthodontic Treatment Modifies the Oxidant–Antioxidant Balance in Saliva of Clinically Healthy Subjects. Adv. Med. Sci. 2017, 62, 129–135. [Google Scholar] [CrossRef] [PubMed]
  253. Lages, R.B.; Bridi, E.C.; Pérez, C.A.; Basting, R.T. Salivary Levels of Nickel, Chromium, Iron, and Copper in Patients Treated with Metal or Esthetic Fixed Orthodontic Appliances: A Retrospective Cohort Study. J. Trace Elem. Med. Biol. 2017, 40, 67–71, Corrigendum in J. Trace Elem. Med. Biol. 2018, 45, 189–190. [Google Scholar] [CrossRef] [PubMed]
  254. Nanjannawar, L.G.; Girme, T.S.; Agrawal, J.M.; Agrawal, M.S.; Fulari, S.G.; Shett, S.S.; Kagi, V.A. Effect of Mobile Phone Usage OnNickel Ions Release and PH of Saliva InPatients Undergoing Fixed OrthodonticTreatment. J. Clin. Diagn. Res. 2017, 11, ZC84–ZC87. [Google Scholar] [CrossRef] [PubMed]
  255. Bhasin, V.; Pustake, S.J.; Joshi, V.; Tiwari, A.; Bhasin, M.; Punia, R.S. Assessment of Changes in Nickel and Chromium Levels in the Gingival Crevicular Fluid during Fixed Orthodontic Treatment. J. Contemp. Dent. Pract. 2017, 18, 675–678. [Google Scholar] [CrossRef] [PubMed]
  256. Jamshidi, S.; Rahmati Kamel, M.; Mirzaie, M.; Sarrafan, A.; Khafri, S.; Parsian, H. Evaluation of Scalp Hair Nickel and Chromium Level Changes in Patients with Fixed Orthodontic Appliance: A One-Year Follow-up Study. Acta Odontol. Scand. 2018, 76, 1–5. [Google Scholar] [CrossRef]
  257. Jurela, A.; Verzak, Ž.; Brailo, V.; Škrinjar, I.; Sudarević, K.; Janković, B. Salivary Electrolytes in Patients with Metallic and Ceramic Orthodontic Brackets. Acta Stomatol. Croat. 2018, 52, 32–36. [Google Scholar] [CrossRef]
  258. Amini, F.; Asadi, E.; Hakimpour, D.; Rakhshan, A. Salivary Nickel and Chromium Levels in Orthodontic Patients with and Without Periodontitis: A Preliminary Historical Cohort Study. Biol. Trace Elem. Res. 2019, 191, 10–15. [Google Scholar] [CrossRef]
  259. Chitra, P.; Prashantha, G.S.; Rao, A. Long-Term Evaluation of Metal Ion Release in Orthodontic Patients Using Fluoridated Oral Hygiene Agents: An in Vivo Study. J. World Fed. Orthod. 2019, 8, 107–111. [Google Scholar] [CrossRef]
  260. Quadras, D.D.; Nayak, U.K.; Kumari, N.S.; Priyadarshini, H.R.; Gowda, S.; Fernandes, B. In Vivo Study on the Release of Nickel, Chromium, and Zinc in Saliva and Serum from Patients Treated with Fixed Orthodontic Appliances. Dent. Res. J. 2019, 16, 209. [Google Scholar] [CrossRef]
  261. Moghadam, M.G.; Hoshyar, R.; Mikulewicz, M.; Chojnacka, K.; Bjørklund, G.; Pen, J.J.; Azadi, N.A.; Pirsaheb, M.; Dashtaki, M.; Mansouri, B. Biomonitorization of Metal Ions in the Serum of Iranian Patients Treated with Fixed Orthodontic Appliances in Comparison with Controls in Eastern Iran. Environ. Sci. Pollut. Res. 2019, 26, 33373–33386. [Google Scholar] [CrossRef] [PubMed]
  262. Butt, M.; Mengal, N.; Ahmed, M. A Comparison of Nickel Release in Saliva between Orthodontic and Non-Orthodontic Patients. Pak. Armed Forces Med. J. 2020, 70, 328–332. [Google Scholar]
  263. de Souza Schacher, H.R.; de Menezes, L.M. Metal Ion Quantification in the Saliva of Patients with Lingual Arch Appliances Using Silver Solder, Laser, or TIG Welding. Clin. Oral Investig. 2020, 24, 2109–2120. [Google Scholar] [CrossRef] [PubMed]
  264. Pritam, A.; Priyadarshini, A.; Hussain, K.; Kumar, A.; Kumar, N.; Malakar, A. Assessment of Nickel and Chromium Level in Gingival Crevicular Fluid in Patients Undergoing Orthodontic Treatment with or without Fluoridated Tooth Paste. J. Pharm. Bioallied Sci. 2021, 13, 1588. [Google Scholar] [CrossRef] [PubMed]
  265. Campos Zeffa, A.; Dias, B.G.; Silva, D.C.M.S.; Rotta, L.O.; Jussiani, E.I.; Andrello, A.C.; de Paula Ramos, S. Influence of Conventional or Invisalign Orthodontic Treatment on Mineral and Trace Element Salivary Levels: Longitudinal Study with Total Reflection X-Ray Fluorescence. Biol. Trace Elem. Res. 2021, 199, 2565–2572. [Google Scholar] [CrossRef] [PubMed]
  266. Hamadamin, S.I. In Vivo Kinetic Release of Five Metal Ions (Iron, Titanium, Nickel, Copper, and Chromium) from Fixed Orthodontic Alloys in Erbil City-Kurdistan Region/Iraq. Environ. Sci. Pollut. Res. 2022, 29, 11730–11735. [Google Scholar] [CrossRef] [PubMed]
  267. Rajendran, R.; Venkatachalapathy, S.; Thiyagarajan, B.; Jeevagan, S.; Chinnasamy, A.; Sivanandham, M. Effect of Mobile Phone with and without Earphones Usage on Nickel Ion Release from Fixed Orthodontic Appliance. J. Contemp. Dent. Pract. 2023, 24, 303–307. [Google Scholar] [CrossRef]
  268. Vaughan, J.L.; Duncanson, M.G.; Nanda, R.S.; Currier, G.F. Relative Kinetic Frictional Forces between Sintered Stainless Steel Brackets and Orthodontic Wires. Am. J. Orthod. Dentofac. Orthop. 1995, 107, 20–27. [Google Scholar] [CrossRef]
  269. Dridi, A.; Bensalah, W.; Mezlini, S.; Tobji, S.; Zidi, M. Influence of Bio-Lubricants on the Orthodontic Friction. J. Mech. Behav. Biomed. Mater. 2016, 60, 1–7. [Google Scholar] [CrossRef]
  270. Fidalgo, T.K.D.S.; Pithon, M.M.; Maciel, J.V.B.; Bolognese, A.M. Friction between Different Wire Bracket Combinations in Artificial Saliva—An in Vitro Evaluation. J. Appl. Oral Sci. 2011, 19, 57–62. [Google Scholar] [CrossRef] [PubMed]
  271. Ranc, H.; Elkhyat, A.; Servais, C.; Mac-Mary, S.; Launay, B.; Humbert, P. Friction Coefficient and Wettability of Oral Mucosal Tissue: Changes Induced by a Salivary Layer. Colloids Surf. A Physicochem. Eng. Asp. 2006, 276, 155–161. [Google Scholar] [CrossRef]
  272. Lu, C.; Zheng, Y.; Zhong, Q. Corrosion of Dental Alloys in Artificial Saliva with Streptococcus Mutans. PLoS ONE 2017, 12, e0174440. [Google Scholar] [CrossRef] [PubMed]
  273. Reichardt, E.; Geraci, J.; Sachse, S.; Rödel, J.; Pfister, W.; Löffler, B.; Wagner, Y.; Eigenthaler, M.; Wolf, M. Qualitative and Quantitative Changes in the Oral Bacterial Flora Occur Shortly after Implementation of Fixed Orthodontic Appliances. Am. J. Orthod. Dentofac. Orthop. 2019, 156, 735–744. [Google Scholar] [CrossRef] [PubMed]
  274. Mei, L.; Chieng, J.; Wong, C.; Benic, G.; Farella, M. Factors Affecting Dental Biofilm in Patients Wearing Fixed Orthodontic Appliances. Prog. Orthod. 2017, 18, 4. [Google Scholar] [CrossRef] [PubMed]
  275. Alavi, S.; Yaraghi, N. The Effect of Fluoride Varnish and Chlorhexidine Gel on White Spots and Gingival and Plaque Indices in Fixed Orthodontic Patients: A Placebo-Controlled Study. Dent. Res. J. 2018, 15, 276–282. [Google Scholar]
  276. Chitra, P.; Prashantha, G.S.; Rao, A. Effect of Fluoride Agents on Surface Characteristics of NiTi Wires. An Ex Vivo Investigation. J. Oral Biol. Craniofacial Res. 2020, 10, 435–440. [Google Scholar] [CrossRef] [PubMed]
  277. Schiff, N.; Grosgogeat, B.; Lissac, M.; Dalard, F. Influence of Fluoridated Mouthwashes on Corrosion Resistance of Orthodontics Wires. Biomaterials 2004, 25, 4535–4542. [Google Scholar] [CrossRef]
  278. Nahidh, M.; MH Garma, N.; Jasim, E.S. Assessment of Ions Released from Three Types of Orthodontic Brackets Immersed in Different Mouthwashes: An in Vitro Study. J. Contemp. Dent. Pract. 2018, 19, 73–80. [Google Scholar] [CrossRef]
  279. Rincic Mlinaric, M.; Karlovic, S.; Ciganj, Z.; Acev, D.P.; Pavlic, A.; Spalj, S. Oral Antiseptics and Nickel–Titanium Alloys: Mechanical and Chemical Effects of Interaction. Odontology 2019, 107, 150–157. [Google Scholar] [CrossRef]
  280. Condò, R.; Carli, E.; Cioffi, A.; Cataldi, M.E.; Quinzi, V.; Casaglia, A.; Giancotti, A.; Pirelli, P.; Lucarini, I.; Maita, F.; et al. Fluorinated Agents Effects on Orthodontic Alloys: A Descriptive In Vitro Study. Materials 2022, 15, 4612. [Google Scholar] [CrossRef] [PubMed]
  281. Pastor, F.; Rodríguez, J.C.; Barrera, J.M.; Delgado García-Menocal, J.A.; Brizuela, A.; Puigdollers, A.; Espinar, E.; Gil, J. Effect of Fluoride Content of Mouthwashes on Superelastic Properties of NiTi Orthodontic Archwires. Materials 2022, 15, 6592. [Google Scholar] [CrossRef] [PubMed]
  282. Sahoo, N.; Bhuyan, L.; Dhull, K.S.; Dash, K.C.; MD, I.; Mishra, P. In Vitro Effect Of Fluoride Prophylactic Agents On Titanium Molybdenum Alloy And Stainless Steel Orthodontic Wires—Scanning Electron Microscope Study. Bangladesh J. Med. Sci. 2023, 22, 47–51. [Google Scholar] [CrossRef]
  283. Mirhashemi, A.; Jahangiri, S.; Kharrazifard, M. Release of Nickel and Chromium Ions from Orthodontic Wires Following the Use of Teeth Whitening Mouthwashes. Prog. Orthod. 2018, 19, 4. [Google Scholar] [CrossRef]
  284. Yanisarapan, T.; Thunyakitpisal, P.; Chantarawaratit, P. Corrosion of Metal Orthodontic Brackets and Archwires Caused by Fluoride-Containing Products: Cytotoxicity, Metal Ion Release and Surface Roughness. Orthod. Waves 2018, 77, 79–89. [Google Scholar] [CrossRef]
  285. Abbassy, M. Fluoride Influences Nickel-Titanium Orthodontic Wires′ Surface Texture and Friction Resistance. J. Orthod. Sci. 2016, 5, 121. [Google Scholar] [CrossRef]
  286. Rajendran, A.; Sundareswaran, S.; Peediyekkal, L.; Santhakumar, P.; Sathyanadhan, S. Effect of Oral Environment and Prescribed Fluoride Mouthwashes on Different Types of TMA Wires—An in-Vivo Study. J. Orthod. Sci. 2019, 8, 8. [Google Scholar] [CrossRef]
  287. Li, X.; Wang, J.; Han, E.; Ke, W. Influence of Fluoride and Chloride on Corrosion Behavior of NiTi Orthodontic Wires. Acta Biomater. 2007, 3, 807–815. [Google Scholar] [CrossRef]
  288. Brandão, G.A.M.; Simas, R.M.; de Almeida, L.M.; da Silva, J.M.; Meneghim, M.d.C.; Pereira, A.C.; de Almeida, H.A.; Brandão, A.M.M. Evaluation of Ionic Degradation and Slot Corrosion of Metallic Brackets by the Action of Different Dentifrices. Dent. Press J. Orthod. 2013, 18, 86–93. [Google Scholar] [CrossRef]
  289. Chantarawaratit, P.; Yanisarapan, T. Exposure to the Oral Environment Enhances the Corrosion of Metal Orthodontic Appliances Caused by Fluoride-Containing Products: Cytotoxicity, Metal Ion Release, and Surface Roughness. Am. J. Orthod. Dentofac. Orthop. 2021, 160, 101–112. [Google Scholar] [CrossRef]
  290. Russell, J.S. Current Products and Practice: Aesthetic Orthodontic Brackets. J. Orthod. 2005, 32, 146–163. [Google Scholar] [CrossRef] [PubMed]
  291. Ali, O.; Makou, M.; Papadopoulos, T.; Eliades, G. Laboratory Evaluation of Modern Plastic Brackets. Eur. J. Orthod. 2012, 34, 595–602. [Google Scholar] [CrossRef] [PubMed]
  292. Cacciafesta, V.; Sfondrini, M.F.; Scribante, A.; Klersy, C.; Auricchio, F. Evaluation of Friction of Conventional and Metal-Insert Ceramic Brackets in Various Bracket-Archwire Combinations. Am. J. Orthod. Dentofac. Orthop. 2003, 124, 403–409. [Google Scholar] [CrossRef] [PubMed]
  293. Arici, N.; Akdeniz, B.S.; Arici, S. Comparison of the Frictional Characteristics of Aesthetic Orthodontic Brackets Measured Using a Modified in Vitro Technique. Korean J. Orthod. 2015, 45, 29. [Google Scholar] [CrossRef] [PubMed]
  294. Varela, J.C.; Velo, M.; Espinar, E.; Llamas, J.M.; Rúperez, E.; Manero, J.M.; Javier Gil, F. Mechanical Properties of a New Thermoplastic Polymer Orthodontic Archwire. Mater. Sci. Eng. C 2014, 42, 1–6. [Google Scholar] [CrossRef]
  295. Burstone, C.J.; Liebler, S.A.H.; Goldberg, A.J. Polyphenylene Polymers as Esthetic Orthodontic Archwires. Am. J. Orthod. Dentofac. Orthop. 2011, 139, e391–e398. [Google Scholar] [CrossRef]
  296. Tada, Y.; Hayakawa, T.; Nakamura, Y. Load-Deflection and Friction Properties of PEEK Wires as Alternative Orthodontic Wires. Materials 2017, 10, 914. [Google Scholar] [CrossRef]
  297. Spendlove, J.; Berzins, D.W.; Pruszynski, J.E.; Ballard, R.W. Investigation of Force Decay in Aesthetic, Fibre-Reinforced Composite Orthodontic Archwires. Eur. J. Orthod. 2015, 37, 43–48. [Google Scholar] [CrossRef]
  298. Chng, C.K.; Foong, K.; Gandedkar, N.H.; Chan, Y.H.; Chew, C.-L. A New Esthetic Fiber-Reinforced Polymer Composite Resin Archwire: A Comparative Atomic Force Microscope (AFM) and Field-Emission Scanning Electron Microscope (FESEM) Study. Prog. Orthod. 2014, 15, 39. [Google Scholar] [CrossRef]
  299. Mikulewicz, M.; Gronostajski, Z.; Wielgus, A.; Chojnacka, K. Transparent Orthodontic Archwires: A Systematic Literature Review. Arch. Civ. Mech. Eng. 2017, 17, 651–657. [Google Scholar] [CrossRef]
  300. Zinelis, S.; Brantley, W. Structure/Property Relationships in Orthodontic Ceramics. In Orthodontic Applications of Biomaterials; Eliades, T., Brantley, W.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 61–71. [Google Scholar]
  301. Krauss, J.; Faltermeier, A.; Behr, M.; Proff, P. Evaluation of Alternative Polymer Bracket Materials. Am. J. Orthod. Dentofac. Orthop. 2010, 137, 362–367. [Google Scholar] [CrossRef] [PubMed]
  302. Faltermeier, A.; Behr, M.; Mussig, D. In Vitro Colour Stability of Aesthetic Brackets. Eur. J. Orthod. 2007, 29, 354–358. [Google Scholar] [CrossRef] [PubMed]
  303. Faltermeier, A.; Reicheneder, C.; Römer, P.; Castro-Laza, A.; Proff, P. Effect of Ionizing Radiation on Polymer Brackets. J. Orofac. Orthop. 2014, 75, 334–344. [Google Scholar] [CrossRef] [PubMed]
  304. Retamoso, L.B.; Luz, T.B.; Marinowic, D.R.; Machado, D.C.; De Menezes, L.M.; Freitas, M.P.M.; Oshima, H.M.S. Cytotoxicity of Esthetic, Metallic, and Nickel-Free Orthodontic Brackets: Cellular Behavior and Viability. Am. J. Orthod. Dentofac. Orthop. 2012, 142, 70–74. [Google Scholar] [CrossRef] [PubMed]
  305. Guignone, B.C.; Silva, L.K.; Soares, R.V.; Akaki, E.; Goiato, M.C.; Pithon, M.M.; Oliveira, D.D. Color Stability of Ceramic Brackets Immersed in Potentially Staining Solutions. Dent. Press J. Orthod. 2015, 20, 32–38. [Google Scholar] [CrossRef] [PubMed]
  306. Johnson, G.; Walker, M.P.; Kula, K. Fracture Strength of Ceramic Bracket Tie Wings Subjected to Tension. Angle Orthod. 2005, 75, 95–100. [Google Scholar] [CrossRef] [PubMed]
  307. Sanchez, D.J.; Walker, M.P.; Kula, K.; Williams, K.B.; Eick, J.D. Fluoride Prophylactic Agents Effect on Ceramic Bracket Tie-Wing Fracture Strength. Angle Orthod. 2008, 78, 524–530. [Google Scholar] [CrossRef]
  308. Guerrero, A.P.; Guariza Filho, O.; Tanaka, O.; Camargo, E.S.; Vieira, S. Evaluation of Frictional Forces between Ceramic Brackets and Archwires of Different Alloys Compared with Metal Brackets. Braz. Oral Res. 2010, 24, 40–45. [Google Scholar] [CrossRef]
  309. Sahoo, N. Comparison of the Perception of Pain during Fixed Orthodontic Treatment with Metal and Ceramic Brackets. J. Pharm. Bioallied Sci. 2019, 11, 30. [Google Scholar] [CrossRef]
  310. Matias, M.; Freitas, M.R.D.; Freitas, K.M.S.D.; Janson, G.; Higa, R.H.; Francisconi, M.F. Comparison of Deflection Forces of Esthetic Archwires Combined with Ceramic Brackets. J. Appl. Oral Sci. 2018, 26, e20170220. [Google Scholar] [CrossRef]
  311. da Silva, D.L.; Mattos, C.T.; de Araújo, M.V.A.; de Oliveira Ruellas, A.C. Color Stability and Fluorescence of Different Orthodontic Esthetic Archwires. Angle Orthod. 2013, 83, 127–132. [Google Scholar] [CrossRef] [PubMed]
  312. Ortiz, A.J.; Fernández, E.; Vicente, A.; Calvo, J.L.; Ortiz, C. Metallic Ions Released from Stainless Steel, Nickel-Free, and Titanium Orthodontic Alloys: Toxicity and DNA Damage. Am. J. Orthod. Dentofac. Orthop. 2011, 140, e115–e122. [Google Scholar] [CrossRef]
  313. Pazzini, C.A.; Marques, L.S.; Pereira, L.J.; Corrêa-Faria, P.; Paiva, S.M. Allergic Reactions and Nickel-Free Braces: A Systematic Review. Braz. Oral Res. 2011, 25, 85–90. [Google Scholar] [CrossRef] [PubMed]
  314. Pantuzo, M.C.G.; Zenóbio, E.G.; Marigo, H.D.A.; Zenóbio, M.A.F. Hypersensitivity to Conventional and to Nickel-Free Orthodontic Brackets. Braz. Oral Res. 2007, 21, 298–302. [Google Scholar] [CrossRef] [PubMed]
  315. Zheng, Y.F.; Zhang, B.B.; Wang, B.L.; Wang, Y.B.; Li, L.; Yang, Q.B.; Cui, L.S. Introduction of Antibacterial Function into Biomedical TiNi Shape Memory Alloy by the Addition of Element Ag. Acta Biomater. 2011, 7, 2758–2767. [Google Scholar] [CrossRef] [PubMed]
  316. Baigonakova, G.; Marchenko, E.; Chekalkin, T.; Kang, J.; Weiss, S.; Obrosov, A. Influence of Silver Addition on Structure, Martensite Transformations and Mechanical Properties of TiNi–Ag Alloy Wires for Biomedical Application. Materials 2020, 13, 4721. [Google Scholar] [CrossRef]
  317. Oh, K.-T.; Joo, U.-H.; Park, G.-H.; Hwang, C.-J.; Kim, K.-N. Effect of Silver Addition on the Properties of Nickel-Titanium Alloys for Dental Application. J. Biomed. Mater. Res. Part B Appl. Biomater. 2006, 76, 306–314. [Google Scholar] [CrossRef]
  318. He, L.; Cui, Y.; Zhang, C. The Corrosion Resistance, Cytotoxicity, and Antibacterial Properties of Lysozyme Coatings on Orthodontic Composite Arch Wires. RSC Adv. 2020, 10, 18131–18137. [Google Scholar] [CrossRef]
  319. Iijima, M.; Zinelis, S.; Papageorgiou, S.N.; Brantley, W.; Eliades, T. Orthodontic Brackets. In Orthodontic Applications of Biomaterials; Elsevier: Amsterdam, The Netherlands, 2017; pp. 75–96. ISBN 9780081003831. [Google Scholar]
  320. Ntasi, A.; Al Jabbari, Y.; Mueller, W.D.; Eliades, G.; Zinelis, S. Metallurgical and Electrochemical Characterization of Contemporary Silver-Based Soldering Alloys. Angle Orthod. 2014, 84, 508–515. [Google Scholar] [CrossRef]
  321. Freitas, M.P.M.; Oshima, H.M.S.; Menezes, L.M.; Machado, D.C.; Viezzer, C. Cytotoxicity of Silver Solder Employed in Orthodontics. Angle Orthod. 2009, 79, 939–944. [Google Scholar] [CrossRef]
  322. Erdogan, A.T.; Nalbantgil, D.; Ulkur, F.; Sahin, F. Metal Ion Release from Silver Soldering and Laser Welding Caused by Different Types of Mouthwash. Angle Orthod. 2015, 85, 665–672. [Google Scholar] [CrossRef] [PubMed]
  323. da Costa, É.C.S.C.; Neves, J.G.; Borges, L.P.S.; Tsuzuki, F.M.; Correr, A.B.; Correr-Sobrinho, L.; Costa, A.R. Comparison of the Physico-Chemical Impact of Chlorhexidine and Silver Nanoparticles on Orthodontic Appliances Made with Laser and Silver Solder: An in Vitro Study. Int. Orthod. 2022, 20, 100631. [Google Scholar] [CrossRef] [PubMed]
  324. Rodrigues, P.F.; Fernandes, F.M.B.; Magalhães, R.; Camacho, E.; Lopes, A.; Paula, A.S.; Basu, R.; Schell, N. Thermo-Mechanical Characterization of NiTi Orthodontic Archwires with Graded Actuating Forces. J. Mech. Behav. Biomed. Mater. 2020, 107, 103747. [Google Scholar] [CrossRef] [PubMed]
  325. Javaid, M.; Haleem, A. Current Status and Applications of Additive Manufacturing in Dentistry: A Literature-Based Review. J. Oral Biol. Craniofacial Res. 2019, 9, 179–185. [Google Scholar] [CrossRef] [PubMed]
  326. Panayi, N.C. 3D Printing of in Office Custom-Made Brackets: Concept, Design, Production and Evidence. Semin. Orthod. 2023, 29, 11–16. [Google Scholar] [CrossRef]
  327. Graf, S.; Thakkar, D.; Hansa, I.; Muthuswamy Pandian, S.; Adel, S.M. 3D Metal Printing in Orthodontics: Current Trends, Biomaterials, Workflows and Clinical Implications. Semin. Orthod. 2023, 29, 34–42. [Google Scholar] [CrossRef]
  328. Abalos, C.; Paúl, A.; Mendoza, A.; Solano, E.; Gil, F.J. Influence of Topographical Features on the Fluoride Corrosion of Ni–Ti Orthodontic Archwires. J. Mater. Sci. Mater. Med. 2011, 22, 2813–2821. [Google Scholar] [CrossRef]
  329. Wichelhaus, A.; Geserick, M.; Hibst, R.; Sander, F.G. The Effect of Surface Treatment and Clinical Use on Friction in NiTi Orthodontic Wires. Dent. Mater. 2005, 21, 938–945. [Google Scholar] [CrossRef]
  330. Neumann, P.; Bourauel, C.; Jäger, A. Corrosion and Permanent Fracture Resistance of Coated and Conventional Orthodontic Wires. J. Mater. Sci. Mater. Med. 2002, 13, 141–147. [Google Scholar] [CrossRef]
  331. Chu, P.; Chen, J.Y.; Wang, L.P.; Huang, N. Plasma-Surface Modification of Biomaterials. Mater. Sci. Eng. R Rep. 2002, 36, 143–206. [Google Scholar] [CrossRef]
  332. Katic, V.; Curkovic, L.; Bosnjak, M.; Peros, K.; Mandic, D.; Spals, S. Effect of PH, Fluoride and Hydrofluoric Acid Concentration on Ion Release from NiTi Wires with Various Coatings. Dent. Mater. J. 2017, 36, 149–156. [Google Scholar] [CrossRef] [PubMed]
  333. de Albuquerque, C.G.; Correr, A.B.; Venezian, G.C.; Santamaria Jr, M.; Tubel, C.A.; Vedovello, S.A.S. Deflection and Flexural Strength Effects on the Roughness of Aesthetic-Coated Orthodontic Wires. Braz. Dent. J. 2017, 28, 40–45. [Google Scholar] [CrossRef] [PubMed]
  334. Iijima, M.; Muguruma, T.; Brantley, W.; Choe, H.-C.; Nakagaki, S.; Alapati, S.B.; Mizoguchi, I. Effect of Coating on Properties of Esthetic Orthodontic Nickel-Titanium Wires. Angle Orthod. 2012, 82, 319–325. [Google Scholar] [CrossRef] [PubMed]
  335. Alsanea, J.; Al Shehri, H. Evaluation of Nanomechanical Properties, Surface Roughness, and Color Stability of Esthetic Nickel-Titanium Orthodontic Archwires. J. Int. Soc. Prev. Community Dent. 2019, 9, 33. [Google Scholar] [CrossRef]
  336. Pinzan-Vercelino, C.R.M.; de Araújo Gurgel, J.; Bramante, F.S.; Pessoa, T.F.; Albertin, S.A.; Fialho, M.P.N.; Pinzan, A. Coated Nickel Titanium Archwires and Their Uncoated Counterparts: In-Vitro Contemporary Comparison of Unloading Forces. Int. Orthod. 2020, 18, 374–379. [Google Scholar] [CrossRef]
  337. Batista, C.H.; Neves, J.G.; Terossi de Godoi, A.P.; Veroni Degan, V.; Custódio, W.; Furletti, V.; Vedovello Filho, M. Comparison of Force Delivery of Thermally Activated Aesthetic and Non-Aesthetic Ni-Ti Wires: An in-Vitro Study. Int. Orthod. 2020, 18, 359–365. [Google Scholar] [CrossRef]
  338. Jusufi Osmani, Z.; Poljšak, B.; Zelenika, S.; Kamenar, E.; Marković, K.; Perčić, M.; Katić, V. Ion Release and Surface Changes of Nickel–Titanium Archwires Induced by Changes in the PH Value of the Saliva—Significance for Human Health Risk Assessment. Materials 2022, 15, 1994. [Google Scholar] [CrossRef]
  339. Katić, V.; Ćurković, L.; Ujević Bošnjak, M.; Špalj, S. Determination of Corrosion Rate of Orthodontic Wires Based on Nickel-Titanium Alloy in Artificial Saliva. Materwiss. Werksttech. 2014, 45, 99–105. [Google Scholar] [CrossRef]
  340. Katić, V.; Ćurković, H.O.; Semenski, D.; Baršić, G.; Marušić, K.; Špalj, S. Influence of Surface Layer on Mechanical and Corrosion Properties of Nickel-Titanium Orthodontic Wires. Angle Orthod. 2014, 84, 1041–1048. [Google Scholar] [CrossRef]
  341. Katić, V.; Buljan, Z.I.; Špalj, S.; Ćurković, H.O. Corrosion Behavior of Coated and Uncoated Nickel-Titanium Orthodontic Wires in Artificial Saliva with Short-Term Prophylactic Fluoride Treatment. Int. J. Electrochem. Sci. 2018, 13, 4160–4170. [Google Scholar] [CrossRef]
  342. de Amorim, M.C.; da Rocha Gomes, S.; da Silva, B.P.; Aoki, I.V.; Basting, R.T. Surface Micromorphology, Ion Release and Resistance to Corrosion of Orthodontic Wires Aesthetic Coating Subject to Degradation. J. Bio-Tribo-Corros. 2022, 8, 22. [Google Scholar] [CrossRef]
  343. Nsaif, Y.A.; Mahmood, A.B. Effect of Fluoride Agent on the Load Deflection of Rhodium-Coated Arch Wires; An In-Vitro Study. Indian J. Public Health Res. Dev. 2019, 10, 823. [Google Scholar] [CrossRef]
  344. Musa Trolić, I.; Turco, G.; Contardo, L.; Serdarević, N.L.; Ćurković, H.O.; Špalj, S. Corrosion of Nickel-Titanium Orthodontic Archwires in Saliva and Oral Probiotic Supplements. Acta Stomatol. Croat. 2017, 51, 316–325. [Google Scholar] [CrossRef] [PubMed]
  345. Musa Trolic, I.; Serdarevic, N.L.; Todoric, Z.; Budimir, A.; Spalj, S.; Otmacic Curkovic, H. Corrosion of Orthodontic Archwires in Artificial Saliva in the Presence of Lactobacillus Reuteri. Surf. Coat. Technol. 2019, 370, 44–52. [Google Scholar] [CrossRef]
  346. Musa Trolic, I.; Todoric, Z.; Pop Acev, D.; Makreski, P.; Pejova, B.; Spalj, S. Effects of the Presence of Probiotic Bacteria in the Aging Medium on the Surface Roughness and Chemical Composition of Two Dental Alloys. Microsc. Res. Tech. 2019, 82, 1384–1391. [Google Scholar] [CrossRef]
  347. Costa Lima, K.C.; Benini Paschoal, M.A.; de Araújo Gurgel, J.; Salvatore Freitas, K.M.; Maio Pinzan-Vercelino, C.R. Comparative Analysis of Microorganism Adhesion on Coated, Partially Coated, and Uncoated Orthodontic Archwires: A Prospective Clinical Study. Am. J. Orthod. Dentofac. Orthop. 2019, 156, 611–616. [Google Scholar] [CrossRef]
  348. Usui, T.; Iwata, T.; Miyake, S.; Otsuka, T.; Koizumi, S.; Shirakawa, N.; Kawata, T. Mechanical and Frictional Properties of Aesthetic Orthodontic Wires Obtained by Hard Chrome Carbide Plating. J. Dent. Sci. 2018, 13, 151–159. [Google Scholar] [CrossRef]
  349. Albawardi, A.; Warunek, S.; Makowka, S.; Al-Jewair, T. Friction Forces Generated by Aesthetic Gummetal® (Ti-Nb) Orthodontic Archwires: A Comparative in Vitro Study. Int. Orthod. 2022, 20, 100683. [Google Scholar] [CrossRef]
  350. Ramasamy, M.; Prabhakar, R.; Thirunavukarasu, R.; Saravanan, R.; Rajvikram, N.; Sowndarya, V. Evalution of Color Stability of Coated Aesthetic Arch Wires. Eur. J. Mol. Clin. Med. 2020, 07, 5194–5200. [Google Scholar]
  351. Cowley, A.; Woodward, B. A Healthy Future: Platinum in Medical Applications. Platin. Met. Rev. 2011, 55, 98–107. [Google Scholar] [CrossRef]
  352. Johnson, A.; Shiraishi, T. Biocompatibility of Precious Metals for Medical Applications. In Precious Metals for Biomedical Applications; Elsevier: Amsterdam, The Netherlands, 2014; pp. 37–55. ISBN 9780857094346. [Google Scholar]
  353. Khonsari, S.K.; Towhidi, N.; Cheraghi, M.S.; Allahkaram, S.R.; Rabizadeh, T. Pt Nanoparticles Coating on Orthodontic Ni-Ti Wires Using Pulse Current. In Proceedings of the 2011 IEEE Nanotechnology Materials and Devices Conference, Jeju, Republic of Korea, 18–21 October 2011; pp. 502–505. [Google Scholar]
  354. Sim, W.; Barnard, R.; Blaskovich, M.A.T.; Ziora, Z. Antimicrobial Silver in Medicinal and Consumer Applications: A Patent Review of the Past Decade (2007–2017). Antibiotics 2018, 7, 93. [Google Scholar] [CrossRef] [PubMed]
  355. Talapko, J.; Matijević, T.; Juzbašić, M.; Antolović-Požgain, A.; Škrlec, I. Antibacterial Activity of Silver and Its Application in Dentistry, Cardiology and Dermatology. Microorganisms 2020, 8, 1400. [Google Scholar] [CrossRef] [PubMed]
  356. Metin-Gürsoy, G.; Taner, L.; Barış, E. Biocompatibility of Nanosilver-Coated Orthodontic Brackets: An in Vivo Study. Prog. Orthod. 2016, 17, 39. [Google Scholar] [CrossRef] [PubMed]
  357. Mhaske, A.R.; Shetty, P.C.; Bhat, N.S.; Ramachandra, C.S.; Laxmikanth, S.M.; Nagarahalli, K.; Tekale, P.D. Antiadherent and Antibacterial Properties of Stainless Steel and NiTi Orthodontic Wires Coated with Silver against Lactobacillus Acidophilus—An in Vitro Study. Prog. Orthod. 2015, 16, 40. [Google Scholar] [CrossRef]
  358. Ghasemi, T.; Arash, V.; Rabiee, S.M.; Rajabnia, R.; Pourzare, A.; Rakhshan, V. Antimicrobial Effect, Frictional Resistance, and Surface Roughness of Stainless Steel Orthodontic Brackets Coated with Nanofilms of Silver and Titanium Oxide: A Preliminary Study. Microsc. Res. Tech. 2017, 80, 599–607. [Google Scholar] [CrossRef]
  359. Meyer-Kobbe, V.; Doll, K.; Stiesch, M.; Schwestka-Polly, R.; Demling, A. Comparison of Intraoral Biofilm Reduction on Silver-Coated and Silver Ion-Implanted Stainless Steel Bracket Material. J. Orofac. Orthop. Fortschritte Kieferorthopädie 2019, 80, 32–43. [Google Scholar] [CrossRef]
  360. Metin-Gürsoy, G.; Taner, L.; Akca, G. Nanosilver Coated Orthodontic Brackets: In Vivo Antibacterial Properties and Ion Release. Eur. J. Orthod. 2017, 39, 9–16. [Google Scholar] [CrossRef]
  361. Espinosa-Cristóbal, L.F.; López-Ruiz, N.; Cabada-Tarín, D.; Reyes-López, S.Y.; Zaragoza-Contreras, A.; Constandse-Cortéz, D.; Donohué-Cornejo, A.; Tovar-Carrillo, K.; Cuevas-González, J.C.; Kobayashi, T. Antiadherence and Antimicrobial Properties of Silver Nanoparticles against Streptococcus Mutans on Brackets and Wires Used for Orthodontic Treatments. J. Nanomater. 2018, 2018, 9248527. [Google Scholar] [CrossRef]
  362. Bindu, S.H.; Kala Vani, S.V.; Nirisha, G.; Madhuri, N.; Sai Deepa, B.; Hemadri, S. Evaluation of Antibacterial Effect of Silver Nanoparticle Coated Stainless Steel Band Material—An In Vitro Study. Orthod. J. Nepal 2019, 9, 13–19. [Google Scholar] [CrossRef]
  363. Prabha, R.D.; Kandasamy, R.; Sivaraman, U.S.; Nandkumar, M.A.; Nair, P.D. Antibacterial Nanosilver Coated Orthodontic Bands with Potential Implications in Dentistry. Indian J. Med. Res. 2016, 144, 580–586. [Google Scholar] [CrossRef]
  364. Bahrami, R.; Pourhajibagher, M.; Badiei, A.; Masaeli, R.; Tanbakuchi, B. Evaluation of the Cell Viability and Antimicrobial Effects of Orthodontic Bands Coated with Silver or Zinc Oxide Nanoparticles: An in Vitro Study. Korean J. Orthod. 2023, 53, 16–25. [Google Scholar] [CrossRef] [PubMed]
  365. Gonçalves, I.S.; Viale, A.B.; Sormani, N.N.; Pizzol, K.E.D.C.; Araujo-Nobre, A.R.D.; Oliveira, P.C.S.D.; Barud, H.G.D.O.; Antonio, S.G.; Barud, H.D.S. Antimicrobial Orthodontic Wires Coated with Silver Nanoparticles. Braz. Arch. Biol. Technol. 2020, 63, e20190339. [Google Scholar] [CrossRef]
  366. Gil, F.J.; Espinar-Escalona, E.; Clusellas, N.; Fernandez-Bozal, J.; Artes-Ribas, M.; Puigdollers, A. New Bactericide Orthodonthic Archwire: NiTi with Silver Nanoparticles. Metals 2020, 10, 702. [Google Scholar] [CrossRef]
  367. Zeidan, N.K.; Enany, N.M.; Mohamed, G.G.; Marzouk, E.S. The Antibacterial Effect of Silver, Zinc-Oxide and Combination of Silver/ Zinc Oxide Nanoparticles Coating of Orthodontic Brackets (an in Vitro Study). BMC Oral Health 2022, 22, 230. [Google Scholar] [CrossRef]
  368. Anand, B.G.; Shejale, K.P.; Rajesh Kumar, R.; Thangam, R.; Prajapati, K.P.; Kar, K.; Mala, R. Bioactivation of an Orthodontic Wire Using Multifunctional Nanomaterials to Prevent Plaque Accumulation. Biomater. Adv. 2023, 148, 213346. [Google Scholar] [CrossRef]
  369. Ryu, H.S.; Bae, I.H.; Lee, K.G.; Hwang, H.S.; Lee, K.H.; Koh, J.T.; Cho, J.H. Antibacterial Effect of Silver-Platinum Coating for Orthodontic Appliances. Angle Orthod. 2012, 82, 151–157. [Google Scholar] [CrossRef] [PubMed]
  370. Shah, P.; Sharma, P.; Goje, S. Comparative Evaluation of Frictional Resistance of Silver-Coated Stainless Steel Wires with Uncoated Stainless Steel Wires: An In Vitro Study. Contemp. Clin. Dent. 2018, 9, 331. [Google Scholar] [CrossRef]
  371. Shah, P.; Sharma, P.; Naik, H.; Patel, K.; Panchal, C. A Comparative Evaluation of Frictional Resistance and Surface Roughness of Silver Coated and Uncoated Stainless-Steel Bracket Wire Assembly- An in-Vitro Study. J. Clin. Exp. Dent. 2023, 15, e411–e419. [Google Scholar] [CrossRef]
  372. Arash, V.; Anoush, K.; Rabiee, S.M.; Rahmatei, M.; Tavanafar, S. The Effects of Silver Coating on Friction Coefficient and Shear Bond Strength of Steel Orthodontic Brackets. Scanning 2015, 37, 294–299. [Google Scholar] [CrossRef]
  373. Shirakawa, N.; Iwata, T.; Miyake, S.; Otsuka, T.; Koizumi, S.; Usui, T.; Kawata, T. Development of the Aesthetic Orthodontic Appliances Using a Silver Plating Process: The Report on Peel Resistance. Biomed. Res. 2017, 28, 3217–3221. [Google Scholar]
  374. Usui, T.; Miyake, S.; Iwata, T.; Otsuka, T.; Koizumi, S.; Shirakawa, N.; Kawata, T. Aesthetic Characteristics of the Orthodontic Wire with Silver Plating. Biomed. Res. 2017, 28, 4937–4941. [Google Scholar]
  375. Ozeki, K.; Yuhta, T.; Aoki, H.; Asaoka, T.; Daisaku, T.; Fukui, Y. Deterioration in the Superelasticity of Ti Sputter Coated on NiTi Orthodontic Wire. Biomed. Mater. Eng. 2003, 13, 355–362. [Google Scholar] [PubMed]
  376. Anuradha, P.; Varma, N.K.S.; Balakrishnan, A. Reliability Performance of Titanium Sputter Coated Ni–Ti Arch Wires: Mechanical Performance and Nickel Release Evaluation. Biomed. Mater. Eng. 2015, 26, 67–77. [Google Scholar] [CrossRef] [PubMed]
  377. Ozeki, K.; Yuhta, T.; Aoki, H.; Fukui, Y. Inhibition of Ni Release from NiTi Alloy by Hydroxyapatite, Alumina, and Titanium Sputtered Coatings. Biomed. Mater. Eng. 2003, 13, 271–279. [Google Scholar]
  378. Prashanth, L.; Kattapagari, K.; Chitturi, R.; Baddam, V.R.; Prasad, L. A Review on Role of Essential Trace Elements in Health and Disease. J. Dr. NTR Univ. Health Sci. 2015, 4, 75. [Google Scholar] [CrossRef]
  379. Tapiero, H.; Tew, K.D. Trace Elements in Human Physiology and Pathology: Zinc and Metallothioneins. Biomed. Pharmacother. 2003, 57, 399–411. [Google Scholar] [CrossRef] [PubMed]
  380. Kogan, S.; Sood, A.; Garnick, M.S. Zinc and Wound Healing: A Review of Zinc Physiology and Clinical Applications. Wounds 2017, 29, 102–106. [Google Scholar]
  381. Karandish, M.; Pakshir, M.; Moghimi, M.; Jafarpour, D. Evaluating the Mechanical Properties of Zinc-Coated Stainless Steel Orthodontic Wires Using Physical Vapor Deposition. Int. J. Dent. 2021, 2021, 6651289. [Google Scholar] [CrossRef]
  382. Mostaed, E.; Sikora-Jasinska, M.; Drelich, J.W.; Vedani, M. Zinc-Based Alloys for Degradable Vascular Stent Applications. Acta Biomater. 2018, 71, 1–23. [Google Scholar] [CrossRef]
  383. Abendrot, M.; Kalinowska-Lis, U. Zinc-Containing Compounds for Personal Care Applications. Int. J. Cosmet. Sci. 2018, 40, 319–327. [Google Scholar] [CrossRef]
  384. Hernández-Escobar, D.; Champagne, S.; Yilmazer, H.; Dikici, B.; Boehlert, C.J.; Hermawan, H. Current Status and Perspectives of Zinc-Based Absorbable Alloys for Biomedical Applications. Acta Biomater. 2019, 97, 1–22. [Google Scholar] [CrossRef] [PubMed]
  385. Sousa, V.F.C.; Silva, F.J.G. Recent Advances on Coated Milling Tool Technology-a Comprehensive Review. Coatings 2020, 10, 235. [Google Scholar] [CrossRef]
  386. Simka, W.; Kaczmarek, M.; Baron-Wiecheć, A.; Nawrat, G.; Marciniak, J.; Żak, J. Electropolishing and Passivation of NiTi Shape Memory Alloy. Electrochim. Acta 2010, 55, 2437–2441. [Google Scholar] [CrossRef]
  387. Espinar, E.; Llamas, J.M.; Michiardi, A.; Ginebra, M.P.; Gil, F.J. Reduction of Ni Release and Improvement of the Friction Behaviour of NiTi Orthodontic Archwires by Oxidation Treatments. J. Mater. Sci. Mater. Med. 2011, 22, 1119–1125. [Google Scholar] [CrossRef] [PubMed]
  388. HORIUCHI, Y.; HORIUCHI, M.; HANAWA, T.; SOMA, K. Effect of Surface Modification on the Photocatalysis of Ti-Ni Alloy in Orthodontics. Dent. Mater. J. 2007, 26, 924–929. [Google Scholar] [CrossRef] [PubMed]
  389. Chun, M.J.; Shim, E.; Kho, E.H.; Park, K.J.; Jung, J.; Kim, J.M.; Kim, B.; Lee, K.H.; Cho, D.L.; Bai, D.H.; et al. Surface Modification of Orthodontic Wires with Photocatalytic Titanium Oxide for Its Antiadherent and Antibacterial Properties. Angle Orthod. 2007, 77, 483–488. [Google Scholar] [CrossRef] [PubMed]
  390. Chhattani, S.; Shetty, P.C.; Laxmikant, S.; Ramachandra, C. In Vitro Assessment of Photocatalytic Titanium Oxide Surface-Modified Stainless Steel and Nickel Titanium Orthodontic Wires for Its Antiadherent and Antibacterial Properties against Streptococcus Mutans. J. Indian Orthod. Soc. 2014, 48, 82–87. [Google Scholar] [CrossRef]
  391. Awata, M.; Okada, M.; Nambu, T.; Matsumoto, N. Improvement of Photocatalytic Activity of TiO2 Coating by the Modified Sol-Gel Method. Nano Biomed. 2015, 7, 51–62. [Google Scholar] [CrossRef]
  392. Özyildz, F.; Uzel, A.; Hazar, A.S.; Güden, M.; Ölmez, S.; Aras, I.; Karaboz, İ. Photocatalytic Antimicrobial Effect of TiO2 Anatase Thin-Film–Coated Orthodontic Arch Wires on 3 Oral Pathogens. Turk. J. Biol. 2014, 38, 289–295. [Google Scholar] [CrossRef]
  393. Salehi, P.; Babanouri, N.; Roein-Peikar, M.; Zare, F. Long-Term Antimicrobial Assessment of Orthodontic Brackets Coated with Nitrogen-Doped Titanium Dioxide against Streptococcus mutans. Prog. Orthod. 2018, 19, 35. [Google Scholar] [CrossRef]
  394. Cao, B.; Wang, Y.; Li, N.; Liu, B.; Zhang, Y. Preparation of an Orthodontic Bracket Coated with an Nitrogen-Doped TiO2-XNy Thin Film and Examination of Its Antimicrobial Performance. Dent. Mater. J. 2013, 32, 311–316. [Google Scholar] [CrossRef] [PubMed]
  395. Shah, A.G.; Shetty, P.C.; Ramachandra, C.S.; Bhat, N.S.; Laxmikanth, S.M. In Vitro Assessment of Photocatalytic Titanium Oxide Surface Modified Stainless Steel Orthodontic Brackets for Antiadherent and Antibacterial Properties against Lactobacillus Acidophilus. Angle Orthod. 2011, 81, 1028–1035. [Google Scholar] [CrossRef] [PubMed]
  396. Kaliaraj, G.S.; Ramadoss, A.; Sundaram, M.; Balasubramanian, S.; Muthirulandi, J. Studies of Calcium-Precipitating Oral Bacterial Adhesion on TiN, TiO2 Single Layer, and TiN/TiO2 Multilayer-Coated 316L SS. J. Mater. Sci. 2014, 49, 7172–7180. [Google Scholar] [CrossRef]
  397. Baby, R.D.; Subramaniam, S.; Arumugam, I.; Padmanabhan, S. Assessment of Antibacterial and Cytotoxic Effects of Orthodontic Stainless Steel Brackets Coated with Different Phases of Titanium Oxide: An in-Vitro Study. Am. J. Orthod. Dentofac. Orthop. 2017, 151, 678–684. [Google Scholar] [CrossRef] [PubMed]
  398. Fatani, E.J.; Almutairi, H.H.; Alharbi, A.O.; Alnakhli, Y.O.; Divakar, D.D.; Muzaheed; Alkheraif, A.A.; Khan, A.A. In Vitro Assessment of Stainless Steel Orthodontic Brackets Coated with Titanium Oxide Mixed Ag for Anti-Adherent and Antibacterial Properties against Streptococcus mutans and Porphyromonas gingivalis. Microb. Pathog. 2017, 112, 190–194. [Google Scholar] [CrossRef]
  399. Zhang, R.; Zhang, W.; Bai, X.; Song, X.; Wang, C.; Gao, X.; Tian, X.; Liu, F. Discussion on the Development of Nano Ag/TiO2 Coating Bracket and Its Antibacterial Property and Biocompatibility in Orthodontic Treatment. Pak. J. Pharm. Sci. 2015, 28, 807–810. [Google Scholar]
  400. Kielan-Grabowska, Z.; Bącela, J.; Zięty, A.; Seremak, W.; Gawlik-Maj, M.; Kawala, B.; Borak, B.; Detyna, J.; Sarul, M. Improvement of Properties of Stainless Steel Orthodontic Archwire Using TiO2:Ag Coating. Symmetry 2021, 13, 1734. [Google Scholar] [CrossRef]
  401. Bącela, J.J.; Kielan-Grabowska, Z.; Borak, B.; Sobieszczańska, B.; Walczuk, U.; Kawala, B.; Zięty, A.; Detyna, J.; Sarul, M. Antiadherent and Antibacterial Properties of TiO2-Coated and TiO2:Ag-Coated Stainless Steel Orthodontic Wires against S. mutans Bacteria. Acta Bioeng. Biomech. 2022, 24, 107–118. [Google Scholar] [CrossRef]
  402. Li, N.; Zhou, H.J.; Han, B.; Zhang, Y.J.; Cao, B.C. Friction Properties of Orthodontic Brackets Coated with TiO2-XNy. Chin. J. Tissue Eng. Res. 2014, 18, 7621–7626. [Google Scholar] [CrossRef]
  403. Mollabashi, V.; Farmany, A.; Alikhani, M.Y.; Sattari, M.; Soltanian, A.R.; Kahvand, P.; Banisafar, Z. Effects of TiO2-Coated Stainless Steel Orthodontic Wires on Streptococcus Mutans Bacteria: A Clinical Study. Int. J. Nanomed. 2020, 15, 8759–8766. [Google Scholar] [CrossRef]
  404. Jung, O.; Becker, J.-P.; Smeets, R.; Gosau, M.; Becker, G.; Kahl-Nieke, B.; Jung, A.-K.; Heiland, M.; Kopp, A.; Barbeck, M.; et al. Surface Characteristics of Esthetic Nickel–Titanium and Beta-Titanium Orthodontic Archwires Produced by Plasma Electrolytic Oxidation (PEO)—Primary Results. Materials 2019, 12, 1403. [Google Scholar] [CrossRef] [PubMed]
  405. Liu, F.; Xu, J.L.; Yu, D.Z.; Wang, F.P.; Zhao, L.C. Wear Resistance of Micro-Arc Oxidation Coatings on Biomedical NiTi Alloy. J. Alloys Compd. 2009, 487, 391–394. [Google Scholar] [CrossRef]
  406. Campeol, D.A.; Fontoura, C.P.; Rodrigues, M.M.; Aguzzoli, C. Assessment of Mechanical and Corrosion Properties of Plasma Oxidized Medical Grade NiTi Wire. Vacuum 2020, 171, 109013. [Google Scholar] [CrossRef]
  407. Supriadi, S.; Suharno, B.; Nugraha, N.K.; Yasinta, A.O.; Annur, D. Adhesiveness of TiO2 PVD Coating on Electropolished Stainless Steel 17–4 PH Orthodontic Bracket. Mater. Res. Express 2019, 6, 094003. [Google Scholar] [CrossRef]
  408. Fu, T.; Wen, C.S.; Lu, J.; Zhou, Y.M.; Ma, S.G.; Dong, B.H.; Liu, B.G. Sol-Gel Derived TiO2 Coating on Plasma Nitrided 316L Stainless Steel. Vacuum 2012, 86, 1402–1407. [Google Scholar] [CrossRef]
  409. Liu, J.; Lou, Y.; Zhang, C.; Yin, S.; Li, H.; Sun, D.; Sun, X. Improved Corrosion Resistance and Antibacterial Properties of Composite Arch-Wires by N-Doped TiO2 Coating. RSC Adv. 2017, 7, 43938–43949. [Google Scholar] [CrossRef]
  410. Solanki, L.A.; Dinesh, S.P.S.; Jain, R.K.; Balasubramaniam, A. Effects of Titanium Oxide Coating on the Antimicrobial Properties, Surface Characteristics, and Cytotoxicity of Orthodontic Brackets—A Systematic Review and Meta Analysis of in-Vitro Studies. J. Oral Biol. Craniofacial Res. 2023, 13, 553–562. [Google Scholar] [CrossRef]
  411. Asiry, M.A.; AlShahrani, I.; Almoammar, S.; Durgesh, B.H.; Al Kheraif, A.A.; Hashem, M.I. Influence of Epoxy, Polytetrafluoroethylene (PTFE) and Rhodium Surface Coatings on Surface Roughness, Nano-Mechanical Properties and Biofilm Adhesion of Nickel Titanium (Ni-Ti) Archwires. Mater. Res. Express 2018, 5, 026511. [Google Scholar] [CrossRef]
  412. Madasamy, R.; Prabhakar, R.; Ramanadhan, T.; Ramachandran, S.; Natrajan, R.; Vasudevan, S. Assessment of Roughness of Fiber-Reinforced Polymer Composite Wires and Other Coated Esthetic Archwires. World J. Dent. 2021, 12, 156–159. [Google Scholar] [CrossRef]
  413. Tawakal, M.S.; Abdelghany Metwally, A.M.; El-Wassefy, N.A.; Tawfik, M.A.; Shamaa, M.S. Static Friction, Surface Roughness, and Antibacterial Activity of Orthodontic Brackets Coated with Silver and Silver Chitosan Nanoparticles. J. World Fed. Orthod. 2023. [Google Scholar] [CrossRef]
  414. Setiyorini, Y.; Pintowantoro, S. Biocompatibility Improvement of NiTi Orthodontic Wire from Various Coatings. Adv. Mater. Res. 2013, 789, 225–231. [Google Scholar] [CrossRef]
  415. Pintowantoro, S.; Setiyorini, Y. Reduction of Nickel Ion Release on a TiO2 Coated onto an Orthodontic Wire. Adv. Mater. Res. 2013, 789, 204–209. [Google Scholar] [CrossRef]
  416. Supriadi, S.; Ovilia, A.; Ilmaniar, N.; Suharno, B. Formation of Tio2 Thin Film on Antibacterial Metal Injection Molding Stainless Steel Orthodontic Bracket 17-4 Ph Using Physical Vapor Deposition Method. Key Eng. Mater. 2020, 846, 169–174. [Google Scholar] [CrossRef]
  417. Kurtoğlu, S.F.; Yağcı, M.B.; Uzun, A.; Ünal, U.; Canadinc, D. Enhancing Biocompatibility of NiTi Shape Memory Alloys by Simple NH3 Treatments. Appl. Surf. Sci. 2020, 525, 146547. [Google Scholar] [CrossRef]
  418. Math, M.; Shah, A.G.; Gangurde, P.; Karandikar, A.G.; Gheware, A.; Jadhav, B.S. In-Vitro Comparative Assessment of Antibacterial and Anti-Adherent Effect of Two Types of Surface Modificants on Stainless Steel Orthodontic Brackets Against Streptococcus Mutans. J. Indian Orthod. Soc. 2021, 56, 282–289. [Google Scholar] [CrossRef]
  419. Kim, H.; Johnson, J.W. Corrosion of Stainless-Steel, Nickel-Titanium, Coated Nickel-Titanium, and Titanium Orthodontic Wires. Angle Orthod. 1999, 69, 39–44. [Google Scholar] [CrossRef]
  420. Kao, C.T.; Ding, S.J.; Chen, Y.C.; Huang, T.H. The Anticorrosion Ability of Titanium Nitride (TiN) Plating on an Orthodontic Metal Bracket and Its Biocompatibility. J. Biomed. Mater. Res. 2002, 63, 786–792. [Google Scholar] [CrossRef]
  421. Gil, F.J.; Solano, E.; Mendoza, A.; Pena, J. Inhibition of Ni Release from NiTi and NiTiCu Orthodontic Archwires by Nitrogen Diffusion Treatment. J. Appl. Biomater. Biomech. 2004, 2, 151–155. [Google Scholar] [CrossRef]
  422. Iijima, M.; Yuasa, T.; Endo, K.; Muguruma, T.; Ohno, H.; Mizoguchi, I. Corrosion Behavior of Ion Implanted Nickel-Titanium Orthodontic Wire in Fluoride Mouth Rinse Solutions. Dent. Mater. J. 2010, 29, 53–58. [Google Scholar] [CrossRef]
  423. Huang, T.-H.; Guo, J.-U.; Kao, C.T. A Comparison of the Friction Associated with Diamond-like Carbon (DLC) or Titanium Nitride (TiN) Plating Metal Brackets. Surf. Coat. Technol. 2010, 205, 1917–1921. [Google Scholar] [CrossRef]
  424. Kao, C.-T.; Guo, J.-U.; Huang, T.-H. Comparison of Friction Force between Corroded and Noncorroded Titanium Nitride Plating of Metal Brackets. Am. J. Orthod. Dentofac. Orthop. 2011, 139, 594–600. [Google Scholar] [CrossRef]
  425. Rongo, R.; Ametrano, G.; Gloria, A.; Spagnuolo, G.; Galeotti, A.; Paduano, S.; Valletta, R.; D’Antò, V. Effects of Intraoral Aging on Surface Properties of Coated Nickel-Titanium Archwires. Angle Orthod. 2014, 84, 665–672. [Google Scholar] [CrossRef] [PubMed]
  426. Zuo, J.; Xie, Y.; Zhang, J.; Wei, Q.; Zhou, B.; Luo, J.; Wang, Y.; Yu, Z.M.; Tang, Z.G. TiN Coated Stainless Steel Bracket: Tribological, Corrosion Resistance, Biocompatibility and Mechanical Performance. Surf. Coat. Technol. 2015, 277, 227–233. [Google Scholar] [CrossRef]
  427. Rongo, R.; Valletta, R.; Bucci, R.; Rivieccio, V.; Galeotti, A.; Michelotti, A.; D’Antò, V. In Vitro Biocompatibility of Nickel-Titanium Esthetic Orthodontic Archwires. Angle Orthod. 2016, 86, 789–795. [Google Scholar] [CrossRef] [PubMed]
  428. Sugisawa, H.; Kitaura, H.; Ueda, K.; Kimura, K.; Ishida, M.; Ochi, Y.; Kishikawa, A.; Ogawa, S.; Takano-Yamamoto, T. Corrosion Resistance and Mechanical Properties of Titanium Nitride Plating on Orthodontic Wires. Dent. Mater. J. 2018, 37, 286–292. [Google Scholar] [CrossRef]
  429. Arici, N.; Akdeniz, B.S.; Oz, A.A.; Gencer, Y.; Tarakci, M.; Arici, S. Effectiveness of Medical Coating Materials in Decreasing Friction between Orthodontic Brackets and Archwires. Korean J. Orthod. 2021, 51, 270–281. [Google Scholar] [CrossRef] [PubMed]
  430. Teixeira, L.P.; Gontijo, L.C.; Franco Júnior, A.R.; Pereira, M.F.; Schuenck, R.P.; Malacarne-Zanon, J. Evaluation of Antimicrobial Potential and Surface Morphology in Thin Films of Titanium Nitride and Calcium Phosphate on Orthodontic Brackets. Am. J. Orthod. Dentofac. Orthop. 2021, 160, 209–214. [Google Scholar] [CrossRef]
  431. Liu, J.K.; Liu, I.H.; Liu, C.; Chang, C.J.; Kung, K.C.; Liu, Y.T.; Lee, T.M.; Jou, J.L. Effect of Titanium Nitride/Titanium Coatings on the Stress Corrosion of Nickel-Titanium Orthodontic Archwires in Artificial Saliva. Appl. Surf. Sci. 2014, 317, 974–981. [Google Scholar] [CrossRef]
  432. Krishnan, V.; Krishnan, A.; Remya, R.; Ravikumar, K.K.; Nair, S.A.; Shibli, S.M.A.; Varma, H.K.; Sukumaran, K.; Kumar, K.J. Development and Evaluation of Two PVD-Coated β-Titanium Orthodontic Archwires for Fluoride-Induced Corrosion Protection. Acta Biomater. 2011, 7, 1913–1927. [Google Scholar] [CrossRef]
  433. Krishnan, V.; Ravikumar, K.K.; Sukumaran, K.; Jyothindra Kumar, K. In Vitro Evaluation of Physical Vapor Deposition Coated Beta Titanium Orthodontic Archwires. Angle Orthod. 2012, 82, 22–29. [Google Scholar] [CrossRef]
  434. Golshah, A.; Feyli, S. Effect of Zirconium Oxide Nano-Coating on Frictional Resistance of Orthodontic Wires. J. Orthod. Sci. 2022, 11, 35. [Google Scholar] [CrossRef]
  435. Wu, H.; Yang, J.; Yan, Y.; Zheng, B.; Algahefi, A.L.; Ma, S.; Liu, Y. Study of Al–SiO2 Aesthetic Composite Coating on Orthodontic Metal Archwire. Coatings 2022, 12, 746. [Google Scholar] [CrossRef]
  436. Krishnan, M.; Seema, S.; Sukumaran, K.; Pawar, V. Phase Transitions in Coated Nickel Titanium Arch Wires: A Differential Scanning Calorimetric and X-Ray Diffraction Analysis. Bull. Mater. Sci. 2012, 35, 905–911. [Google Scholar] [CrossRef]
  437. Ramazanzadeh, B.; Jahanbin, A.; Yaghoubi, M.; Shahtahmassbi, N.; Ghazvini, K.; Shakeri, M.; Shafaee, H. Comparison of Antibacterial Effects of ZnO and CuO Nanoparticles Coated Brackets against Streptococcus Mutans. J. Dent. 2015, 16, 200–205. [Google Scholar]
  438. Kachoei, M.; Divband, B.; Eskandarinejad, F.; Khatamian, M. Deposition of ZnO Nano Particles on Stainless Steel Orthodontic Wires by Chemical Solution Method for Friction Reduction Propose. Res. J. Pharm. Biol. Chem. Sci. 2015, 6, 104–112. [Google Scholar]
  439. Behroozian, A.; Kachoei, M.; Khatamian, M.; Divband, B. The Effect of ZnO Nanoparticle Coating on the Frictionalresistance between Orthodontic Wires and Ceramic Brackets. J. Dent. Res. Dent. Clin. Dent. Prospect. 2016, 10, 106–111. [Google Scholar] [CrossRef] [PubMed]
  440. Kachoei, M.; Nourian, A.; Divband, B.; Kachoei, Z.; Shirazi, S. Zinc-Oxide Nanocoating for Improvement of the Antibacterial and Frictional Behavior of Nickel-Titanium Alloy. Nanomedicine 2016, 11, 2511–2527. [Google Scholar] [CrossRef]
  441. Hammad, S.M.; El-Wassefy, N.A.; Shamaa, M.S.; Fathy, A. Evaluation of Zinc-Oxide Nanocoating on the Characteristics and Antibacterial Behavior of Nickel-Titanium Alloy. Dent. Press J. Orthod. 2020, 25, 51–58. [Google Scholar] [CrossRef]
  442. Gholami, M.; Esmaeilzadeh, M.; Kachoei, Z.; Kachoei, M.; Divband, B. Influence of Physical Dimension and Morphological-Dependent Antibacterial Characteristics of ZnO Nanoparticles Coated on Orthodontic NiTi Wires. Biomed Res. Int. 2021, 2021, 6397698. [Google Scholar] [CrossRef]
  443. Elhelbawy, N.; Ellaithy, M. Comparative Evaluation of Stainless-Steel Wires and Brackets Coated with Nanoparticles of Chitosan or Zinc Oxide upon Friction: An in Vitro Study. Int. Orthod. 2021, 19, 274–280. [Google Scholar] [CrossRef]
  444. Palanivel, J.; Srinivasan, D.; Chakravathy, N.C.S. Comparison of the Frictional Resistance and Optical Properties of Aluminum Oxide and Zinc Oxide Coated Nickel Titanium Archwires—An in Vitro Study. APOS Trends Orthod. 2022, 12, 168. [Google Scholar] [CrossRef]
  445. Tanbakuchi, B.; Kharrazi, S.; Nikfarjam, M.; Akhoundi, M.S.A.; Shahroudi, A.S. Comparative Assessment of the Orthodontic Wire’s Friction Coated with Zinc Oxide Nanoparticles by Two Methods of Chemical Precipitation and Hydrothermal Process. Folia Med. 2022, 64, 945–952. [Google Scholar] [CrossRef] [PubMed]
  446. Ameli, N.; Ghorbani, R.; Asadi, S.; Zarrinzade, Z. Investigation of the Effects of Orthodontic Brackets Coated by Silver Hydroxyapatite, Copper Oxide, and Titanium Oxide Nanoparticles on Wire-Bracket Friction. APOS Trends Orthod. 2022, 12, 27. [Google Scholar] [CrossRef]
  447. Ameli, N.; Asadi, S.; Ghorbani, R.; Mohebi, S.; Hans, M. Comparative Antibacterial Efficacy of Orthodontic Brackets Coated with Titanium Dioxide, Copper Oxide, and Hydroxyapatite-Silver Nanoparticles against Streptococcus mutans. Middle East J. Rehabil. Health Stud. 2022, 9, e119536. [Google Scholar] [CrossRef]
  448. Venkatesan, K.; Kailasam, V.; Padmanabhan, S. Evaluation of Titanium Dioxide Coating on Surface Roughness of Nickel-Titanium Archwires and Its Influence on Streptococcus Mutans Adhesion and Enamel Mineralization: A Prospective Clinical Study. Am. J. Orthod. Dentofac. Orthop. 2020, 158, 199–208. [Google Scholar] [CrossRef]
  449. Silveira, R.E.; Elias, C.N.; Amaral, F.L.B. Assessment of Frictional Resistance and Surface Roughness in Orthodontic Wires Coated with Two Different Nanoparticles. Microsc. Res. Tech. 2022, 85, 1884–1890. [Google Scholar] [CrossRef]
  450. Chaturvedi, T.; Indumathi, P.; Sharma, V.; Agrawal, A.; Singh, D.; Upadhyay, C. Evaluation of Surface-Modified Orthodontic Wires by Different Concentration and Dipping Duration of Titanium Oxide (TiO2) Nanoparticles. J. Orthod. Sci. 2023, 12, 3. [Google Scholar] [CrossRef]
  451. Redlich, M.; Gorodnev, A.; Feldman, Y.; Kaplan-Ashiri, I.; Tenne, R.; Fleischer, N.; Genut, M.; Feuerstein, N. Friction Reduction and Wear Resistance of Electro-Co-Deposited Inorganic Fullerene-like WS 2 Coating for Improved Stainless Steel Orthodontic Wires. J. Mater. Res. 2008, 23, 2909–2915. [Google Scholar] [CrossRef]
  452. Amorodnitzky-Naveh, G.R.S.; Redlich, M.; Katz, A.; Adini, A.R.; Gorodnev, A.; Rapoport, L.; Moshkovich, A.; Cohen, S.R.; Rosentsveig, R.; Moshonov, J.; et al. Towards Medical Applications of Self-Lubricating Coatings with Fullerene-like (IF) WS2 Nanoparticles. Int. J. Nano Biomater. 2010, 3, 140. [Google Scholar] [CrossRef]
  453. Katz, A.; Redlich, M.; Rapoport, L.; Wagner, H.D.; Tenne, R. Self-Lubricating Coatings Containing Fullerene-like WS2 Nanoparticles for Orthodontic Wires and Other Possible Medical Applications. Tribol. Lett. 2006, 21, 135–139. [Google Scholar] [CrossRef]
  454. Redlich, M.; KATZ, A.; RAPOPORT, L.; WAGNER, H.; FELDMAN, Y.; TENNE, R. Improved Orthodontic Stainless Steel Wires Coated with Inorganic Fullerene-like Nanoparticles of WS2 Impregnated in Electroless Nickel–Phosphorous Film. Dent. Mater. 2008, 24, 1640–1646. [Google Scholar] [CrossRef] [PubMed]
  455. Gracco, A.; Dandrea, M.; Deflorian, F.; Zanella, C.; De Stefani, A.; Bruno, G.; Stellini, E. Application of a Molybdenum and Tungsten Disulfide Coating to Improve Tribological Properties of Orthodontic Archwires. Nanomaterials 2019, 9, 753. [Google Scholar] [CrossRef] [PubMed]
  456. Samorodnitzky-Naveh, G.R.; Redlich, M.; Rapoport, L.; Feldman, Y.; Tenne, R. Inorganic Fullerene-like Tungsten Disulfide Nanocoating for Friction Reduction of Nickel–Titanium Alloys. Nanomedicine 2009, 4, 943–950. [Google Scholar] [CrossRef] [PubMed]
  457. Fu, Y.; Wu, X.; Wang, Y.; Li, B.; Yang, S. Study of Corrosion Resistance Property and Microstructure of TiNi Shape Memory Alloy Modified by Pulsed High-Energy Density Plasma. Appl. Surf. Sci. 2000, 157, 167–177. [Google Scholar] [CrossRef]
  458. Shevchenko, N.; Pham, M.-T.; Maitz, M.F. Studies of Surface Modified NiTi Alloy. Appl. Surf. Sci. 2004, 235, 126–131. [Google Scholar] [CrossRef]
  459. Zhang, M.; Liu, X.; Shang, H.; Lin, J. Comparison of TiN and CNx Coatings on Orthodontic Stainless Steel: Tribological and Biological Evaluation. Surf. Coat. Technol. 2019, 362, 381–387. [Google Scholar] [CrossRef]
  460. Jin, S.; Zhang, Y.; Wang, Q.; Zhang, D.; Zhang, S. Influence of TiN Coating on the Biocompatibility of Medical NiTi Alloy. Colloids Surf. B Biointerfaces 2013, 101, 343–349. [Google Scholar] [CrossRef] [PubMed]
  461. Zhang, J.; Lou, J.; He, H.; Xie, Y. Comparative Investigation on the Tribological Performances of TiN, TiCN, and Ti-DLC Film-Coated Stainless Steel. JOM 2019, 71, 4872–4879. [Google Scholar] [CrossRef]
  462. Wang, X.; Bai, S.; Li, F.; Li, D.; Zhang, J.; Tian, M.; Zhang, Q.; Tong, Y.; Zhang, Z.; Wang, G.; et al. Effect of Plasma Nitriding and Titanium Nitride Coating on the Corrosion Resistance of Titanium. J. Prosthet. Dent. 2016, 116, 450–456. [Google Scholar] [CrossRef]
  463. Shukla, K.; Rane, R.; Alphonsa, J.; Maity, P.; Mukherjee, S. Structural, Mechanical and Corrosion Resistance Properties of Ti/TiN Bilayers Deposited by Magnetron Sputtering on AISI 316L. Surf. Coat. Technol. 2017, 324, 167–174. [Google Scholar] [CrossRef]
  464. Bellini, H.; Moyano, J.; Gil, J.; Puigdollers, A. Comparison of the Superelasticity of Different Nickel–Titanium Orthodontic Archwires and the Loss of Their Properties by Heat Treatment. J. Mater. Sci. Mater. Med. 2016, 27, 158. [Google Scholar] [CrossRef] [PubMed]
  465. Rahmati, M.; Mozafari, M. Biocompatibility of Alumina-based Biomaterials–A Review. J. Cell. Physiol. 2019, 234, 3321–3335. [Google Scholar] [CrossRef] [PubMed]
  466. Gautam, C.; Joyner, J.; Gautam, A.; Rao, J.; Vajtai, R. Zirconia Based Dental Ceramics: Structure, Mechanical Properties, Biocompatibility and Applications. Dalt. Trans. 2016, 45, 19194–19215. [Google Scholar] [CrossRef]
  467. Kaliaraj, G.S.; Vishwakarma, V.; Kirubaharan, K.; Dharini, T.; Ramachandran, D.; Muthaiah, B. Corrosion and Biocompatibility Behaviour of Zirconia Coating by EBPVD for Biomedical Applications. Surf. Coat. Technol. 2018, 334, 336–343. [Google Scholar] [CrossRef]
  468. Louro, C.S.; Fróis, A.; Cerqueira, M. Are W Based Coatings a Surface Alternative to the Adverse Ni Release in Orthodontic Applications? (Poster). In Proceedings of the Iberian Vacuum Conference—RIVA-X, Bilbao, Spain, 4–6 October 2017. [Google Scholar]
  469. Sánchez-López, E.; Gomes, D.; Esteruelas, G.; Bonilla, L.; Lopez-Machado, A.L.; Galindo, R.; Cano, A.; Espina, M.; Ettcheto, M.; Camins, A.; et al. Metal-Based Nanoparticles as Antimicrobial Agents: An Overview. Nanomaterials 2020, 10, 292. [Google Scholar] [CrossRef] [PubMed]
  470. Moradpoor, H.; Safaei, M.; Mozaffari, H.R.; Sharifi, R.; Imani, M.M.; Golshah, A.; Bashardoust, N. An Overview of Recent Progress in Dental Applications of Zinc Oxide Nanoparticles. RSC Adv. 2021, 11, 21189–21206. [Google Scholar] [CrossRef]
  471. Puts, G.J.; Crouse, P.; Ameduri, B.M. Polytetrafluoroethylene: Synthesis and Characterization of the Original Extreme Polymer. Chem. Rev. 2019, 119, 1763–1805. [Google Scholar] [CrossRef]
  472. Roina, Y.; Auber, F.; Hocquet, D.; Herlem, G. EPTFE Functionalization for Medical Applications. Mater. Today Chem. 2021, 20, 100412. [Google Scholar] [CrossRef]
  473. Kameda, T.; Sato, H.; Oka, S.; Miyazaki, A.; Ohkuma, K.; Terada, K. Low Temperature Polytetrafluoroethylene (PTFE) Coating Improves the Appearance of Orthodontic Wires without Changing Their Mechanical Properties. Dent. Mater. J. 2020, 39, 721–734. [Google Scholar] [CrossRef]
  474. Abdulkader, Y.C.; Kamaruddin, A.F.; Mydin, R.B.S.M.N. Effects of Salivary PH on Coating Durability of Two Different Aesthetic Archwire Coatings under a Simulated Intraoral Environment. Saudi Dent. J. 2020, 32, 306–313. [Google Scholar] [CrossRef]
  475. Mareci, D.; Earar, K.; Zetu, I.; Bolat, G.; Crimu, C.; Istrate, B.; Munteanu, C.; Matei, M. Comparative Electrochemical Behaviour of Uncoated and Coated NiTi for Dental Orthodontic Wires. Mater. Plast. 2015, 2, 150–153. [Google Scholar]
  476. Earar, K.; Matei, M.N.; Mareci, D.; Trinca, L.C.; Ariton, M.A.; Arotaritei, D.; Cerghizan, D.; Bica, C.; Ciupilan, C. The Electrochemical Behaviour of NiTi Orthodontic Wires with Polytetrafluoroethylene Coating in Red Ruffled Pimiento Peppers Paste. Rev. Chim. 2016, 67, 1850–1853. [Google Scholar]
  477. Matei, M.N.; Earar, K.; Trinca, L.C.; Mareci, D.; Fotea, L.; Peptu, C.A.; Bica, C. Degradation Characteristics of Poly-Tetrafluoroethylene Coatings on Stainless Steel Orthodontic Wires Immersed in Tuna Fish Derived Products. Rev. Chim. 2016, 67, 800–807. [Google Scholar]
  478. Farronato, G.; Maijer, R.; Carìa, M.P.; Esposito, L.; Alberzoni, D.; Cacciatore, G. The Effect of Teflon Coating on the Resistance to Sliding of Orthodontic Archwires. Eur. J. Orthod. 2012, 34, 410–417. [Google Scholar] [CrossRef] [PubMed]
  479. Husmann, P.; Bourauel, C.; Wessinger, M.; Jäger, A. The Frictional Behavior of Coated Guiding Archwires. J. Orofac. Orthop. Fortschritte Kieferorthopädie 2002, 63, 199–211. [Google Scholar] [CrossRef] [PubMed]
  480. Demling, A.; Elter, C.; Heidenblut, T.; Bach, F.-W.; Hahn, A.; Schwestka-Polly, R.; Stiesch, M.; Heuer, W. Reduction of Biofilm on Orthodontic Brackets with the Use of a Polytetrafluoroethylene Coating. Eur. J. Orthod. 2010, 32, 414–418. [Google Scholar] [CrossRef] [PubMed]
  481. Ryu, S.H.; Lim, B.S.; Kwak, E.J.; Lee, G.J.; Choi, S.; Park, K.H. Surface Ultrastructure and Mechanical Properties of Three Different White-Coated NiTi Archwires. Scanning 2015, 37, 414–421. [Google Scholar] [CrossRef]
  482. Elsaka, S.; Hassan, A.; Elnaghy, A. Effect of Gastric Acids on Surface Topography and Bending Properties of Esthetic Coated Nickel-Titanium Orthodontic Archwires. Clin. Oral Investig. 2021, 25, 1319–1326. [Google Scholar] [CrossRef]
  483. Dokku, A.; Peddu, R.; Prakash, A.S.; Padhmanabhan, J.; Kalyani, M.; Devikanth, L. Surface and Mechanical Properties of Different Coated Orthodontic Archwires. J. Indian Orthod. Soc. 2018, 52, 238–242. [Google Scholar] [CrossRef]
  484. Choi, S.; Park, D.-J.; Kim, K.-A.; Park, K.-H.; Park, H.-K.; Park, Y.-G. In Vitro Sliding-Driven Morphological Changes in Representative Esthetic NiTi Archwire Surfaces. Microsc. Res. Tech. 2015, 78, 926–934. [Google Scholar] [CrossRef]
  485. Rego, M.V.N.N.D.; Lau, G.W.T.; Araújo, Y.C.; Silva, R.M. e Color Stability of Esthetic Coatings Applied to Nickel-Titanium Archwires. Rev. Odontol. UNESP 2017, 46, 307–311. [Google Scholar] [CrossRef]
  486. Jejurikar, H.; Contractor, T.; Nene, S.; Kalia, A.; Patil, W.; Khan, N. A Comparison of Surface Characteristics, Coating Stability and Friction Coefficients of Esthetic Archwires: A Comparative Study. J. Indian Orthod. Soc. 2021, 55, 56–63. [Google Scholar] [CrossRef]
  487. Shahabi, M.; Salari, S.; Poosti, M.; Abtahi, M. Static and Kinetic Frictional Forces of Silica-Insert Ceramic Brackets with Coated Archwires in Artificial Saliva. Dent. Res. J. 2017, 14, 412. [Google Scholar] [CrossRef]
  488. Argalji, N.; Da Silva, E.M.; Cury-Saramago, A.; Mattos, C.T. Characterization and Coating Stability Evaluation of Nickel-Titanium Orthodontic Esthetic Wires: An in Vivo Study. Braz. Oral Res. 2017, 31, e68. [Google Scholar] [CrossRef] [PubMed]
  489. Abdulkhabeer, H.; Alyassiri, A.; Farhan Kamaruddin, A.; Ismail, K.; Shafiai, N.A.A.; Rahman, N.A.; Amir, W.M.; Ahmad, W. Preliminary Result of Randomised Controlled Trial of Three Different Coated Archwires Part 1: Tooth Alignment and Coating Loss. Malays. J. Med. Health Sci. 2020, 16, 2636–9346. [Google Scholar]
  490. Zegan, G.; Sodor, A.; Munteanu, C. Surface Characteristics of Retrieved Coated and Nickel-Titanium Orthodontic Archwires. Rom. J. Morphol. Embryol. 2012, 53, 935–939. [Google Scholar] [PubMed]
  491. da Silva, D.L.; Santos, E.; de Souza Camargo, S.; de Oliveira Ruellas, A.C. Infrared Spectroscopy, Nano-Mechanical Properties, and Scratch Resistance of Esthetic Orthodontic Coated Archwires. Angle Orthod. 2015, 85, 777–783. [Google Scholar] [CrossRef]
  492. Lin, C.; Kim, J.S.; Lin, E.T.; Lin, E.T. Prolonged Water Immersion Alters Resistance to Sliding of Aesthetic Orthodontic Coated Wires. Orthod. Craniofac. Res. 2021, 24, 111–120. [Google Scholar] [CrossRef]
  493. Rudawska, A.; Sarna-Boś, K.; Rudawska, A.; Olewnik-Kruszkowska, E.; Frigione, M. Biological Effects and Toxicity of Compounds Based on Cured Epoxy Resins. Polymers 2022, 14, 4915. [Google Scholar] [CrossRef]
  494. Raji, S.H.; Shojaei, H.; Ghorani, P.S.; Rafiei, E. Bacterial Colonization on Coated and Uncoated Orthodontic Wires: A Prospective Clinical Trial. Dent. Res. J. 2014, 11, 680–683. [Google Scholar]
  495. Shamohammadi, M.; Hormozi, E.; Moradinezhad, M.; Moradi, M.; Skini, M.; Rakhshan, V. Surface Topography of Plain Nickel-Titanium (NiTi), as-Received Aesthetic (Coated) NiTi, and Aesthetic NiTi Archwires Sterilized by Autoclaving or Glutaraldehyde Immersion: A Profilometry/SEM/AFM Study. Int. Orthod. 2019, 17, 60–72. [Google Scholar] [CrossRef] [PubMed]
  496. Hosseini, N.; Alavi, S. Load-Deflection and Surface Properties of Coated and Conventional Superelastic Orthodontic Archwires in Conventional and Metal-Insert Ceramic Brackets. Dent. Res. J. 2012, 9, 133. [Google Scholar] [CrossRef]
  497. Pop, S.I.; Dudescu, M.; Bratu, D.C.; Merie, V.V.; Pacurar, M. Effect of Esthetic Coating on the Load Deflection and Surface Characteristics of the NiTi Orthodontic Archwires. Rev. Chim. 2015, 66, 364–367. [Google Scholar]
  498. Elayyan, F.; Silikas, N.; Bearn, D. Mechanical Properties of Coated Superelastic Archwires in Conventional and Self-Ligating Orthodontic Brackets. Am. J. Orthod. Dentofac. Orthop. 2010, 137, 213–217. [Google Scholar] [CrossRef] [PubMed]
  499. Dragomirescu, A.O.; Teodorescu, E.; Tarmure, V.; Baluta, A.; Pacurar, M.; Nenovici, D.; Chibelean, M.; Ionescu, E. Variation of Static Frictional Forces in the Fixed Orthodontic System. Rev. Chim. 2019, 70, 3954–3956. [Google Scholar] [CrossRef]
  500. Aboalnaga, A.A.; Shahawi, A.M. El Comparison of Surface Roughness and Hardness of Three Different Brands of Esthetic Coated NiTi Archwires: Invitro Study. BMC Oral Health 2023, 23, 816. [Google Scholar] [CrossRef]
  501. Elayyan, F.; Silikas, N.; Bearn, D. Ex Vivo Surface and Mechanical Properties of Coated Orthodontic Archwires. Eur. J. Orthod. 2008, 30, 661–667. [Google Scholar] [CrossRef]
  502. Ping, S.; Xue, F.; Jie He, S.; Wei, C.; Tao, W.; Wei, M. Color Schemes and Biocompatibility of Epoxy Resin/Polytetrafluorethylene Coat on the Surface of Tini Arth Wires. Int. J. Mod. Phys. B 2009, 23, 1578–1583. [Google Scholar]
  503. Sheiko, N.; Kékicheff, P.; Marie, P.; Schmutz, M.; Jacomine, L.; Perrin-Schmitt, F. PEEK (Polyether-Ether-Ketone)-Coated Nitinol Wire: Film Stability for Biocompatibility Applications. Appl. Surf. Sci. 2016, 389, 651–665. [Google Scholar] [CrossRef]
  504. Shirakawa, N.; Iwata, T.; Miyake, S.; Otuka, T.; Koizumi, S.; Kawata, T. Mechanical Properties of Orthodontic Wires Covered with a Polyether Ether Ketone Tube. Angle Orthod. 2018, 88, 442–449. [Google Scholar] [CrossRef]
  505. do Rego, M.V.N.N.; de Araújo, G.M.M.; Martinez, E.F.; de Sousa Lima, K.R.; Fortes, P.T.F.; Leal, L.M.P. Influence of Aesthetic Coating on the Load-Deflection Ratio of Nickel–Titanium Archwires. Braz. J. Oral Sci. 2017, 15, 293. [Google Scholar] [CrossRef]
  506. Zhou, A.; Makowka, S.; Warunek, S.; Chen, M.-Y.; Al-Jewair, T. Effects of Various Coating Methods on the Mechanical, Physical, and Aesthetic Properties of GUMMETAL® Archwires: In Vitro Study. Int. Orthod. 2023, 21, 100753. [Google Scholar] [CrossRef] [PubMed]
  507. Bravo, L.A.; de Cabañes, A.G.; Manero, J.M.; Rúperez, E.; Gil, F.J. NiTi Superelastic Orthodontic Archwires with Polyamide Coating. J. Mater. Sci. Mater. Med. 2014, 25, 555–560. [Google Scholar] [CrossRef] [PubMed]
  508. Tupinambá, R.A.; Claro, C.A.D.A.; Pereira, C.A.; Nobrega, C.J.P.; Claro, A.P.R.A. Bacterial Adhesion on Conventional and Self-Ligating Metallic Brackets after Surface Treatment with Plasma-Polymerized Hexamethyldisiloxane. Dent. Press J. Orthod. 2017, 22, 77–85. [Google Scholar] [CrossRef] [PubMed]
  509. Oliveira, A.S.; Kaizer, M.R.; Azevedo, M.S.; Ogliari, F.A.; Cenci, M.S.; Moraes, R.R. (Super)Hydrophobic Coating of Orthodontic Dental Devices and Reduction of Early Oral Biofilm Retention. Biomed. Mater. 2015, 10, 065004. [Google Scholar] [CrossRef]
  510. Kunimatsu, R.; Tsuka, Y.; Nakajima, K.; Sumi, K.; Yoshimi, Y.; Kado, I.; Inada, A.; Kiritoshi, Y.; Tanimoto, K. The Influence of 2-Methacryloyloxyethyl Phosphorylcholine Polymer Materials on Orthodontic Friction and Attachment of Oral Bacteria. Materials 2022, 15, 5770. [Google Scholar] [CrossRef] [PubMed]
  511. Wang, Y.; Xu, J.; Yu, C.; Zhou, X.; Chang, L.; Liu, J.; Peng, Q. Prevention of Bacterial Biofilm Formation on Orthodontic Brackets by Non-Crosslinked Chitosan Coating. Int. J. Biol. Macromol. 2023, 251, 126283. [Google Scholar] [CrossRef]
  512. Kumarasinghe, L.S.; Ninan, N.; Dabare, P.R.L.; Cavallaro, A.; Doğramacı, E.J.; Rossi-Fedele, G.; Dreyer, C.; Vasilev, K.; Zilm, P. Bioactive Plasma Coatings on Orthodontic Brackets: In Vitro Metal Ion Release and Cytotoxicity. Coatings 2021, 11, 857. [Google Scholar] [CrossRef]
  513. Liu, X.; Peng, L.; Meng, J.; Zhu, Z.; Han, B.; Wang, S. Protein-Mediated Anti-Adhesion Surface against Oral Bacteria. Nanoscale 2018, 10, 2711–2714. [Google Scholar] [CrossRef]
  514. Li, M.; Tang, C.; Yu, X.; Shi, X.; Yu, H.; Yin, H.; You, M.; Chen, Q.; Ding, X. Tough Adhesion and Antifouling Poly(Vinyl Alcohol) Hydrogel Coating on the Arch Wire for Antibacterial Adhesion. Sci. China Technol. Sci. 2023, 66, 2786–2796. [Google Scholar] [CrossRef]
  515. Wang, Y.; Ding, C.; Ge, Z.; Li, Z.; Chen, L.; Guo, X.; Dong, G.; Zhou, P. A Novel Antibacterial and Fluorescent Coating Composed of Polydopamine and Carbon Dots on the Surface of Orthodontic Brackets. J. Mater. Sci. Mater. Med. 2023, 34, 10. [Google Scholar] [CrossRef]
  516. Ahmed, R.A.; Farghali, R.A.; Alshahrani, W.A. Influence of Fluoride and/or Bovine Albumin on Electrochemical Properties of Bare and Ionic Liquid- Coated Ni47Ti49Co4 Orthodontic Archwires in Artificial Saliva Solution. Int. J. Electrochem. Sci. 2021, 16, 211230. [Google Scholar] [CrossRef]
  517. Xu, J.L.; Lai, T.; Luo, J.M. Preparation and Characterization of the Aesthetic Coating on Nickel-Titanium Orthodontic Archwire by Electrophoretic Deposition. Prog. Org. Coat. 2019, 137, 105271. [Google Scholar] [CrossRef]
  518. Lee, B.-S.; Lin, Y.-C.; Hsu, W.-C.; Hou, C.-H.; Shyue, J.-J.; Hsiao, S.-Y.; Wu, P.-J.; Lee, Y.-T.; Luo, S.-C. Engineering Antifouling and Antibacterial Stainless Steel for Orthodontic Appliances through Layer-by-Layer Deposition of Nanocomposite Coatings. ACS Appl. Bio Mater. 2020, 3, 486–494. [Google Scholar] [CrossRef] [PubMed]
  519. Szcześ, A.; Hołysz, L.; Chibowski, E. Synthesis of Hydroxyapatite for Biomedical Applications. Adv. Colloid Interface Sci. 2017, 249, 321–330. [Google Scholar] [CrossRef]
  520. Chen, L.; Al-Bayatee, S.; Khurshid, Z.; Shavandi, A.; Brunton, P.; Ratnayake, J. Hydroxyapatite in Oral Care Products—A Review. Materials 2021, 14, 4865. [Google Scholar] [CrossRef] [PubMed]
  521. Jiang, H.C.; Rong, L.J. Effect of Hydroxyapatite Coating on Nickel Release of the Porous NiTi Shape Memory Alloy Fabricated by SHS Method. Surf. Coat. Technol. 2006, 201, 1017–1021. [Google Scholar] [CrossRef]
  522. Kocijan, A. Electrodeposition of a Hydroxyapatite Coating on a Biocompatible NiTi Alloy. Mater. Tehnol. 2018, 52, 67–70. [Google Scholar] [CrossRef]
  523. Surmenev, R.A.; Ryabtseva, M.A.; Shesterikov, E.V.; Pichugin, V.F.; Peitsch, T.; Epple, M. The Release of Nickel from Nickel–Titanium (NiTi) Is Strongly Reduced by a Sub-Micrometer Thin Layer of Calcium Phosphate Deposited by Rf-Magnetron Sputtering. J. Mater. Sci. Mater. Med. 2010, 21, 1233–1239. [Google Scholar] [CrossRef]
  524. Dimasruhin, F.; Ichwan, M.; Nur, A.; Yuwana, M. Electrodeposition of Hydroxyapatite-SiO2 Composite Particles for Biomedical Applications. AIP Conf. Proc. 2014, 1586, 119–123. [Google Scholar]
  525. Kawaguchi, K.; Iijima, M.; Endo, K.; Mizoguchi, I. Electrophoretic Deposition as a New Bioactive Glass Coating Process for Orthodontic Stainless Steel. Coatings 2017, 7, 199. [Google Scholar] [CrossRef]
  526. Kawaguchi, K.; Iijima, M.; Muguruma, T.; Endo, K.; Mizoguchi, I. Effects of Bioactive Glass Coating by Electrophoretic Deposition on Esthetical, Bending, and Frictional Performance of Orthodontic Stainless Steel Wire. Dent. Mater. J. 2020, 39, 593–600. [Google Scholar] [CrossRef] [PubMed]
  527. Rapiejko, C.; Fouvry, S.; Grosgogeat, B.; Wendler, B. A Representative Ex-Situ Fretting Wear Investigation of Orthodontic Arch-Wire/Bracket Contacts. Wear 2009, 266, 850–858. [Google Scholar] [CrossRef]
  528. VDI 2840; Carbon Films—Basic Knowledge, Film Types and Properties. VDI-Rischtlinien 2012; VDI (Verein Deutscher Ingenieure): Düsseldorf, Germany, 2012.
  529. Aisenberg, S.; Chabot, R. Ion-Beam Deposition of Thin Films of Diamondlike Carbon. J. Appl. Phys. 1971, 42, 2953–2958. [Google Scholar] [CrossRef]
  530. Grill, A. Diamond-like Carbon Coatings as Biocompatible Materials—An Overview. Diam. Relat. Mater. 2003, 12, 166–170. [Google Scholar] [CrossRef]
  531. Hauert, R.; Thorwarth, K.; Thorwarth, G. An Overview on Diamond-like Carbon Coatings in Medical Applications. Surf. Coat. Technol. 2013, 233, 119–130. [Google Scholar] [CrossRef]
  532. Fedel, M. Blood Compatibility of Diamond-like Carbon (DLC) Coatings. In Diamond-Based Materials for Biomedical Applications; Narayan, R., Ed.; Woodhead Publishing Limited: Sawston, UK, 2013; pp. 71–102. ISBN 9780857093400. [Google Scholar]
  533. Love, C.A.; Cook, R.B.; Harvey, T.J.; Dearnley, P.A.; Wood, R.J.K. Diamond like Carbon Coatings for Potential Application in Biological Implants—A Review. Tribol. Int. 2013, 63, 141–150. [Google Scholar] [CrossRef]
  534. Robertson, J. Diamond-like Amorphous Carbon. Mater. Sci. Eng. R Rep. 2002, 37, 129–281. [Google Scholar] [CrossRef]
  535. Kobayashi, S.; Ohgoe, Y.; Ozeki, K.; Sato, K.; Sumiya, T.; Hirakuri, K.K.; Aoki, H. Diamond-like Carbon Coatings on Orthodontic Archwires. Diam. Relat. Mater. 2005, 14, 1094–1097. [Google Scholar] [CrossRef]
  536. Kobayashi, S.; Ozeki, K.; Ohgoe, Y.; Gei, L.; Hirakuri, K.K.; Aoki, H. Biocompatibility of Diamond-Like Carbon Coated NiTi Orthodontic Wire and Acrylic Resin Teeth. Key Eng. Mater. 2005, 284–286, 783–786. [Google Scholar] [CrossRef]
  537. Ohgoe, Y.; Hirakuri, K.K.; Ozeki, K.; Fukui, Y. Investigation of Diamond-like Carbon Coating for Orthodontic Archwire. New Diam. Front. Carbon Technol. 2007, 17, 281–288. [Google Scholar]
  538. Ohgoe, Y.; Kobayashi, S.; Ozeki, K.; Aoki, H.; Nakamori, H.; Hirakuri, K.K.; Miyashita, O. Reduction Effect of Nickel Ion Release on a Diamond-like Carbon Film Coated onto an Orthodontic Archwire. Thin Solid Film. 2006, 497, 218–222. [Google Scholar] [CrossRef]
  539. Huang, S.Y.; Huang, J.J.; Kang, T.; Diao, D.F.; Duan, Y.Z. Coating NiTi Archwires with Diamond-like Carbon Films: Reducing Fluoride-Induced Corrosion and Improving Frictional Properties. J. Mater. Sci. Mater. Med. 2013, 24, 2287–2292. [Google Scholar] [CrossRef] [PubMed]
  540. Muguruma, T.; Iijima, M.; Brantley, W.A.; Nakagaki, S.; Endo, K.; Mizoguchi, I. Frictional and Mechanical Properties of Diamond-like Carbon-Coated Orthodontic Brackets. Eur. J. Orthod. 2013, 35, 216–222. [Google Scholar] [CrossRef] [PubMed]
  541. Akaike, S.; Kobayashi, D.; Aono, Y.; Hiratsuka, M.; Hirata, A.; Hayakawa, T.; Nakamura, Y. Relationship between Static Friction and Surface Wettability of Orthodontic Brackets Coated with Diamond-like Carbon (DLC), Fluorine- or Silicone-Doped DLC Coatings. Diam. Relat. Mater. 2016, 61, 109–114. [Google Scholar] [CrossRef]
  542. Muguruma, T.; Iijima, M.; Brantley, W.A.; Mizoguchi, I. Effects of a Diamond-like Carbon Coating on the Frictional Properties of Orthodontic Wires. Angle Orthod. 2011, 81, 141–148. [Google Scholar] [CrossRef]
  543. Akaike, S.; Hayakawa, T.; Kobayashi, D.; Aono, Y.; Hirata, A.; Hiratsuka, M.; Nakamura, Y. Reduction in Static Friction by Deposition of a Homogeneous Diamond-like Carbon (DLC) Coating on Orthodontic Brackets. Dent. Mater. J. 2015, 34, 888–895. [Google Scholar] [CrossRef]
  544. Muguruma, T.; Iijima, M.; Kawaguchi, M.; Mizoguchi, I. Effects of Sp2/Sp3 Ratio and Hydrogen Content on In Vitro Bending and Frictional Performance of DLC-Coated Orthodontic Stainless Steels. Coatings 2018, 8, 199. [Google Scholar] [CrossRef]
  545. Kobayashi, S.; Ohgoe, Y.; Ozeki, K.; Hirakuri, K.; Aoki, H. Dissolution Effect and Cytotoxicity of Diamond-like Carbon Coatings on Orthodontic Archwires. J. Mater. Sci. Mater. Med. 2007, 18, 2263–2268. [Google Scholar] [CrossRef]
  546. Zhang, H.; Guo, S.; Wang, D.; Zhou, T.; Wang, L.; Ma, J. Effects of Nanostructured, Diamondlike, Carbon Coating and Nitrocarburizing on the Frictional Properties and Biocompatibility of Orthodontic Stainless Steel Wires. Angle Orthod. 2016, 86, 782–788. [Google Scholar] [CrossRef]
  547. Tantiwinyupong, N.; Chintavalakorn, R.; Santiwong, P.; Khantachawana, A. Frictional and Mechanical Properties of Surface Modified Nickel-Titanium Orthodontic Archwires. Key Eng. Mater. 2019, 801, 39–43. [Google Scholar] [CrossRef]
  548. Kang, T.; Huang, S.-Y.; Huang, J.-J.; Li, Q.-H.; Diao, D.-F.; Duan, Y.-Z. The Effects of Diamond-Like Carbon Films on Fretting Wear Behavior of Orthodontic Archwire-Bracket Contacts. J. Nanosci. Nanotechnol. 2015, 15, 4641–4647. [Google Scholar] [CrossRef] [PubMed]
  549. Danisman, H.; Celebi, F.; Danisman, S.; Bicakci, A.A. Effects of Diamond-like Carbon Coating on Frictional and Mechanical Properties of Orthodontic Brackets: An in Vitro Study. APOS Trends Orthod. 2021, 12, 13. [Google Scholar] [CrossRef]
  550. Almaguer-Flores, A.; Olivares-Navarrete, R.; Lechuga-Bernal, A.; Ximénez-Fyvie, L.A.; Rodil, S.E. Oral Bacterial Adhesion on Amorphous Carbon Films. Diam. Relat. Mater. 2009, 18, 1179–1185. [Google Scholar] [CrossRef]
  551. Carvalho, I.; Rodrigues, L.; Lima, M.J.; Carvalho, S.; Cruz, S.M.A. Overview on the Antimicrobial Activity and Biocompatibility of Sputtered Carbon-Based Coatings. Processes 2021, 9, 1428. [Google Scholar] [CrossRef]
  552. Almaguer-Flores, A.; Olivares-Navarrete, R.; Ximénez-Fyvie, L.A.; García-Zarco, O.; Rodil, S.E. Biocompatibility and Anti-Microbial Properties of Silver Modified Amorphous Carbon Films. MRS Proc. 2009, 1244, 2. [Google Scholar] [CrossRef]
  553. Wei, S.; Shao, T.; Ding, P. Study of CNx Films on 316L Stainless Steel for Orthodontic Application. Diam. Relat. Mater. 2010, 19, 648–653. [Google Scholar] [CrossRef]
  554. Wei, S.; Shao, T.; Ding, P. Improvement of Orthodontic Friction by Coating Archwire with Carbon Nitride Film. Appl. Surf. Sci. 2011, 257, 10333–10337. [Google Scholar] [CrossRef]
  555. Pan, Z.; Zhou, Q.; Wang, P.; Diao, D. Robust Low Friction Performance of Graphene Sheets Embedded Carbon Films Coated Orthodontic Stainless Steel Archwires. Friction 2022, 10, 142–158. [Google Scholar] [CrossRef]
  556. Wang, P.; Luo, X.; Qin, J.; Pan, Z.; Zhou, K. Effect of Graphene Sheets Embedded Carbon Films on the Fretting Wear Behaviors of Orthodontic Archwire–Bracket Contacts. Nanomaterials 2022, 12, 3430. [Google Scholar] [CrossRef]
Figure 2. Typical Fixed Orthodontic Appliances: (a) by courtesy of Professor Sónia A. Pereira, Faculty of Medicine, University of Coimbra, Portugal; (b) used with permission of Dental Press Publishing from [22]. (i) Brackets with polymeric ligatures; (ii) tube; (iii) bands with attached tubes; (iv) archwires.
Figure 2. Typical Fixed Orthodontic Appliances: (a) by courtesy of Professor Sónia A. Pereira, Faculty of Medicine, University of Coimbra, Portugal; (b) used with permission of Dental Press Publishing from [22]. (i) Brackets with polymeric ligatures; (ii) tube; (iii) bands with attached tubes; (iv) archwires.
Metals 13 01955 g002
Figure 3. Schematic summary of the current review work, concerning corrosion of fixed orthodontic appliances.
Figure 3. Schematic summary of the current review work, concerning corrosion of fixed orthodontic appliances.
Metals 13 01955 g003
Figure 4. Six-phase oral biofilm formation mechanism over tooth and root surfaces, and representative bacterial colonizers. (used with permission of Georg Thieme Verlag KG, from [67]; permission conveyed through Copyright Clearence Center, Inc.).
Figure 4. Six-phase oral biofilm formation mechanism over tooth and root surfaces, and representative bacterial colonizers. (used with permission of Georg Thieme Verlag KG, from [67]; permission conveyed through Copyright Clearence Center, Inc.).
Metals 13 01955 g004
Figure 5. Passivation model of a metallic surface in an electrolyte. Me: metal element; X: complexing anions, e.g., Cl and F. (used with permission of John Wiley & Sons—Books, from [97]; permission conveyed through Copyright Clearance Center, Inc.).
Figure 5. Passivation model of a metallic surface in an electrolyte. Me: metal element; X: complexing anions, e.g., Cl and F. (used with permission of John Wiley & Sons—Books, from [97]; permission conveyed through Copyright Clearance Center, Inc.).
Metals 13 01955 g005
Figure 6. Crevice corrosion detected in a groove on a retrieved archwire (used for at least 6 months) caused by orthodontic pliers (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [106]).
Figure 6. Crevice corrosion detected in a groove on a retrieved archwire (used for at least 6 months) caused by orthodontic pliers (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [106]).
Metals 13 01955 g006
Figure 8. MIC mechanisms on a metallic surface in the oral cavity: (a) cells with varying degrees of oxygenation; (b) cells with different concentration of metal ions; and (c) passive–active cells. EPS: extracellular polymeric substances (reproduced from [73]).
Figure 8. MIC mechanisms on a metallic surface in the oral cavity: (a) cells with varying degrees of oxygenation; (b) cells with different concentration of metal ions; and (c) passive–active cells. EPS: extracellular polymeric substances (reproduced from [73]).
Metals 13 01955 g008
Figure 9. SEM micrographs of retrieved archwires: (a) SS archwires before cleaning, showing plaque and food remnants on the surface and evident marks of the location of the bracket borders (black arrows) (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [106]); (b) SEM micrographs and EDS elemental distribution maps of an NiTi archwire after 17 weeks of intraoral use (reproduced from [116]).
Figure 9. SEM micrographs of retrieved archwires: (a) SS archwires before cleaning, showing plaque and food remnants on the surface and evident marks of the location of the bracket borders (black arrows) (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [106]); (b) SEM micrographs and EDS elemental distribution maps of an NiTi archwire after 17 weeks of intraoral use (reproduced from [116]).
Metals 13 01955 g009
Figure 10. Examples of allergic symptoms related to the use of fixed orthodontic appliances: (ac) extraoral allergic reaction (face and forearms) after placement of fixed appliances, without intraoral symptoms (reproduced from [192] with permission from SNCSC).
Figure 10. Examples of allergic symptoms related to the use of fixed orthodontic appliances: (ac) extraoral allergic reaction (face and forearms) after placement of fixed appliances, without intraoral symptoms (reproduced from [192] with permission from SNCSC).
Metals 13 01955 g010
Figure 11. Schematic representations of the resistance to sliding components in a bracket/archwire contact (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [218]).
Figure 11. Schematic representations of the resistance to sliding components in a bracket/archwire contact (reproduced by permission of Oxford University Press on behalf of the European Orthodontic Society, from [218]).
Metals 13 01955 g011
Figure 12. The two types of friction present when a force is applied to an object. Static friction occurs until a certain limit above which the movement starts. Kinetic friction opposes the on-going movement (reprinted from [215], Copyright (2009), with permission from Elsevier).
Figure 12. The two types of friction present when a force is applied to an object. Static friction occurs until a certain limit above which the movement starts. Kinetic friction opposes the on-going movement (reprinted from [215], Copyright (2009), with permission from Elsevier).
Metals 13 01955 g012
Figure 13. Optical and SEM micrographs of a retrieved SS tube after 2 years of intraoral exposure: (a) cavity filled with organic material; (b) wear track and debris coupled with the EDS chemical composition (reproduced from [116]).
Figure 13. Optical and SEM micrographs of a retrieved SS tube after 2 years of intraoral exposure: (a) cavity filled with organic material; (b) wear track and debris coupled with the EDS chemical composition (reproduced from [116]).
Metals 13 01955 g013
Figure 14. Metal ions (in μg/L) released by wire-bracket combinations in three different solutions (reprinted by permission of Oxford University Press on behalf of the European Orthodontic Society, from [108]; permission conveyed through Copyright Clearance Center, Inc.).
Figure 14. Metal ions (in μg/L) released by wire-bracket combinations in three different solutions (reprinted by permission of Oxford University Press on behalf of the European Orthodontic Society, from [108]; permission conveyed through Copyright Clearance Center, Inc.).
Metals 13 01955 g014
Figure 15. Identification of the bracket parts: wings, slot, and base (reprinted from [319], Copyright (2017), with permission from Elsevier).
Figure 15. Identification of the bracket parts: wings, slot, and base (reprinted from [319], Copyright (2017), with permission from Elsevier).
Metals 13 01955 g015
Figure 16. Summary of biomaterials surface properties (reprinted from [331], Copyright (2002), with permission from Elsevier).
Figure 16. Summary of biomaterials surface properties (reprinted from [331], Copyright (2002), with permission from Elsevier).
Metals 13 01955 g016
Figure 17. Classification of coating processes used at industrial level (adapted from [39]).
Figure 17. Classification of coating processes used at industrial level (adapted from [39]).
Metals 13 01955 g017
Figure 18. SEM micrographs of as-deposited W–N sputtered coatings (a) and its surface morphology after (b) Fusayama–Meyer solution aging (37 °C, 30 days, pH 8.3) coupled with EDS elemental maps distribution (adapted from [468], own work).
Figure 18. SEM micrographs of as-deposited W–N sputtered coatings (a) and its surface morphology after (b) Fusayama–Meyer solution aging (37 °C, 30 days, pH 8.3) coupled with EDS elemental maps distribution (adapted from [468], own work).
Metals 13 01955 g018
Figure 19. Photomicrographs of coated wires: (A) as-received Titanol® Cosmetic, (B,C) post clinical Titanol® Cosmetic, (D) as-received Biocosmetic, (E,F) post clinical Biocosmetic (reproduced from [488]).
Figure 19. Photomicrographs of coated wires: (A) as-received Titanol® Cosmetic, (B,C) post clinical Titanol® Cosmetic, (D) as-received Biocosmetic, (E,F) post clinical Biocosmetic (reproduced from [488]).
Metals 13 01955 g019
Figure 20. Epoxy coating loss on an archwire when exposed to acidic artificial saliva (reprinted from [474], Copyright (2020), with permission from Elsevier).
Figure 20. Epoxy coating loss on an archwire when exposed to acidic artificial saliva (reprinted from [474], Copyright (2020), with permission from Elsevier).
Metals 13 01955 g020
Figure 21. Biofilm retention after 12 and 24 h of incubation as a function of the water contact angle on organosilane coated brackets (used with permission of IOP Publishing, Ltd., from [509]; permission conveyed through Copyright Clearence Center, Inc.).
Figure 21. Biofilm retention after 12 and 24 h of incubation as a function of the water contact angle on organosilane coated brackets (used with permission of IOP Publishing, Ltd., from [509]; permission conveyed through Copyright Clearence Center, Inc.).
Metals 13 01955 g021
Figure 22. A complex polymeric-based composite deposited onto an SS flat substrate to confer the surface with antibacterial and antifouling properties (reprinted with permission from [518]. Copyright (2020) American Chemical Society).
Figure 22. A complex polymeric-based composite deposited onto an SS flat substrate to confer the surface with antibacterial and antifouling properties (reprinted with permission from [518]. Copyright (2020) American Chemical Society).
Metals 13 01955 g022
Figure 23. Low friction mechanism of graphene sheets embedded carbon (GSEC) coating on an SS archwire against an SS bracket: (a) initial state of the contact combination; (b) stable low friction with the formation of graphete-rich tribofilm and salivary adsorbed layer; (c) accumulation of wear debris detached from the GSEC film with micro-groove; and (d) flow out of wear debris with artificial saliva from the micro-groove (reproduced from [556]).
Figure 23. Low friction mechanism of graphene sheets embedded carbon (GSEC) coating on an SS archwire against an SS bracket: (a) initial state of the contact combination; (b) stable low friction with the formation of graphete-rich tribofilm and salivary adsorbed layer; (c) accumulation of wear debris detached from the GSEC film with micro-groove; and (d) flow out of wear debris with artificial saliva from the micro-groove (reproduced from [556]).
Metals 13 01955 g023
Table 3. Composition of several orthodontic stainless-steel grades (reprinted from [10], with permission from Elsevier, after [47] (used with permission of Elsevier Science & Technology Journals; permission conveyed through Copyright Clearance Center, Inc.).
Table 3. Composition of several orthodontic stainless-steel grades (reprinted from [10], with permission from Elsevier, after [47] (used with permission of Elsevier Science & Technology Journals; permission conveyed through Copyright Clearance Center, Inc.).
DesignationComposition (wt.%)
AISIFeCrNiMnMoCPSiSOther
303Bal.17–198–1020.60.150.21.00.15-
304LBal.18–208–122-0.03004100.03-
316LBal.16–18101422.50.030.041.00.03-
630/17-4Bal.15–173–51-0.070.041.00.044 Cu/3 Nb
630/17-7Bal.16–186.5–7.51-0.090.041.00.040.08–1.5 Al
SAF 2205Bal.225.5230.030.031.00.020.16 N
18–8 PlusBal.80.161810.150.0451.00.030.5 N
431Bal.26--4-----
AI29Bal.290.30.540.020.0350.350.010.5 Ti
Table 4. Main functions assigned to saliva and its constituents [69,70,71,72].
Table 4. Main functions assigned to saliva and its constituents [69,70,71,72].
FunctionDescriptionAgents
Tissue lubrication, repairing, and protectionSeromucous covering of the oral tissues.
Barrier against irritants.
Lubrication of hard and soft tissues, and prosthesis.
Mastication, speech, and deglutition aid due to lubrication.
Selective modulation of microbial adhesion to oral tissues. Modulation of dental plaque metabolism.
Faster tissue repair.
Mucins and other proteins.
Clearance and pH maintenanceAcids neutralization (e.g., bicarbonate buffer).
Alkalinization of dental plaque’s pH through urea metabolization by its microbiome.
pH modulation to prevent reaching optimal conditions for oral colonization by pathogens.
Bicarbonate, phosphate, urea, amphoteric proteins, and enzymes.
Maintenance of dental integrityModulation of pathogens activity to control the progression of caries and enamel damage.
Maintenance of the enamel mineralization/demineralization equilibrium.
The presence of fluoride in saliva enhances mineralization and forms a fluorapatite-like coating, which is more resistant to caries than the original teeth material.
Calcium, phosphate, fluoride, and several proteins (including statherin, histatins, cystatins, and proline-rich proteins).
Antibacterial activitySelective action of protein-based immunological and non-immunological agents, allowing the growth of non-cariogenic microorganisms.
Among other mechanisms, the non-immunological action involves the adhesion inhibition of colonizers to the oral tissues, namely by aggregation (clumping).
Immunoglobulins, enzymes, and other proteins (including glycoproteins, staherins, agglutinins, histidine-rich proteins, and proline-rich proteins).
Digestion, taste, and smellBesides lubricating food and tissues, saliva starts the chemical oral digestion, namely by the initial action of the α-amylase (converting complex carbohydrates into simple sugars).
The hypotonicity of saliva (low sodium, glycose, bicarbonate, and urea levels) regarding plasma, which enhances the dissolution of the substances.
The presence of proteins (such as gustin) is necessary to the growth of gustatory buds.
α-amylase, gustin, lipases and other proteins.
Table 5. Oral microbiome (reproduced from [73]).
Table 5. Oral microbiome (reproduced from [73]).
Oral Bacteria Microbiome
SalivaActinobacteria, Bacteroides, Firmicutes, Fusobacteria, Proteobacteria, Spirochaetes, TM7 (The Human Microbiome Consortium)
Dental plaqueFirmicutes, Actinobacteria
Oral mucosaStreptococcus salivarius, Rothia mucilaginosa, Eubacterium strain FTB41
Oral Bacteria Related to Oral Diseases
Dental cariesStreptococcus, Veillonella, Actinomyces, Granulicatella, Leptotrichia, Thiomonas, Bifidobacterium, Prevotella, Lactobacillus, Propionibacterium, Pseudoramibacter, Selenomonas
Periapical infections (periapical periodontitis, root canal infection)Proteobacteria, Firmicutes, Bacteroidetes, Fusobacteria, Actinobacteria, Olsenella uli, Prevotella baroniae, Porphyromonas endodontalis, Fusobacterium nucleatum, Tannerella forsythia, Propionibacterium propionicum, Porphyromonas gingivalis, Prevotella intermedia, Prevotella oralis, Parvimonas micra, Porphyromonas endodontalis, Fusobacterium nucleatum, Tannerella forsythia
Periodontal diseases (gingivitis, periodontitis)Actinomycetes, Capnocytophaga, Campylobacter, Eikenella, Fusobacterium, Prevotella, Porphyromonas gingivalis, Treponema denticola, Tannerella forsythia, Bacteroidetes spp., Eubacterium saphenum, Porphyromonas endodontalis, Prevotella denticola, Parvimonas micra, Peptostreptococcus spp., Filifactor alocis, Desulfobulbus spp., Dialister spp., Synergistetes
HalitosisSolobacterium moorei, Atopobium parvulum, Eubacterium sulci
Table 7. Ni-containing orthodontic materials and the corresponding Ni-free substitutes (used with permission of The Edward H. Angle Education and Research Foundation, from [28]; permission conveyed through Copyright Clearence Center, Inc.).
Table 7. Ni-containing orthodontic materials and the corresponding Ni-free substitutes (used with permission of The Edward H. Angle Education and Research Foundation, from [28]; permission conveyed through Copyright Clearence Center, Inc.).
CategoryMaterialNi-Free Substitute and Modifications
Standard appliancesBrackets.Ni-free SS, ceramic, plastic, Ti, gold-plated or coated with other precious metals (Pd, Pt) brackets.
Treatment utilitiesBands.
SS archwires.
Gold-plated bands.
No alternative currently available; development of polymeric wires in progress.
Mechanic helpersCoCrNi archwires.
Sliding yokes, transpalatal. and lingual arches.
No alternative currently available
β-Ti (TMA), plastic or inert metal (gold) coatings of wire segments.
Miscellaneous helpersSS ligatures.
Kobayashi hooks.
Coil springs.
Teflon-coated ligatures.
Teflon-coated Kobayashi hooks; Ni-free brackets with hooks.
Elastomeric ligatures.
Fixed expansion appliancesSS appliances (Quad Helix). Rapid palatal expander.
SS headgear.
NiTi spring screws.
β-Ti (TMA) wires for Quad-Helix.
Teflon-coated SS facebow.
No alternative currently available.
Removable appliancesSS components of Hawley appliances and variations.Plastic or elastic retainers; elastic positioners or acrylic splints invisagenTM technique.
Complex therapeutic interventionsOrthognathic surgery lag screws and plates.
Distraction osteogenesis apparatus.
Resorbable polylactic-polyglycolic lag screws and plates.
No alternative currently available.
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

Fróis, A.; Santos, A.C.; Louro, C.S. Corrosion of Fixed Orthodontic Appliances: Causes, Concerns, and Mitigation Strategies. Metals 2023, 13, 1955. https://doi.org/10.3390/met13121955

AMA Style

Fróis A, Santos AC, Louro CS. Corrosion of Fixed Orthodontic Appliances: Causes, Concerns, and Mitigation Strategies. Metals. 2023; 13(12):1955. https://doi.org/10.3390/met13121955

Chicago/Turabian Style

Fróis, António, Ana Cristina Santos, and Cristina Santos Louro. 2023. "Corrosion of Fixed Orthodontic Appliances: Causes, Concerns, and Mitigation Strategies" Metals 13, no. 12: 1955. https://doi.org/10.3390/met13121955

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