Next Article in Journal
Experimental Measurement of the Thermal Conductivity of Fused Deposition Modeling Materials with a DTC-25 Conductivity Meter
Next Article in Special Issue
Recent Advances in Quaternary Ammonium Monomers for Dental Applications
Previous Article in Journal
A Simple Approach for Regenerating Electrolyzed Hydrogen Production Using Non-De-Ionized Water Sources
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Dental Implants: Modern Materials and Methods of Their Surface Modification

by
Catherine Sotova
1,*,
Oleg Yanushevich
2,
Natella Kriheli
2,
Sergey Grigoriev
1,
Vladimir Evdokimov
2,
Olga Kramar
2,
Margarita Nozdrina
2,
Nikita Peretyagin
1,2,
Nika Undritsova
1,2,
Egor Popelyshkin
1 and
Pavel Peretyagin
1,2
1
Department of High-Efficiency Machining Technologies, Moscow State University of Technology “STANKIN”, Vadkovsky Lane 3a, 127055 Moscow, Russia
2
Scientific Department, A.I. Evdokimov Moscow State University of Medicine and Dentistry, Delegatskaya St., 20, p.1, 127473 Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2023, 16(23), 7383; https://doi.org/10.3390/ma16237383
Submission received: 26 October 2023 / Revised: 20 November 2023 / Accepted: 21 November 2023 / Published: 27 November 2023
(This article belongs to the Special Issue Research Progress in Functional Dental Materials)

Abstract

:
The development of dental implantology is based on the detailed study of the interaction of implants with the surrounding tissues and methods of osteogenesis stimulation around implants, which has been confirmed by the increasing number of scientific publications presenting the results of studies related to both the influence of the chemical composition of dental implant material as well as the method of its surface modification on the key operational characteristics of implants. The main materials for dental implant manufacturing are Ti and its alloys, stainless steels, Zr alloys (including ceramics based on ZrO2), and Ta and its alloys, as well as other materials (ceramics based on Al2O3, Si3N4, etc.). The review presents alloy systems recommended for use in clinical practice and describes their physical–mechanical and biochemical properties. However, when getting into the body, the implants are subjected to various kinds of mechanical influences, which are aggravated by the action of an aggressive biological environment (electrolyte with a lot of Cl and H+); it can lead to the loss of osteointegration and to the appearance of the symptoms of the general intoxication of the organism because of the metal ions released from the implant surface into the biological tissues of the organism. Since the osteointegration and biocompatibility of implants depend primarily on the properties of their surface layer (it is the implant surface that makes contact with the tissues of the body), the surface modification of dental implants plays an important role, and all methods of surface modification can be divided into mechanical, physical, chemical, and biochemical methods (according to the main effect on the surface). This review discusses several techniques for modifying dental implant surfaces and provides evidence for their usefulness.

1. Introduction

According to the World Health Organization (WHO), oral diseases are prevalent noncommunicable diseases that afflict almost half of the world’s population, i.e., around 45% or 3.5 billion individuals of all ages [1,2]. Complete tooth loss, or adentia, is steadily increasing with a peak occurrence in older age groups. Currently, there are over 350 million cases worldwide, resulting in a global prevalence rate of almost 6.82% [1]. In Russia, partial tooth loss is estimated to affect between 40% and 75% of the population, corresponding to a range of 58 million to 108 million people. And, the number of patients with such problems will increase as the average age of the population rises [3]. In addition to tooth loss caused by accidents (trauma), dental decay and gum diseases are key factors contributing to teeth loss. Untreated caries in permanent teeth affect 2.3 billion individuals, while over 514 million children suffer from untreated deciduous teeth caries, and more than 1 billion people, which represent 19% of the world’s population over 15 years of age, have periodontal disease [1].
Thus, the development of the dental implants market is driven by the high prevalence of oral diseases and the goal of improving people’s quality of life. The worldwide market for dental implants was estimated at $9.27 billion in 2022 and $10.09 billion in 2023, with a forecasted compound annual growth rate (CAGR) of 8.95% to hit $18.42 billion by 2030 [4]. The dental implants market is growing due to a rise in tooth loss cases, increasing demand for cosmetic dentistry, and advancements in dental implant technology [5].
The advancement of dental implantology is based on a detailed study of the interaction of implants with surrounding tissues and methods of stimulating osteogenesis around implants. This has been demonstrated by the increasing number of scientific articles presenting the results of studies on the influence of the chemical composition of the dental implant material and the method of its surface modification on the key characteristics of dental implants (Figure 1). From 1989 (the year of the first patent registration for a titanium implant) to the present day, a search query with the keywords “dental implant” using bibliographic databases Scopus and ScienceDirect yielded 89,535 publications. Over the last decade (from 2012 to 2022), the number of publications has more than doubled (from 42,762 to 89,535).
Figure 2 and Figure 3 display the word cloud generated by keywords from the previously mentioned sample, limited to the period from 2014 to the present, where keywords related only to implant material and surface modification methods and properties were selected. According to the results presented, the main requirements for implants are to provide and/or enhance the following properties (these have been the focus of scientific research on the improvement of dental implants):
  • Osseointegration, or the fusion of the implant surface with the bone [6,7,8,9,10];
  • Bone regeneration, the process by which the body builds new bone tissue [11,12,13,14,15];
Figure 2. Overlay visualization of word cloud based on the keywords (dental implant) of publications from 2014 to present (according Scopus, ScienceDirect).
Figure 2. Overlay visualization of word cloud based on the keywords (dental implant) of publications from 2014 to present (according Scopus, ScienceDirect).
Materials 16 07383 g002
  • Antibacterial ability, the ability of the implant to inhibit or slow down bacterial growth [16,17,18,19,20];
  • Biocompatibility, the ability of the material to be integrated into the body without clinical complications and to induce the required cellular or tissue response [9,21,22,23,24,25].
Based on their biocompatibility, materials for dental implants can be divided into biotolerant, bioinert, and bioactive [26]. Biotolerant materials, such as stainless steel and cobalt–chromium alloys, induce osteogenesis in the bone to respond to the irritating effect of the implant in the tissue contact zone. In this case, a layer of soft fibrous tissue separates the bone from the implant composed of these materials. The application of bioinert materials, including alumina, zirconia, titanium, its alloys, and tantalum, leads to the development of contact osteogenesis given favorable mechanical conditions. This means that these materials directly bond with the bone tissue. Bone integration occurs because the surface of such materials is chemically inert to the surrounding tissues and fluids. Bioactive materials such as calcium phosphate ceramics, glass, and glass ceramics cause connective osteogenesis—the direct chemical bonding of the implant with the surrounding bone—due to the presence of free calcium and phosphate on the surface and of their interaction.
In addition, a number of scientific studies have been devoted to the influence of the above-mentioned factors on other properties of implants, for example:
  • Bone resorption—the destruction (resorption, degradation) of bone tissue under the action of osteoclasts [27,28,29];
  • Biomechanical stability—the ability of the implant to be reliably fixed in the bone tissue without clinical mobility [30,31,32];
  • Osteoconductivity—the ability of the implant to provide bone tissue formation and growth on its surface (a special case of osteointegration) [33,34,35];
Figure 3. Density visualization of word cloud based on the keywords (dental implant) of publications from 2014 to present (according Scopus, ScienceDirect).
Figure 3. Density visualization of word cloud based on the keywords (dental implant) of publications from 2014 to present (according Scopus, ScienceDirect).
Materials 16 07383 g003
  • Osteoinductivity—the ability of the implant surface to induce differentiation in mesenchymal cells within the surrounding tissues, resulting in the formation of osteoblasts (a special case of osteointegration) [36,37,38];
  • Osteogenesis—the ability of osteoblasts to form a new bone matrix (a special case of biocompatibility) [30,39,40,41];
  • Osseoperception—the sensory feedback of sensorimotor, orofacial, and masticatory functions resulting from the direct transmission of forces and vibrations to bone tissue [42,43,44];
  • Corrosion (or tribocorrosion) resistance—the ability of the implant not to be destroyed under the influence of the biological environment (saliva, blood, etc.) (reviewed in [45,46,47,48,49,50]).
Currently, according to numerous fundamental and applied studies, the main materials for dental implants are titanium and its alloys, stainless steels, zirconium alloys (including ceramics based on zirconium dioxide), and tantalum and its alloys, as well as other materials (ceramics based on aluminum oxide, silicon nitride, etc.). It should be noted that due to the lowest osseointegration and the presence of harmful impurities, the number of publications where the objects of research are implants made of stainless steels is significantly reduced (Figure 4). The great variety of dental implants based on chemical composition and surface layer modification technology indicates that the problem of choosing the optimal material for implant manufacturing has not been finally solved to date.
The purpose of this study was to analyze the different materials used in dental implantation at present (namely, for the production of dental implants representing the root of an artificial tooth), as well as to find ways to improve their performance properties by modifying the implant surface.
The material of the research comprised the literature data presented in scientific publications indexed in the bibliographic databases PubMed, ScienceDirect, and Scopus, using, as search terms, the keywords “dental implant*” with specifications on different groups of materials and methods of additional implant processing (surface modification techniques). The end date of the search was August 2023.

2. Dental Implant Materials

An artificial tooth (denture) is a complex construction (Figure 5) consisting of an implant (1), abutment (2) and crown (3). The implant itself is completely immersed deep into the bone, and the crown is located above the gingiva; the abutments connect them with each other, except in the use of one-piece implants, in which the intraosseous part and the abutment are immediately connected into a monolithic element (one-piece type of implants). However, this variant is mainly used for full jaw restorations, i.e., basal implantation (in one-stage implantation protocols involving immediate loading with a prosthesis).
The implant material must meet several requirements [51]:
  • To be strong enough to withstand chewing pressure (sometimes exceeding 100 kgf/cm2) and not to destroy the bone with its weight;
  • To be well machinable at the manufacturing stage in order to retain its shape throughout its entire service life;
  • To not be destroyed (not corroded) by the action of the biological environment (saliva, blood, etc.);
  • To not show toxic, allergenic, and carcinogenic effects on the body;
  • To not provoke an increase in galvanic currents when interacting with metal structures installed in the patient’s mouth.
The main materials for the production of dental implants, as mentioned before, are mainly metal alloys based on titanium, iron, and tantalum, as well as ceramics based on zirconium (Figure 6). At the same time, starting in the last twenty years, the number of studies in which the implants made of stainless steels are the objects of studies has considerably decreased (from 38% to 20%), and this has been connected with the presence of harmful impurities in the steel composition and the insufficient corrosion resistance and biocompatibility of these implants. At the same time, titanium and its alloys, during this period, have steadily occupied the leading positions (~50%), and today, they are the main materials for the production of commercialized dental implants.

2.1. Titanium and Its Alloys

The main materials for the production of dental implants are titanium and its alloys. For commercialized dental implants, the following grades of alloys are mainly used [52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67]:
  • Grade 2 (cp-Ti) [53,54,55,56,57]—commercial pure titanium containing up to 0.3% iron;
  • Grade 4 (cp-Ti) [57,58,59,60,61], which contains about 99% titanium and up to 0.5% iron;
  • Grade 5 (Ti-6Al-4V alloy) [62,63,64,65,66,67], which is an alloy of titanium (88%) with aluminum (up to 6.75%) and vanadium (up to 4.5%);
  • Roxolid®, which is a high-strength alloy of titanium (about 85%) with zirconium (~13–17%), specially developed for use in dental implantology [68,69,70,71].
The chemical and mechanical properties of these alloys meet the requirements for materials used in dental implantation. They are characterized by high strength at low density, high elasticity (five times higher than the elasticity of bone [26]), and bioinertness [72], but not all titanium alloys are used in dentistry. However, with all these positive properties of titanium and its alloys, they are characterized by low resistance to shear and wear, especially under friction conditions [26].
At the same time, dental implants should be designed with a high coefficient of friction and a low modulus of elasticity to avoid excessive stress on the bone [73]. The modulus of elasticity of solid titanium alloys is 100–115 GPa, while the values for different types of bone are 0.5–20 GPa. This mechanical mismatch is the reason why healthy bones are underloaded and the “stress shielding” effect occurs [74]. The “stress shielding” effect leads to the healthy bone resorption and loosening of the implant, as a result of which the implant functioning period decreases.
This is why studies on obtaining “new”-generation titanium alloys with mechanical properties approaching the properties of bone tissue are conducted widely enough. All of them are aimed at the volumetric alloying of titanium alloys with β-stabilizers (Nb, Ta, Zr, and Mo) to obtain a single-phase β-alloy structure since β-Ti alloy has a lower modulus of elasticity and higher strength than α-Ti alloys and (α+β)-Ti alloys [75]. In addition, the β-type titanium alloy has good technological properties (in particular, cold-pressure machinability), which reduces the cost of implant production [76].
Table 1 shows the compositions of titanium alloys, used for the production of dental implants, identified by the above search query.
Ti-6Al-4V alloy is 3.0–3.5 times stronger than commercially pure titanium and cheaper to produce [136], which makes it indispensable in the manufacture of thin, reliable implants with special compression threads that are placed in the dense basal regions of the jawbone [52]. However, vanadium and aluminum contained in the alloys have been shown [137,138] to have toxic effects on biological entities and can lead to long-term health problems, such as neurological diseases and Alzheimer’s disease [139,140,141,142], while aluminum and iron (although it is not a toxic element) lead to the formation of a connective tissue layer around the implant and to significant tissue contamination, which is a sign of the insufficient bioinertness of the metal. In addition, iron suppresses the growth of organic cultures. Also, the degree of tissue adhesion to the implants made of titanium alloys is somewhat worse than that to unalloyed titanium [143].
The introduction of Nb into the alloy composition, instead of toxic vanadium, improves the bioinertness of alloys and leads to a significant increase in the wear resistance of alloys of the Ti-Al-Nb system in comparison with Ti-Al-V alloys, with comparable values in terms of fatigue strength and the service lives of products. Thus, it was shown in [77] that the wear resistance of Ti-21Al-29Nb and Ti-15Al-33Nb alloys was more than three times higher than that of Ti-6Al-4V. At the same time, at low- and multi-cycle fatigue tests on samples from both alloys, the following were observed: the nucleation of surface cracks, furrows in the zone of propagation of a stable fatigue crack, and equiaxed dimples in the zone of propagation of a fast-fatigue crack. The Ti-15Al-33Nb alloy exhibited ductile fracture morphology whereas the Ti-21Al-29Nb alloy tended to exhibit a more brittle fracture morphology.
The introduction of yttrium-stabilized zirconium oxide into the alloy of the Ti-Al-Nb system increases the Vickers microhardness and biocorrosion resistance of the alloy and improves the corrosion resistance of the alloy [81].
In recent years, non-toxic titanium alloys with a lower modulus of elasticity and that are free of aluminum and vanadium, such as Ti-Nb, Ti-Zr, Ti-Ta, Ti-Fe, Ti-Mo, and others, have been used as alternatives to Ti-Al-based alloys (see Table 1). Among them, Ti-Nb-based alloys have a low elastic modulus, shape memory, and hyperelasticity, so these alloys are chosen as the base alloys for alloying studies by most biomedical materials researchers [53,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98]. Thus, the elastic modulus of Ti-Nb alloys depends on the Nb content and has two minima: ~ 68 GPa at 15 wt.% Nb (the alloy has an (α+β)-structure) and 62 GPa at 42 wt.% Nb (the alloy has a β-structure) [82] (for comparison, the elastic modulus of Ti-6Al-4V alloy is ~ 115 GPa [94]).
Moreover, Nb is non-toxic and has no harmful effect on the human body, which can be a result of its ability to form a protective oxide film on the implant surface like titanium. According to a study [144], Nb demonstrated good compatibility in contact with cells, providing mitochondrial activity and cell growth. Moreover, Ti-Nb alloys release fewer metal ions into the surrounding tissues compared to Ti-Al-V, Ti-Ni, and Ti-Mo alloys [83,145]. Ti-Nb alloys also have good mechanical strength (e.g., for Ti-42Nb alloy, the tensile strength is 683.17 ± 16.67 MPa and the compressive strength is 1330.74 ± 53.45 MPa [82]) and hardness (microhardness increases with increasing Nb content in the alloy—276 HV0.5 for Ti-13Nb and 287 HV0.5 for Ti-28Nb [146]) because Nb dissolves in the Ti crystal lattice, forming a solid substitution solution, which leads to the hardening of the solid solution β [147].
While, in binary Ti-Zr alloys, zirconium has almost no effect on β-phase stabilization, in multicomponent Ti alloys containing Nb or Ta, Zr acts as an effective β-stabilizer [76]. At the same time, Ti-Nb-Zr (TNZ) alloys have high strength (the tensile strengths of the alloys are in the range of 704–839 MPa) while remaining at the level of Ti-Nb alloys in terms of their low modulus of elasticity (in the range of 62–65 GPa) [76]. In addition, the new series of β-TNZ alloys have excellent cytocompatibility [76,85]. According to [85], due to its good mechanical properties (E = 84.1 GPa), high corrosion resistance, and lack of cytotoxicity toward MC3T3 and NHDF cells, this Ti-13Nb-13Zr alloy can be successfully used in implants, including in bone tissue engineering and dental products.
The additional alloying of a TNZ alloy with tantalum (TNZT alloy) leads to an increase in the corrosion resistance of the alloy in the environment of biological fluids (e.g., saliva, biofilm, and fluoride) while maintaining (with a decreasing trend) a relatively low modulus of elasticity (E = 82 GPa) [88]. The excellent biological and corrosion performance results are achieved mainly due to the oxide film (TiO2, Nb2O5, ZrO2, or Ta2O5) spontaneously formed on Ti and its alloys when exposed to atmospheric air [88]. The introduction of silicon into the composition of the TNZT alloy makes it possible to further reduce the elastic modulus (up to 50 GPa) [90]. At the same time, the Ti-Nb-Zr-Ta-Si alloy has a significant compatibility with bone tissue in comparison with the cp-Ti alloy, which is widely used in dental implants. Thus, in in vitro tests, in comparison with cp-Ti, Ti-Nb-Zr-Ta-Si alloy showed better adhesion, proliferation, and alkaline phosphatase (ALP) activity and promoted the expression of osteocalcin (OCN) mRNA in MG63 cells [90].
The addition of Mo not only stabilizes the β-phase of alloys but also increases the strength and maintains the ductility of Ti-Nb alloys, while Ti-Nb-Mo alloys are not prone to riveting [75]. This leads to a decrease in the modulus of elasticity, which is lower the more Mo is present in the alloy composition. Thus, the reduced elastic modulus (Er measured via nanoindentation) decreases from 67.0 GPa for the Ti-26Nb-2Mo alloy to 54.5 GPa for the Ti-26Nb-8Mo alloy (with the lowest elastic modulus) [75]. In addition, the Ti-26Nb-8Mo alloy has good wear resistance and higher impact toughness than the cp-Ti alloy.
Despite the biocompatible composition, improved corrosion resistance, and low modulus of elasticity of complex-alloyed titanium alloys, serious aspects limiting the effectiveness of titanium alloys for biomedical applications are low tribocorrosion resistance in biological body environments and an inability to prevent bacterial infection in the early stages of recovery [148,149]. The latter can be addressed by alloying with antibacterial components such as Cu, Zn, Ag, and Ga [150].
In one study, although the addition of alloying elements Ga and/or Cu to the Ti-Nb matrix resulted in a slight increase in the elastic modulus to 73–78 GPa (for the base alloy Ti-45Nb, E = 64 GPa), the elastic modulus was still lower than that of the alloys used in clinical practice [93]. Among β-Ti-Nb-Ga alloys, increased wear resistance was noted for the Ti-45Nb-8Ga alloy [94], which is associated with increased microhardness (it increased by 32–44% with the addition of Ga and/or Cu, which the authors [93] attributed to the hardening of solid solutions). Also, the addition of Ga to Ti-45Nb leads to both improved corrosion resistance under mechanical loading as well as the increased strength of the alloy [94].
Another group of alloys used for the fabrication of dental implants, which have been clinically approved and already successfully commercialized, comprises Ti-Zr alloys [68,69,70]. For example, Ti-Zr alloys can compete in the fabrication of small-diameter dental implants. In these cases, cp-Ti is more prone to fracture (due to its relatively low compressive strength), and the use of Ti-Al-V alloy is undesirable due to the toxicity associated with the release of Al and V.
For binary Ti-Zr alloys (as mentioned earlier), Zr is a neutral component with respect to titanium and easily dissolves in both α- and β-solid substitution solutions of unlimited solubility of Ti(Zr) [151] (while strengthening it due to distortion of the crystal lattice of titanium when its atoms are replaced by zirconium atoms [152,153]) since it undergoes a similar allotropic transformation at a close phase transition temperature [154].
Ti-Zr alloys have a predominantly α-crystalline structure, which predetermines the enhanced mechanical properties and excellent electrochemical properties of these alloys [150,153]. Thus, the addition of 5 and 10 wt.% Zr to Ti allowed more than doubling the microhardness of the alloy (HV 416–434) compared to the microhardness of pure titanium cp-Ti (HV ~188), surpassing even the hardness of the Ti-6Al-4V alloy (HV 354) [155]. The tensile strength of binary Ti-Zr alloys also increases with increasing zirconium concentration in the alloy. In [101], it was shown that the tensile strength of Ti-15Zr alloy exceeds that of cp-Ti by about 10–15% (950 MPa vs. 860 MPa, approximately), which the authors attributed, first, to the hardening of the solid solution due to alloying, second, to the refinement of its grain size (1–2 μm in Ti-15Zr as opposed to 20–30 μm in cp-Ti), and third, to the riveting obtained during sample fabrication (strain hardening). However, Ti-Zr alloys have an increased modulus of elasticity, and the modulus of elasticity increases with increasing zirconium concentration in the alloy (for Ti-5Zr, Ti-10Zr, and Ti-15Zr alloys the values were ~86, ~95, and ~110 GPa, respectively [152]), but for Ti-Zr alloy with a Zr concentration of up to 10 wt.%, the modulus of elasticity was significantly lower than for the widely used Ti-6Al-4V alloy, which had a higher elastic modulus (~115 GPa [94]). At the same time, the minimum modulus of elasticity has an alloy containing about 7.5 wt.% Zr [152].
Also, Ti-Zr alloys demonstrate increased corrosion resistance in comparison with traditional titanium alloys, which can be associated with the formation of passivating ZrO2 film on the implant surface. Thus, Zr for Ti is an anodic alloying component, which directly reduces its anodic activity [156], i.e., ZrO2 is a more stable oxide than TiO2. And, although Ti4+ ions are more mobile than Zr4+ ions, which leads to a significant reduction in the amount of ZrO2 oxide in the outer surface layer, the amount of the latter is proportional to the concentration of Zr in the alloy [155]. Higher corrosion protection for Ti-Zr alloys may also be facilitated by their increased hardness (and therefore wear resistance), which may prevent damage to the passivation film by mechanical stresses and ensure long-term rehabilitation success by reducing both the likelihood of corrosion in physiologic environments and the likelihood of impaired osseointegration [157]. In addition, it was shown in a study [155] that higher Zr concentrations resulted in enhanced albumin adsorption (albumin adsorption values for cp-Ti, Ti-6Al-4V, Ti-5Zr, and Ti-10Zr were 600, 650, 650, 750 mg/mL, respectively), suggesting, according to the authors, that there was no negative effect on initial cell adhesion.
Due to the sensitivity of mechanical properties to the structure of binary Ti-Zr alloys, it is possible to reduce the modulus of elasticity by applying strengthening heat treatment (accelerated cooling after heating above the β→α transformation temperature) or strain hardening, which will change the distance between atoms, and this, in turn, can lead to a change in the bonding force between atoms and, as a result, the modulus of elasticity [152,158]. In addition, the use of additive technologies to create porous structures can also contribute to a decrease in the elastic modulus while maintaining the relatively high strength of the alloys. For this connection, additional studies are needed to reveal the influence of technologies for producing products from binary Ti-Zr alloys on their mechanical properties.
The additional alloying of Ti-Zr alloys with β-stabilizers (Nb, Mo and/or Ta) leads to a decrease in the elastic modulus [104,105,155] since the introduction of these components promotes the formation of β-phase in the structure, characterized by the lowest elastic modulus among all phases of titanium alloys [159]. Nb reduces the hardness of Ti-Zr alloy to a value close to the value of cp-Ti (~ 200 HV) [155], which is also associated with the appearance of a softer β-phase in the structure. The influence of zirconium concentration on the hardness of ternary Ti-Nb-Zr alloys is similar to that of binary Ti-Zr alloys; a study [155] showed the higher hardness of Ti-Nb-Zr alloys with higher Zr concentrations (in the range of zirconium concentrations of up to 10 wt.%).
However, as shown by studies, in complex alloys such as Ti-Zr-Nb-Mo, increasing the concentration of Zr from 34 to 40% (by replacing it with niobium while keeping the content of titanium and molybdenum constant) leads to a decrease in hardness (243 ± 4, 2 HV, 238 ± 2.74 HV, and 237 ± 3.07 HV for Ti-34Zr-17Nb-1Mo, Ti-37Zr-14Nb-1Mo, and Ti-40Zr-11Nb-1Mo alloys, respectively) and alloy elastic modulus (75, 72, and 69.5 GPa for Ti-34Zr-17Nb-1Mo, Ti-37Zr-14Nb-1Mo, and Ti-40Zr-11Nb-1Mo alloys, respectively) [105] (in contrast to alloys with low zirconium content, i.e., up to 15 wt.%, for which hardness and elastic modulus increase with an increasing Zr concentration in the alloy [152]).
Biosafety evaluation tests (in accordance with the ISO 10993 series, [160]) of Ti-Zr alloyed with Nb and Ta [104] did not reveal any adverse (negative) effects either during extraction simulating normal use or under conditions of exaggerated extraction, which the authors attributed to the small number of released Ti ions. Thus, the concentrations of Ti in 0.9% NaCl + HCl solution when the samples from Ti-15Zr-4Nb, Ti-15Zr-4Nb-1Ta, and Ti-15Zr-4Nb-4Ta alloys were immersed in it were 0.4, 0.27, and 0.32 μg/mL, respectively. The authors of the study [104] substantiated such results via the presence of a passivating TiO2 film formed on the Ti alloy surface and containing Nb2O5-, ZrO2-, and Ta2O5- oxide films, which inhibited the yield of metal ions.
Another study [106] showed the prospects of the biomedical application of Ti-Zr alloys additionally alloyed with Fe. The Ti-6Zr-xFe alloy has an (α+β)-structure. With an increasing Fe content, from 4 to 7 wt.%, the volume fraction of α-phase decreases, the fraction of β-phase increases, and the grain size of the alloy as a whole decreases. This leads to an increase in microhardness (264–348 NV), the tensile strength (748–994 MPa) of cast Ti-6Zr-xFe alloys due to the hardening of the solid solution (due to iron alloying) and formation of ω-phase, and elastic modulus (90–94 GPa). At the same time, the alloys Ti-6Zr-4Fe and Ti-6Zr-5Fe (with the lowest Fe content) showed better corrosion resistance.
In each of the binary titanium alloys Ti-Ta and Ti-Mo, alloying with tantalum or molybdenum leads to higher strength and lower modulus of elasticity in the alloy compared to cp-Ti (similar to the effect of niobium) since Ta and Mo are also β-stabilizers. Thus, in a study, Ti-40Ta and Ti-50Ta alloys had higher tensile strength values compared to Ti-6Al-4V (786, 724, and 689 MPa, respectively), which were 14% and 8.9% higher than that of Ti-6Al-4V material [161], and the Ti-40Ta alloy with a biomimetic lamellar structure (obtained using sequential spark plasma sintering at 1200 °C followed by hot rolling at a strain rate of 60% and annealing) had a suitable combination of strength (tensile strength of 980 MPa, which was much higher than cp-Ti and comparable to Ti-6Al-4V) and low modulus of elasticity (80.4 GPa) [108]. It is worth noting that the strength and modulus of elasticity of Ti-Ta alloys are very sensitive to the Ta content; the dependence of these properties on the tantalum content in the binary Ti-Ta alloy has a complex character. Thus, the elastic modulus of the alloy first decreases almost linearly with an increasing Ta content and reaches a local minimum of 69 GPa at 30% Ta, then gradually increases and reaches 88 GPa at 50% Ta, and then gradually decreases and reaches a second local minimum of 67 GPa at 70% Ta. A further increase in Ta content leads to an increase in the elastic modulus, approaching the modulus of pure Ta [162]. The opposite dependence is observed for the tensile strength: at first, with an increasing Ta content, the strength increases, and the tensile strength reaches the first local maximum (595 MPa) at 30% Ta, then the strength slightly decreases to 530 MPa at 50% Ta. After that, it increases again and reaches the peak value of 690 MPa at 60% Ta; a further increase in Ta content slightly decreases the strength of the alloys [162]. The authors of [162] attribute these changes to the influence of tantalum concentration on the structure of binary Ti-Ta alloys (in their study, the alloys had a hexagonal (α’) martensitic structure with a Ta content of up to 20%, orthorhombic (α”) martensitic structure at a Ta content from 30 to 50%, (α” +β)-structure in the alloys with 60% Ta, and single-phase metastable β structure at a Ta content of more than 60%). In addition to tantalum concentration, the structure and, consequently, the properties of binary Ti-Ta alloys (by changing the phase composition and their ratio) are affected by the hardening heat treatment and plastic deformation (as confirmed by studies [162]).
In addition, Ti-Ta alloys have high biocompatibility and corrosion resistance. Thus, the results of a study [111] have confirmed that the corrosion resistance of Ti-Ta alloys with tantalum contents of 30, 40, 50 and 60 wt.% are not inferior to that of the Ti-6Al-7Nb alloy, and in fluorinated acidified saliva, even exceed it, which can be explained by the formation of Ta2O5 oxide film on the surface of a Ti-Ta alloy. At the same time, the authors of another study [112] state that a minimum of 40 wt.% Ta is recommended to achieve excellent corrosion properties in Ti-Ta alloys.
Tantalum also exhibits antibacterial activity against various pathogens such as Staphylococcus aureus and Escherichia coli [163,164], so further studies on the antibacterial activity of Ti-Ta alloys are needed.
As mentioned above, molybdenum acts as a strong β-stabilizer, which reduces the elastic modulus and increases the corrosion resistance of binary Ti-Mo alloys. Thus, Ti-Mo alloys with Mo content of up to 12% have lower elastic moduli than cp-Ti (138.56 GPa) [165]. However, the character of the influence of Mo concentration (in the concentration range from 3.2 to 12 at.%) on the elastic modulus of the alloy is identical to that of tantalum: the first local minimum corresponds to the Ti-3.2Mo alloy (83.8 GPa), the local maximum to Ti-6Mo (112.092 GPa), and the second local minimum to Ti-8Mo (82.98 GPa). It is worth noting that a significant increase in the elastic modulus between Ti-4.5Mo and Ti-6Mo alloys is observed when the Mo content increases up to 6 at.%, then its value falls sharply when the Mo content increases from 6 to 8 at.%, and when the Mo content changes from 8 to 12 at.%, only a slow increase is observed [165]. A further increase in the amount of molybdenum in binary alloys, as well as their additional alloying with zirconium, leads to an increase in the amount of β-phase in the alloy structure, which contributes to an even greater decrease in the elastic modulus while maintaining the increased strength of the alloy. In [123], it was shown that the ratios of α/β-phase (in %) for Ti-12Mo, Ti-15Mo, Ti-12Mo-6Zr, and Ti-15Mo-6Zr alloys were 53.7/46.3, 50.61/49.39, 40.03/59.97, and 16.26/83.74, respectively. This also led to an increase in the hardness of the alloys and the elastic modulus of the alloys (Figure 7).
The introduction of Fe into the ternary alloy of the Ti-Mo-Zr system reduces the elastic modulus of the alloy (from 105 GPa to 93 GPa with the addition of Fe from 1 to 3% [125]), while Ti-Mo-Zr-Fe alloys (containing from 1 to 4% Fe) are pseudo-β-alloys, i.e., the structure of the alloys contains up to 99% of the β-phase, the grain size of which decreases with increasing iron content [125].
Ti-Cu alloys are promising antibacterial biomaterials that have the potential to be tools against peri-implantitis and antibiotic resistance [117]. It was found that by varying the Cu concentration (1 to 10 wt.%) and aging temperature, the antibacterial properties of Ti-Cu alloys could be controlled. Thus, the best antibacterial properties were observed in the alloy with the maximum Cu content, Ti-10Cu, which was aged at 400 °C for 6 h [117], which may have been due to the high content of Ti2Cu in the aged alloy and, as a consequence, a higher rate of release of Cu ions. It is worth noting that this alloy was the only material that showed an antibacterial effect after two hours of testing, while after six hours, bacteria were destroyed in all alloys with a Cu content of more than 5 wt.%.
Ti-Mg alloys have a low modulus of elasticity while maintaining the high specific strength and corrosion resistance inherent in cp-Ti (while the corrosion resistance and compressive modulus of Ti-Mg alloys decrease with increasing amounts of Mg) [126,127,128]. Moreover, Mg can dissolve in the modeled body fluids and contribute to the formation of a calcium phosphate layer; after the degradation of Mg, it is possible to obtain a porous titanium framework that promotes bone sprouting into the implant volume, i.e., the osseointegration and bioactivity of the implant are increased [128].
The introduction of Sr slows down the degradation rate of Ti-Mg alloys and increases their biocompatibility and bioactivity [130]. Sr promotes bone tissue regeneration and osteointegration at the interface (bone–implant), stimulates protein amalgamation, prevents osteoporosis, and suppresses bone resorption.
The alloys of the Ti-Mg-Sr system showed high mechanical properties in comparison with traditional alloys (cp-Ti, Ti-Al-V). Thus, the Ti-10Mg-20Sr alloy showed a low elastic modulus of 36 ± 7 GPa (measured by nanoindentation) with a hardness of 1.8 ± 0.8 GPa [130].
The addition of calcium to both titanium as well as noble metals (in particular, Pd, Pt) leads to an increase in the corrosion resistance of the binary alloys Ti-Ca, Ti-Pt, and Ti-Pd in acidic fluoride aqueous solutions (e.g., HF + NaF) adapted to different pH values (from 4.7 to 3.4) [133]. This was confirmed in a study [134], where it was shown that titanium alloys Ti-0.2Pd and Ti-0.3Mo-0.8Ni exhibited higher corrosion resistance than cp-Ti at fluoride concentrations below 0.002 M due to the accelerating effect of Pd and Ni on the cathodic process and the inhibiting effect of Mo on the anodic process.
In addition, the introduction of calcium and palladium into the composition of a titanium alloy contributes to the improvement of the osteointegration of the alloy. Thus, after 60 days of exposure to immersion in SBF solution with the addition of bovine serum albumin (0.8 g/l), precipitated CaP compounds (with Ca:P ratios from 1:0.7 to 1:4.4) were found on the surfaces of Ti-0.2Pd alloy samples [135].
Thus, alloying titanium alloys with β-stabilizers (Nb, Mo, Zr (for triple and more alloys), Fe, etc.) leads to a significant decrease in the elastic modulus and increase in the strength of alloys in comparison with traditional commercially available alloys (cp-Ti, Ti-6Al-4V), but it is not enough (the minimum elastic modulus of titanium alloys is 40–50 GPa, and that of bone is 4–30 GPa). The further reduction of Young’s modulus is possible by creating porous structures with pore sizes of 100–500 μm for better ingrowth of new bone tissue [127,166]. The prospectivity of this has been supported by some studies. For example, it has been shown [127] that porous Ti fabricated using the high-pressure solid-state sintering (HPSSS) method has a high strength (~115 MPa), relatively low Young’s modulus (~18 GPa), excellent ductility, and suitable pore sizes (50–300 μm), which make it attractive for load-bearing implant applications. Meanwhile, the strength of porous titanium can also be improved by alloying with Mo and Nb. Thus, porous alloy Ti-12.5Mo, with a porosity of 40–45%, and alloy Ti-25Nb, with a porosity of 39–48%, sintered at 1050 °C for 2 h, have excellent properties close to the properties of human cortical bone (Young’s modulus is in the range of 5–18 GPa, with compressive strength in the range of 141–286 MPa), which correspond to the properties of bone [167]. In addition, in the study, the compressive strength and Young’s modulus increased linearly with the decreasing porosity of the samples.
As our study has shown, titanium and its alloys remain some of the promising materials for dental implants with excellent biocompatibility and osseointegration. And, their development follows the way of making the mechanical properties of titanium alloys approach the properties of bone tissue (due to alloying and the creation of meta-materials with porous 3D structures), increasing tribocorrosion properties (in particular, in acidic fluoride media, where titanium has low corrosion resistance), and giving the alloys antibacterial and bone tissue regenerating properties.

2.2. Zirconium and Its Alloys

In modern practice zirconium, and its alloys are used for dental implant manufacturing [168,169,170,171,172,173]. Comparing the values of the electrode potentials of titanium (-1.63 mV) and zirconium (-1.4 mV), it can be assumed that implants based on zirconium alloy are more preferable, which is due to the negative influence of the negative potential of the implantation material surface on the surrounding tissues [174]. The higher the negative value of the standard electrode potential of a metal is, the greater its solubility and reactivity will be [175].
The conducted analysis of the results of the above-mentioned search query has shown that two groups of zirconium-based alloys are used for dental implant manufacturing:
  • Metal alloys of the Zr-Nb [176,177,178], Zr-Ti [153,179,180,181,182], Zr-Ti-Mo [168], Zr-Ti-Nb [183,184], Zr-Ti-Ag [179], Zr-Ta, and Zr-Ta-Ti [9] systems;
  • Ceramic materials based on zirconium dioxide:
    ZrO2 [185,186,187,188,189,190];
    ZrO2 stabilized with cerium (Ce-TZP) [191,192];
    ZrO2 stabilized with yttrium (Y-TZP or YSZ) [193,194,195,196];
    Ce-TZP/Al2O3 nanocomposite, which is a Ce-stabilized ZrO2-based composite material reinforced with Al2O3 crystals or fibers (Al-TZP) [197,198,199,200].
Zirconium is a rather soft gray metal, but it has a lower modulus of elasticity (92 GPa), better corrosion resistance than Ti [201], and good biocompatibility and osteoinductivity [202]. Zr, similar to titanium, is allotropic. Therefore, metal alloys based on it (in particular, Zr-Nb) belong to the group of alloys with solid-solution hardening and differ from intermetallic alloys, i.e., those inclined to magnetization, which include titanium, based on the high characteristics of fatigue endurance that do not depend much on the metal structure [174].
In [178], the microstructure and mechanical properties of cast Zr-(0-24)Nb alloys were investigated. It was found that when Nb was introduced in small amounts (up to 6 wt.%), Zr-Nb alloys consisting mainly of α’-phase (containing less than 6 wt.% Nb) showed the highest strength (786–881 MPa), moderate ductility (6.5–3.7%), and relatively high Young’s modulus (74 GPa) while maintaining low magnetic susceptibility. A further increase in niobium content led to a decrease in alloy strength (tensile strength of Zr-22Nb alloy was 605 MPa), which was associated with both the appearance of ω-phase as well as an increase in the amount of a less strong β-phase. It is worth noting that Young’s modulus of Zr-Nb alloys at a niobium concentration of up to 24% (over the whole investigated range) was lower than the elastic modulus of the titanium alloy Ti-6Al-4V, with the minimum value of Young’s modulus (48.4 GPa) being associated with alloy Zr-20Nb. Although the strength of this alloy was inferior to that of the Ti-6Al-4V alloy (687 MPa vs. 994 MPa), its value exceeded that of bone strength, and the magnetic susceptibility of the Zr-20Nb alloy was half that of the Ti-6Al-4V alloys, making it a preferred implant material for patients requiring MRI studies [178].
As mentioned above, zirconium and titanium have close atomic sizes and allotropic transformation temperatures, which allows these elements to form solid solutions of unlimited solubility. The properties of binary alloys Ti-xZr (where x is up to 90 at.%) have been discussed in detail above. Let us only add that metallic zirconium has excellent corrosion resistance in acid or alkali solutions but is subject to localized corrosion in media containing chloride or fluoride. The introduction of titanium into the alloy composition reduces the local corrosion of Zr-Ti alloys (while increasing the titanium content contributes to a decrease in the corrosion rates of the alloys), but their resistance to pitting corrosion is still inferior to that of cp-Ti [179]. Zr-20Ti and Zr-40Ti alloys exhibit good mechanical properties (the ultimate strength values are 1630 MPa and 1884 Mpa, and the microhardness values are 330 HV and 340 HV, respectively) and good machinability in cold deformation (alloy deformations at fracture are 25.2% and 26.8%, respectively) [179].
Binary Zr-Ti alloys have a predominantly α-crystalline structure, and the introduction of molybdenum, which is a β-stabilizer, leads to a change in the alloy structure. With the gradual addition of Mo from 0 to 12.5 wt.%, the phase composition of the alloys changes as α → α + β → β, with the amount of β-phase increasing with the increasing Mo content [168]. The properties of Zr-Ti-Mo alloys fully meet the requirement of aligning with the mechanical properties of biomedical dental implants. Thus, the alloys Zr-25Ti-xMo (where x is from 0 to 12.5 wt.%) have a reduced compressive strength in the range of 1154.4 to 1310.8 MPa, yield strength of 673.2 to 996.0 MPa, and Young’s modulus of 17.74 to 24.44 GPa, and the alloy Zr-25Ti-7.5Mo shows the best wear and corrosion resistance [168].
The introduction of Nb into the composition of Zr-Ti alloys also leads to the change of its structure to the β-type. Thus, the microstructure of the cast Zr-19Ti-21,4Nb alloy is represented by β-grains of 150–200 μm in size. The mechanical properties of this alloy in the deformed state correspond to the properties of the Ti-6Al-4V alloy (the tensile strength of the Zr-19Ti-21,4Nb alloy was 850–900 MPa, and the deformation of the alloy at fracture, characterizing plasticity, was 10.5%), but the modulus of elasticity is much closer to that of natural bone (37.6 GPa) [183]. The experiment with cell cultivation demonstrated the adequate adhesion of osteoblasts on the first day to both sr-Ti and Zr-Ti-Nb alloys, with subsequent proliferation on days 3 and 7. These data indicate the biocompatibility of Zr-19Ti-21,4Nb alloy and the absence of toxicity to cells [183].
The introduction of silver (6%) into Zr-Ti alloys leads to increased resistance to electrochemical corrosion (in particular, in artificial saliva containing fluorine). Moreover, the effect of Ag is stronger the more titanium is in the alloy composition since the joint action of titanium and silver promotes the formation of a thick, compact, and stable passive film [179]. At the same time, silver slightly changes the mechanical properties of Zr-xTi alloys. Thus, in a study, the introduction of 6% Ag increased the hardness of the Zr-20Ti alloy from HV 329 to HV 338, and the hardness of the Zr-40Ti alloy increased from HV 340 to HV 349 [179].
The Zr-30Ta and Zr-25Ta-5Ti alloys also showed improved biocompatibility and osteogenic activity in comparison with cp-Ti due to, among other things, favorable surface properties (increased surface hydrophilicity and roughness) [9]. Thus, the Zr-25Ta-5Ti alloy had the smallest contact angle (~30.6°) and the largest roughness value (Ra = 34.8 nm) while, for the Zr-30Ta alloy, the values of the contact angle of wettability and roughness were 46.6° and 20.7 nm, respectively (for comparison, these characteristics for cp-Ti were 52° and 13.1 nm). The differences between the samples were statistically significant (p < 0.01). In addition, the Zr-30Ta and Zr-25Ta-5Ti alloys also induced the balanced expression of M1 and M2 macrophage phenotypes during the first 24 h. A Zr-25Ta-5Ti sample adsorbed three times more protein ((147.4± 13.26) μg) compared to Zr-30Ta ((32.1 ± 4.57) μg) and CP-Ti ((32.4 ± 3.7) μg) samples.
Unlike metallic zirconia, ZrO2, zirconium dioxide, is an ultra-hard material; at the same time, it has a white color similar to that of the natural tooth root, which is favorable for restorations in the smile area [186], and it is biocompatible, does not provoke allergies, and is one of the most durable in service [174]. The zirconia surface is poorly adhered to by bacteria (40% less adhesion than metals) [203,204], and ceramic materials are less prone to plaque accumulation than metal substrates [200].
Zirconium dioxide ceramics have higher strength (the bending strength of zirconium ceramics varies from 500 to 1200 MPa [205,206,207]) and crack resistance (not less than 15 MPa-m1/2 [206]); high acid-, corrosion-, and wear resistance due to high hardness (Vickers hardness of not less than 10 GPa [206]) and heat resistance; and biocompatibility in comparison with alumina analogs [174,208]. This material is characterized by an extremely low coefficient of friction with metals and the possibility of obtaining very high surface cleanliness [174]. However, compared to titanium alloys, zirconium dioxide has an increased modulus of elasticity [208] (100–200 GPa [206] vs. 30–115 GPa [75,76,82,90,94,209]), comparable to the modulus of elasticity of stainless steel (210 GPa [209]). In addition, ZrO2 has a high value of compressive strength (more than 2000 MPa [210]), significantly exceeding the compressive strength of titanium alloys (~400–1000 MPa [93]). It should be noted that the strength of ceramics depends on porosity and grain size, which, in turn, are influenced by the sintering technology and the addition of various modifiers to the ceramic composition, and the strength of titanium alloys depends on alloying additives and alloy manufacturing technology. Thus, [25] has shown that the compressive strengths of ZrO2 ceramics and Ti-6Al-4V alloy are practically equal (1525 MPa and 1565 MPa, respectively).
The introduction of stabilizing oxides of ceramics cerium (CeO2), yttrium (Y2O3), and aluminum (Al2O3), which form a solid solution with ZrO2, prevents spontaneous tetragonal–monoclinic transformation from occurring during the cooling of ceramics from the sintering temperature and contributes to grain refinement and leads to an increase in the bending strength of ceramics and ductility (it is known that when the grain size is reduced to values of the order of 1–5 microns, there is a decrease in porosity and an increase in strength and the amount of tetragonal zirconium dioxide [206]), which make it possible to withstand high torsional moments during implant placement. At the same time, the high compressive strength values of the ceramics are maintained.
Zirconium dioxide is also bioinert to other materials in the oral cavity and is particularly suitable for patients who are allergic or intolerant to metals [211]. The clinical use of zirconium oxide implants is practically free of peri-implantitis, and gingival epithelial tissues are able to firmly and reliably attach to their surfaces [212].
It was found that the osseointegration processes of implants made of ZrO2 and Ti are similar. Thus, the degree of contact of the implant with the surrounding bone after 4 weeks of functional loading (p = 0.505) for ZrO2 was 76.76%, and it was 77.25% for titanium, and the average total bone density after 4 weeks was 57.51% around zirconium implants and 60.89% around titanium implants (p = 0.223) [187]. A greater number of formed and shaped bone areas were found around the perimeter of bone contact areas with zirconium implants compared to bone contact areas with titanium implants, where erosion lacunae were more numerous [213].
It was also found that the stabilization of ZrO2 with cerium oxide leads to the increased bioactivity of ceramics (Ce-TZP nanocrystals directly bind to hydroxyapatite crystals deposited by osteoblastic cells without any preliminary chemical treatment of the substrate surface) [191]. Ce-TZP has extremely high values of fracture toughness (up to 30 MPa-m1/2) [206], but it is possible to increase the plasticity of Ce-TZP zirconia ceramics only by introducing no more than 10 mol% [192]. However, from an aesthetic point of view, Ce-TZP ceramics have a yellowish color, which limits the material’s use as a restoration for frontal teeth [214].
ZrO2 stabilized by yttrium (yttrium oxide) also has high values in terms of strength (up to 1300 MPa), Young’s modulus (~ 205 GPa), and hardness (10 GPa) [206]. At the same time, an increase in the toughness of ceramics is possible at a content of more than 3 mol% Y2O3. The results of a study on Y-TZP ceramics biocompatibility testify to the increased adhesion, viability, and proliferation of osteoblasts and gingival fibroblasts on zirconium surfaces in comparison with titanium ones, which the authors [195] attributed to the increased wettability of the zirconium implant surface. However, one of the disadvantages of Y-TZP is degradation during low-temperature aging; exposure to moisture for a long period of time, even at low temperatures in the mouth, can lead to material cracking and catastrophic implant failure [200,215].
Ce-TZP/Al2O3 nanocomposite is a Ce-stabilized ZrO2-based composite material reinforced with Al2O3 crystals of several hundred nanometers in size. Ce-TZP/Al2O3, containing about 30% aluminum oxide particles, exhibits high fracture toughness (19 MPa-m1/2) and extremely high bending strength (1400 MPa) [199]. In addition, the increased resistance to low-temperature fracture suggests that the material will have long-term reliability. A histologic analysis of cross-sections of implanted Ce-TZP/Al2O3 ceramic fixators showed not only perfect biocompatibility but also a high rate of bone integration [200]. For example, the bone-to-implant contact (BIC) rate after eight weeks was 80%, which was comparable to the BIC rates with cp-Ti (85%) and Ti-Zr (72%) implants and much higher than those for ZrO2 (22–48% [216]) and Ti-6Al-4V (30%) implants. In addition, Ce-TZP/Al2O3 implants adhered tightly to the surrounding soft tissues with no signs of inflammation [200].
The wide implementation of ceramic implants is hindered by the relatively low machinability of ZrO2 [217,218], increased brittleness of ceramics [219,220], low thermal conductivity and sensitivity to thermal shock [206,221,222,223,224,225], and CAD/CAM technologies of mechanical processing of products, which have led to the increase in the volume of studies, the object of which are implants from ceramic materials (for the last 20 years, the number of publications with studies on zirconium implant properties has increased 2.5 times; see Figure 6), with the market for dental implants made of zirconium dioxide increasing by more than 12% per year [226].

2.3. Tantalum and Its Alloys

Tantalum, like titanium, is a strong and bioinert metal. However, it is used much less frequently because it is a rare-earth metal (titanium is found in the Earth’s crust almost 2500 times more often than tantalum [227,228]). Moreover, tantalum is less strong but more ductile (1.4 times) and refractory (1.8 times) than titanium. At the same time, the complexity of the production of metallic tantalum from ore containing hundredths of a percent of (Ta,Nb)2O5 and then the use of expensive processes for manufacturing tantalum products (vacuum-arc, plasma melting, or powder metallurgy methods) [228] eventually sharply increase the cost of the material.
Compared to titanium, Ta has higher thermal and electrical conductivity [227,228] and high corrosion resistance [229] and has been shown to have better biocompatibility (the Ta surface promotes the better adhesion, proliferation, and differentiation of osteoblasts compared to the titanium alloy Ti-6Al-4V) [230,231]. In addition, tantalum (even coated on a titanium substrate) shows good antibacterial activity against Streptococcus mutans and Porphyromonas gingivalis [232].
Basically, tantalum implants are made of highly porous “pure” Ta, the so-called Trabecular Metal™ (TM) [30,233,234,235,236]. Its open-cell structure has interconnected pores, as a result of which the structure becomes highly porous (75–80%) and resistant to fatigue failure and retains its strength throughout the healing process [237]. Such tantalum implants are capable of osteointegration and biological fixation with the growth of new bone tissue in the pores and trabeculae (small cells or labyrinths, which resemble the own trabeculae of the bone substance) of the implants, and no fibrotic changes are detected at the “bone–implant” boundary [238].
The modulus of elasticity of porous Ta (1.3–10 GPa [229]) is similar to the modulus of elasticity of subchondral bone [237,239] and more similar to that of natural cortical (12–18 GPa [238]) and spongy bone (0.1–0.5 GPa [229]) in comparison with traditional cp-Ti and titanium alloys (106–115 GPa [94,229]), which contributes to stress reduction and the preservation of bone stock. Simultaneously, the high coefficient of friction allows TM to demonstrate higher biomechanical stability compared to conventional cementless implants even in conditions of bone deficiency or insufficient bone density [240].
However, the creation of a material with a highly porous structure leads to a decrease in its mechanical properties [241], which can limit its use in load-bearing structures, while the decrease in the strength of the implant material can impair its mechanical stability in vivo and increase the frequency of implant failures [108]. Thus, the compressive strength of highly porous Ta is 60 MPa (which is significantly higher than the values of compressive strength, though, compared to porous (75%) titanium (~25 MPa) [242]), the tensile strength is 63 MPa, and the bending strength is 110 MPa; in a study, the compressive fatigue endurance limit was 23 MPa at 5–106 cycles, and the samples showed significant plastic deformation [243]. On the other hand, tantalum implants with low porosity may not provide the necessary surface area for cell attachment and tissue ingrowth [219]. Also, the efficiency of bone tissue “sprouting” into the pores of tantalum implants is influenced by the pore size. A histological analysis of highly porous tantalum specimens showed a significant increase in the degree of filling of the implant pores with bone tissue with time. The authors [237] noted that implants with a smaller pore diameter (430 microns) have better mechanical properties in comparison with more “porous” (650 microns) implants. The optimal pore size for promoting bone ingrowth in metallic porous materials, as previously mentioned, is 100–500 μm [127,219], providing a large surface area for cell attachment and proliferation, which can promote both new bone formation as well as the diffusion of nutrients and waste products, which is important for promoting cell survival and proliferation [219].
However, tantalum has a number of disadvantages, such as the high cost of the material, and there are difficulties in manufacturing the precise internal relief of the implants, which also increases the cost of tantalum implants, so they are not available to all patients. As a consequence, the dental implant can be made in a hybrid design (titanium alloy anchoring base with a trabecular tantalum middle part) [30,234,244,245] and also form tantalum coatings on implants made of other materials [126,246,247,248,249,250].
In addition, search studies are underway to develop biocompatible alloys of increased strength based on tantalum, in particular Ta-Ti [108,162] or bioactive Ta-Cu alloys [251].
The influence of tantalum content on the properties of binary Ti-xTa alloys (where x = 10–80 wt.%) has been discussed in detail above. Let us only note that the Young’s modulus and tensile strength values of Ta-Ti alloys with tantalum contents of 50 to 80% are in the ranges of 67–102 GPa and 530–685 MPa, respectively [162] (for comparison, the tensile strength of pure compact Ta is 206 MPa [228]). At the same time, the maximum strength has been shown by the Ti-60Ta alloy, and the minimum elastic modulus has been shown by the Ti-70Ta alloy.
The addition of copper to tantalum leads to an increase in its antibacterial properties but can cause minor cytotoxicity and reduce the corrosion resistance of Ta [251]. Thus, the sintered Ta-5Cu alloy shows increased antibacterial activity against Escherichia coli (E. coli) due to the prolonged release of Cu ions. At the same time, tribological tests in pin-on-disk sliding geometry on stainless steel (pin-on-disk wear tests) have revealed a much lower coefficient of friction but higher wear rate for the Ta-5Cu alloy compared to pure tantalum [251]. This can be explained by the fact that softer and more ductile copper acts as a “hard lubricant”. The increased wear rate of the Ta-5Cu alloy is also consistent with the data on the decrease in its hardness compared to Ta (3.6 GPa vs. 4 GPa) [251].
Thus, tantalum and its alloys are promising materials for application in dental implantology, but additional research is required for highly porous structures of alloyed tantalum alloys, which give an opportunity to combine high strength with a low modulus of elasticity.

2.4. Other Materials for Dental Implants

A number of other materials are also known that are capable of forming a strong bond with bone tissue.
As previously stated, some of the first materials used for implant fabrication were stainless steels, particularly the austenitic steels AISI 316 and 316L [252,253,254] and the ferritic steels AISI 444 and NeoM (the steel of commercially marketed dental implants) [255]. However, the most recent studies where steel implants were the subject date from the 2010s; in the 2020s, steel implants have only been mentioned in review articles or used for control groups (for comparison). Although stainless steels are a stronger and mechanically reliable class of metal alloys, their use as biomaterials is limited due to their lack of proper biocompatibility and tendency to corrode over time in biological environments [219].
As a result, modern dental implant manufacturers have abandoned steel implants because steels have the highest modulus of elasticity (~210 GPa [209]), the lowest corrosion resistance in biological media [254], and the lowest biocompatibility [219,253]. Thus, comparative in vitro tests of the corrosion resistance of commercially available mini-implants made of stainless steel AISI 316 and titanium alloys Ti-6Al-4V (Grade 5 and Grade 23 with reduced oxygen content), when immersed in artificial saliva, saliva with probiotic bacteria Lactobacillus reuteri, and saliva with an oral antiseptic—chlorhexidine (CHX)—revealed an increase in the roughness of steel implants when exposed to CHX, as well as a decrease in their microhardness compared to unexposed implants when exposed to CHX and probiotics [254]. It is worth noting that exposure to probiotics significantly increased the roughness of class 5 titanium implants compared to other media, but the microhardness of the samples did not change after exposure to either medium. At the same time, the corrosion resistance of austenitic steels can be increased by applying appropriate heat treatment (by correctly selecting the temperatures of the hardening and subsequent aging of the steel) [252].
In addition, stainless steel AISI 316L demonstrates low integration in contact with the surrounding bone tissue, and the duration of implant healing can reach several months [253].
The osseointegration of steel implants has been attempted to be improved by fabricating products using additive manufacturing techniques for dental implants without the use of subsequent surface treatment [256]. The samples in one study had higher roughness and lower surface energy than commercially realized implants.
The ferromagnetic properties of ferritic stainless steels make them potential materials for use as magnetically bonded implantable dental devices [255]. Comparative studies on the corrosion resistance and cytotoxicity of ferritic stainless steel AISI 444 (ferritic steel alloyed with Nb and Ti) have been carried out, and NeoM steel (ferritic steel for commercialized dental implants) and austenitic steel (the composition of which corresponds to ISO 5832-1, [257]) have revealed an increased tendency towards pitting corrosion in phosphate buffer solution (PBS). The resistance to pitting corrosion of AISI 444 steel is comparable to that of austenitic steel [255].
Taking into account the biocompatibility and high chemical resistance of ceramic materials, there have been studies suggesting Al2O3- [258,259] and Si3N4 [24,260,261]-based ceramics as promising biomedical materials for the production of dental implants. Table 2 presents the mechanical properties of ceramic materials used for the production of dental implants in comparison with traditional titanium alloys.
Aluminum oxide (Al2O3) has high hardness and wear resistance, hydrophilicity, corrosion resistance, good thermal conductivity, and a low coefficient of friction, but has low bending strength [274]. It has been shown [258] that the surface wettability of Al2O3 and Y-TZP samples is within the hydrophilic regions (the contact angle for Al2O3 was ~64°, and for Y-TZP, it was ~58°; for comparison, Ti-6Al-4V alloy has a high wetting angle of ~85° [262]). This accounts for the better cellular adhesion of ceramic materials (approximately similar for both types of ceramics) compared to titanium alloy. Meanwhile, the cell adhesion of human osteoblasts (HOB), human osteoblast-like cells (MG-63), and human mesenchymal stromal cells (hMSC), as well as cell spreading and the number of focal cell contact points, were further increased via the covalent immobilization of alkaline phosphatase (ALP), an enzyme involved in bone mineralization, to the surfaces of ceramics [258].
In recent years, silicon nitride (Si3N4) has attracted increasing attention. Nitride ceramics have increased hardness and strength compared to other types of ceramics (see Table 2), also showing excellent biocompatibility [24,260,261]. It was found that when Si3N4 samples were immersed in SBF solution, the pH and ionic conductivity values of the samples varied widely in the first days, but then stabilized around SBF values (pH = 7.26; C = 9.148 mS/sm) afterwards (5–6 days). The exposure of silicon nitride ceramics to the MG-63 cell line for 24 h results in low lactate dehydrogenase (LDH) activity and, therefore, a good percentage of cell viability. In addition, the cells grow and proliferate near the samples under normal conditions [261].
Also, nitride ceramics showed good osseointegration in a study. For example, three months after surgery, histologic sections showed superior new bone formation around the tested Si3N4, polyether-ether-ether-ketone (PEEK), and cp-Ti implants compared to Ti and PEEK (69%, 24%, and 36%, respectively) [275].
At the same time, the strength of Si3N4 depends on porosity, which, in turn, depends on the sample fabrication technology—in particular, on the sintering temperature (the sintering pressure and holding time had almost no effect on the strength of the product) [261].
In addition, Si3N4 exhibits antibacterial properties, which proves a threefold increase in the biofilm inhibition area on Si3N4 samples compared to cp-Ti samples under the same conditions [261]. This gives an advantage in using nitride ceramics for dental implants over titanium and zirconium alloys—a decisive advantage over available Ti or Zi dental implants, as the antibacterial properties may help cope with the continuous increase in the prevalence of peri-implantitis [260].
As biomaterials for dental implant manufacturing, polymeric materials have also been tried; in particular, polyether-ether-ketone (PEEK) [23,26,195,262,275,276,277,278,279,280] is an X-ray-negative semi-crystalline thermoplastic that combines strength (100–230 MPa depending on the polymer base), hardness, and wear resistance, as well as high chemical resistance and biocompatibility [26]. In addition, unlike metallic and, even more so, ceramic materials, PEEK has an elasticity modulus close to the values of the bone tissue elasticity modulus, and it can be regulated by adding carbon particles or fibers to the PEEK composition (which also leads to an increase in polymer strength) [26,277]. Thus, when short carbon fibers are added, the elastic modulus of PEEK is 4–18 GPa [26].
At the same time, in a study, PEEK samples had better adhesion, viability, and proliferation of osteoblasts and gingival fibroblasts compared to titanium, and had indicators similar to those to ZrO2 samples, which correlated with the increased wettability of these materials [195], but in comparison with titanium implants, PEEK is less osteoconductive than titanium (less bone–implant contact is observed than with titanium implants) [278]. PEEK also affects the biofilm structure and reduces the likelihood of inflammation around the implant [23].
Moreover, the osteointegration of PEEK implants can be improved by modifying treatment, increasing the surface hydrophilicity [23,262,276,279,280], or by forming bioactive coatings on the surface, which, according to the authors of [23], enhances cell adhesion, proliferation, biocompatibility, and the osteogenic properties of the polymer. Thus, in a study, sandblasting treatment with Al2O3 microparticles (110 μm in size) of PEEK implants increased bone ingrowth and the degree of implant contact with bone tissue (51.1% for treated PEEK, 30.9% for untreated PEEK), and the osteointegration of treated PEEK was similar to the osteointegration of the titanium implant, the degree of implant contact with bone tissue of which was 54.0% [276,280]. In another study [279], it was shown that laser-modified titanium and PEEK surfaces led to the directed attachment of gingival fibroblasts.
It has also been shown that the shear strengths at the interface between the bone and implants made of carbon fiber-reinforced PEEK (CFR/PEEK) and cp-Ti with hydroxyapatite (HA) coatings were comparable [277], which can be explained by the fact that in both cases, the same materials (bone and hydroxyapatite) come into contact, and the adhesion of HA with implant materials (titanium and PEEK) is the same [277]. At the same time, HA-coated specimens have better osseointegration in comparison with uncoated specimens (in a study, the shear strength at the interface between bone and implants was 8 MPa for uncoated specimens and 15 MPa for coated specimens) [277].
However, there have been studies that do not recommend the use of unmodified PEEK (without surface modification) for implants [278,281]. PEEK compared to Ti and ZrO2 showed the highest values of total strain, demonstrating reduced Mises stresses in implants and abutments, but had high tensile stresses in trabecular bone, reaching the critical values [276]. Additional comparative animal and clinical studies are needed to determine the potential of PEEK as a biomaterial for dental implants.
In addition, there have been studies in which, for the production of dental implants, the prospect of composite material application has been considered, i.e., materials consisting of two or more heterogeneous materials (with essentially different physical and/or chemical properties), with a clear interface between them, combining the properties inherent in several constituent materials—for example, metal–ceramic materials, metal–polymer materials, etc.
In [259], the microstructural and mechanical properties and biocompatibility of the composites “hydroxyapatite–ceramic” HA-Al2O3 and HA-ZrO2, with the addition of 5 and 10 wt.% of commercial inert glass (CIG; contains ~69% SiO2, ~17% Na2O, ~9% CaO, and ~2% Al2O3 and MgO and other oxides [259]), were investigated. With an increasing CIG content, the compressive strength and microhardness of HA-Al2O3 composites increase while the microhardness of HA-ZrO2 composites also increases and the compressive strength decreases. The biocompatibility (in vitro and in vivo) and mechanical properties of HA-Al2O3 composites (microhardness of HV 43 and compressive strength of 36 MPa at sintering temperature of 1200 °C and 10% CIG content) are lower than those of HA-ZrO2 composites (HV 47 and 53 MPa under the same conditions).
One of the most important properties to be considered in the fabrication of all-ceramic dental implants is cyclic fatigue because implants are subjected during mastication processes: precisely, during cyclic loading, which can lead to fatigue cracking (appearance of subcritical cracks) and sudden implant failure (implant fracture). It has been shown [282] that ZrO2-Al2O3 (80:20) composite can have up to 20 years of service if the implant diameter is chosen correctly with regard to cyclic fatigue. At the same time, the hardness of the ZrO2-Al2O3 composite is HV 15.2 GPa, and the tensile strength is 700 MPa, which is twice as high as the strength of some types of ceramics and comparable to the tensile strength of titanium alloys (see Table 2).
To improve tribocorrosion properties, a composite material, (Ti-6Al-4V)-PEEK, was created [283]. Ti-6Al-4V alloy is characterized, as noted above, by high corrosion properties due to the formation of a passive TiO2 film on the surface, but, as a result of tribochemical reactions occurring at the interface of the “implant-bone”, it can be destroyed (under the action of both wear and corrosion), which becomes the reason of metal ions released in a toxic concentration (metallosis phenomenon) [283]. In one study, the impregnation of a Ti-6Al-4V alloy with a lattice structure obtained using selective laser melting with polyether-ether-ether-ketone (PEEK) at hot pressing allowed to increase the wear and corrosion resistance of composite (Ti-6Al-4V)-PEEK in comparison with samples from titanium alloys, both dense and cellular structured (the specific wear rate of the composite on tribological tests in phosphate buffer solution decreased by 450% in comparison with the sample from Ti-6Al-4V, cut from a rolled bar, i.e., obtained using the traditional method) [283].
To avoid problems related to both the mobility of the components of a prefabricated dental implant (between the implant and abutment and the abutment and crown) as well as bacterial colonization in the gaps of the mechanical connections, thus preventing the fusion of the gingiva or bone tissue with the implant, which can lead to dental implant rejection [284], a one-piece implant consisting of two layers was developed: a zirconia ceramic was used for the crown and a titanium alloy was used for the implant body. The two-layer monolithic metal–ceramic dental implant (Ti-6Al-4V)-YSZ, fabricated using spark-plasma sintering, showed good mechanical properties and biological properties investigated in vitro (good biocompatibility, non-cytotoxicity, cell adhesion, and hemocompatibility) [25].
Thus, there is a great variety of biomaterials that are promising for use in dental implantology. There is a large amount of clinical and statistical material that confirms the well-known thesis that medicine is the science of alternatives—in this case, the alternative consists of using different materials in different clinical situations [174]. However, unfortunately, to date, there has been no clearly structured methodology for the selection of an implant material for specific conditions of use (specific clinical case). Apart from one recommendation, a comparison of the aesthetic performance and durability of implants made of titanium and its alloys (as the main materials of dental implants) with their ceramic counterparts indicates that for implants of load-bearing structures, more durable metal alloys based on titanium and/or zirconium are preferred for masticatory teeth, and in the smile area, more aesthetic white ceramic materials are preferred (based on stabilized zirconium dioxide or silicon nitride (to a lesser extent, it has a gray shade)); however, it is necessary to take into account the following factors [285].

3. Implant Surface Modification

The positive course of the processes of dental implant integration into the bone is characterized by at least three indisputable criteria [286]:
(1)
An absence of rejection reactions expressed in the development of inflammation in the adjacent tissues, local necrotic changes, and systemic manifestations such as allergic and immune reactions;
(2)
The formation of morphofunctional determinants of the integration process in the area of contact, i.e., the “implant–tissue medium”: bone or bone-like substance (in the case of osteointegration);
(3)
The relative stability (including mechanical stability) of the above-mentioned morphofunctional determinants for a certain period of time as a reflection of the dynamic equilibrium occurring in the system, i.e., the “implant–tissue substrate”.
Thus, the main operational properties of dental implants, such as osteointegration and biocompatibility, depend primarily on the properties of their surface layer, as it is the implant surface that makes contact with biological tissues of the body (bone, gingiva).
In this connection, the surface modification of dental implants plays an important role, and the majority of scientific studies are devoted to it. Thus, all methods of surface modification can be divided conditionally into groups (Table 3): mechanical, physical, chemical, biochemical (according to the main influence on the surface), and combined (using several types of processing or influences).
According to the influence of the implants’ surface modification on their functional properties, all the methods can be divided into two large groups:
(1)
Methods aimed at changing the surface roughness of implants to improve their integration;
(2)
Methods forming protective and/or bioactive coatings on the implants to improve their corrosion resistance, biocompatibility, biomechanical stability, and antibacterial properties and to promote bone tissue regeneration.
The requirements of biomechanical compatibility and the fixation of the implant in the body tissues can be satisfactorily solved if we use materials with rough surfaces to which the living tissues are able to firmly attach. In this case, two ways of connection between the implant and the living tissue are created: mechanical coupling as a result of the tissue formation (sprouting) in the implant pores and chemical coupling due to the interaction of the tissue with the components of the elemental composition of the implant. In a study [372], it was shown that due to the rough surfaces of ADiN Touareg implants, the surrounding tissues firmly attach to them and grow into the pores, and when unscrewing such products, there is a significant traumatization of tissues around the implantation site. Smooth polished 3S implants have a smaller area of contact with body tissues. They are fixed using bicortical implantation, are easy to unscrew, and do not tear the surrounding tissues during removal. Thus, on the removed ADiN Touareg implants, there were many tissue fragments, and on the 3S products with smooth surfaces, there were practically no tissue fragments. This indicates the better osseointegration of the rough implant. At the same time, there were no significant histological differences in the bone tissue after the implants with rough and polished surfaces were added, and no significant differences between the states of the surrounding tissues 2 and 6 months after the implantation of each product.
The most frequently used methods for changing the implant surface microgeometry are anodic oxidation, acid/alkaline etching, a combination of these methods, hydrogen peroxide treatment, the formation of coatings by using different methods (the sol-gel method, chemical deposition, etc.), and the mechanical treatment of the implant surface (sandblasting, machine dashing, and laser ablation) (see Table 3).
Another important factor affecting osseointegration is hydrophilicity/hydrophobicity. Thus, hydrophilic materials with a surface tension higher than 30 dyne/cm interact more closely with biological fluids, cells, and tissues and, consequently, contribute to a better osseointegration process [350,377]. The roughness of the implant surface also plays a certain role in this process. Fibroblasts and epithelial cells adhere more strongly to smooth surfaces, and the ability to ensure osteoblast proliferation and collagen synthesis is more pronounced on a surface with moderate roughness [377].
Recently, surface modification via applying various functional coatings has become widespread. This is mainly aimed at improving the integration of implants (foreign bodies) with the body tissues and fighting against peri-implant infection, which can lead to the rejection of artificial products.
Thus, several new technologies of applying hydroxyapatite (Ca10(PO4)6(OH)2), calcium phosphates (CaP), and tricalcium phosphates (Ca3(PO4)2) on the implant surface have been developed [7,277,378,379,380,381,382,383]. It has been proved that CaP compounds promote the formation of direct connections with bone tissue in comparison with implants without the additional coating of the titanium surface.
A separate group consists of bioactive surface modifications of dental implants [384,385,386,387]. Mainly, various osteoinductive growth factors are used [11,375,388]. Moreover, to decrease the risk of infectious complication development, the application of antibacterial preparations on the implant surface [20,389,390] and introduction of metal particles (antibacterial components such as Cu, Zn, Ag, and Ga) into the implant surface or into the coating composition are widely used [16,17,18,19,117,149,163,164,232,251].
The analysis of a large number of controversial results of studies devoted to the contact processes at the “implant–bone” interface allows us to conclude that there is no unambiguous opinion about the “correct” way of modifying the surface of dental implants. The problems of improving the quality of dental implantation and combating the complications of this procedure cannot be solved only by applying other substances on the surfaces of implanted materials. It is necessary to solve the multifactorial problem of increasing the efficiency of dental implantology by choosing the optimal implant material, its manufacturing technology, and the method of surface modification, including the formation of a branched “porous” structure on it and giving it bioactive properties.

4. Conclusions

Based on the results of this review, the following conclusions were drawn:
  • There is a great variety of biomaterials that are promising for application in dental implantology; however, unfortunately, to date, there has been no clearly structured methodology for implant material selection for specific operating conditions (specific clinical cases) or rational technology of implant manufacturing from the selected material;
  • There is a need for a complex approach to improve the quality of dental implants, including the choice of the optimal material, implant manufacturing technology, and the method of its surface modification;
  • The modification of the dental implant surface should combine different methods aimed at creating the surface texture and formation of bioactive coatings;
  • A comparison of the aesthetic indicators and durability of metal and ceramic implants indicates the following for implants of bearing structures: for masticatory teeth, more durable metal alloys based on titanium and/or zirconium are preferable, and in the smile zone, more aesthetic white ceramic materials are preferable (based on stabilized zirconium dioxide or silicon nitride (to a lesser extent, it has a gray shade));
  • Among metal alloys, the most promising are alloys of the Ti-Nb-Zr system alloyed additionally with Ga and/or Sn, but the elasticity modulus of these alloys (dense structure) exceeds the elasticity modulus of bone tissue; to reduce the elasticity modulus and increase the bioactivity and corrosion resistance of metal alloys, it is promising to develop composite meta-materials from these alloys (i.e., manufacturing of alloys of the specified composition of the lattice structure using additive technologies, with further “impregnation” of them with polymeric bioactive materials);
  • A promising direction is the creation of biomedical nanomaterials; nanotextured surfaces have a positive effect on bone tissue cells and have an antibacterial effect. Promising in the production of dental implants is the use of nanostructured titanium, the advantage of which is the absence of toxic elements (aluminum and vanadium) and the higher strength and corrosion resistance inherent in unalloyed titanium;
  • Due to the development of additive technologies, it has become possible to manufacture products with a “controlled” gradient of properties by volume; a one-piece implant consisting of two layers has been developed: zirconium ceramic, used for the crown, and titanium alloy, used for the body of the implant. This has avoided the problems of both the mobility of the components of a prefabricated dental implant (between the implant and abutment and the abutment and crown) as well as the colonization of bacteria in the gaps of mechanical connections, thus preventing the fusion of the gingiva or bone tissue with the implant, which can lead to the rejection of the dental implant; however, it should be taken into account that the heterogeneity of implant and abutment materials may increase corrosion processes due to galvanic processes, so further studies both in vitro and in vivo are required.
Thus, further clinical and experimental studies are needed to select implant materials, processing methods, and surface modification.
To date, there has been insufficient information on dental materials that have optimal mechanical properties, i.e., high strength and low modulus of elasticity (equal to the modulus of elasticity of bone tissue). According to the existing data, varying and/or modifying the chemical composition of materials has not yet allowed the simultaneous achievement of these parameters. Therefore, one of the promising directions of research in this field can be comparative studies of dental implants made of the same alloy but using different methods (traditional and additive technology methods). These studies will provide new insights into the influence of material architecture (macrostructure, porosity, meta-structure) on its mechanical properties (compressive strength, modulus of elasticity). It is also worthwhile to continue research in the field of composite metal–polymer meta-materials or partially biodegradable materials that initially have, or form in the process of use, “macro-pores” necessary for the better osteointegration of implants.

Author Contributions

Draft conception, C.S., S.G., O.K. and N.P.; writing—original draft preparation, C.S. and E.P.; writing—review and editing, C.S., N.U., M.N. and P.P.; supervision, O.Y., N.K. and V.E.; project administration, O.K. and M.N.; funding acquisition, O.Y., N.K. and V.E. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Health of the Russian Federation under project 056-00041-23-00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The study used the equipment from the Centre for Collective Use of Moscow State Technological University STANKIN (agreement No. 075-15-2021-695, 26/07/2021).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Global Oral Health Status Report: Towards Universal Health Coverage for Oral Health by 2030. Available online: https://www.who.int/publications/i/item/9789240061484 (accessed on 15 August 2023).
  2. Follow-Up to the Political Declaration of the Third High-Level Meeting of the General Assembly on the Prevention and Control of Noncommunicable Diseases. Available online: https://apps.who.int/gb/ebwha/pdf_files/WHA75/A75_10Add1-ru.pdf (accessed on 15 August 2023). (In Russian).
  3. Russian Nanotitanium for Bioimplants Has No Analogues in the World. Available online: https://new.ras.ru/activities/news/rossiyskiy-nanotitan-dlya-bioimplantatov-ne-imeet-analogov-v-mire (accessed on 15 August 2023). (In Russian).
  4. Global Dental Implants Market by Design, Type, Price, Procedure, Material, Component, End User—Cumulative Impact of COVID-19, Russia Ukraine Conflict, and High Inflation—Forecast 2023–2030. Available online: https://www.researchandmarkets.com/report/dental-implant#reld0-4620505 (accessed on 15 August 2023).
  5. Dental Implants Market—Growth, Trends, COVID-19 Impact and Forecasts (2023–2028). Available online: https://www.mordorintelligence.com/ru/industry-reports/dental-implants-market (accessed on 24 August 2023).
  6. Thalji, G.; Cooper, L.F. Molecular assessment of osseointegration in vitro: A review of current literature. Int. J. Oral Maxillofac. Implant. 2014, 29, 171–199. [Google Scholar] [CrossRef]
  7. Yuan, K.; Chan, Y.-J.; Kung, K.-C.; Lee, T.-M. Comparison of Osseointegration on Various Implant Surfaces After Bacterial Contamination and Cleaning: A Rabbit Study. Int. J. Oral Maxillofac. Implant. 2014, 29, 32–40. [Google Scholar] [CrossRef]
  8. Yao, L.; Al-Bishari, A.M.; Shen, J.; Wang, Z.; Liu, T.; Sheng, L.; Wu, G.; Lu, L.; Xu, L.; Liu, J. Osseointegration and anti-infection of dental implant under osteoporotic conditions promoted by gallium oxide nano-layer coated titanium dioxide nanotube arrays. Ceram. Int. 2023, 49, 22961–22969. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Yang, H.; Juaim, A.N.; Chen, X.; Lu, C.; Zou, L.; Wang, Y.; Zhou, X. Biocompatibility and osteogenic activity of Zr–30Ta and Zr–25Ta–5Ti sintered alloys for dental and orthopedic implants. Trans. Nonferrous Met. Soc. China 2023, 33, 851–864. [Google Scholar] [CrossRef]
  10. Wang, S.; Zhao, X.; Hsu, Y.; He, Y.; Wang, F.; Yang, F.; Yan, F.; Xia, D.; Liu, Y. Surface modification of titanium implants with Mg-containing coatings to promote osseointegration. Acta Biomater. 2023, 169, 19–44. [Google Scholar] [CrossRef] [PubMed]
  11. Gaur, S.; Chugh, A.; Chaudhry, K.; Bajpayee, A.; Jain, G.; Chugh, V.K.; Kumar, P.; Singh, S. Efficacy and Safety of Concentrated Growth Factors and Platelet- Rich Fibrin on Stability and Bone Regeneration in Patients with Immediate Dental Implants: A Randomized Controlled Trial. Int. J. Oral Maxillofac. Implant. 2022, 37, 784–792. [Google Scholar] [CrossRef] [PubMed]
  12. Han, C.-H.; Kim, S.; Chung, M.-K.; Heo, S.-J.; Rhyu, I.-C.; Kwon, Y.; Chang, J.-S. Regenerated Bone Pattern Around Exposed Implants with Various Designs. Int. J. Oral Maxillofac. Implant. 2019, 34, 61–67. [Google Scholar] [CrossRef]
  13. Bang, S.-M.; Moon, H.-J.; Kwon, Y.-D.; Yoo, J.-Y.; Pae, A.; Kwon, I.K. Osteoblastic and osteoclastic differentiation on SLA and hydrophilic modified SLA titanium surfaces. Clin. Oral. Implant. Res. 2014, 25, 831–837. [Google Scholar] [CrossRef]
  14. Alavi, S.E.; Alavi, S.Z.; Gholami, M.; Sharma, A.; Sharma, L.A.; Shahmabadi, H.E. Biocomposite-based strategies for dental bone regeneration. Oral Surg. Oral Med. Oral Pathol. Oral Radiol. 2023, in press. [CrossRef] [PubMed]
  15. Tangri, S.; Hasan, N.; Kaur, J.; Mohammad, F.; Maan, S.; Kesharwani, P.; Ahmad, F.J. Drug loaded bioglass nanoparticles and their coating for efficient tissue and bone regeneration. J. Non-Cryst. Solids 2023, 616, 122469. [Google Scholar] [CrossRef]
  16. Shi, C.; Gao, J.; Wang, M.; Fu, J.; Wang, D.; Zhu, Y. Ultra-trace silver-doped hydroxyapatite with non-cytotoxicity and effective antibacterial activity. Mater. Sci. Eng. C Mater. Biol. Appl. 2015, 55, 497–505. [Google Scholar] [CrossRef] [PubMed]
  17. Li, Y.; Liu, X.; Tan, L.; Cui, Z.; Yang, X.; Yeung, K.W.K.; Pan, H.; Wu, S. Construction of N-halamine labeled silica/zinc oxide hybrid nanoparticles for enhancing antibacterial ability of Ti implants. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 76, 50–58. [Google Scholar] [CrossRef]
  18. Li, M.; Liu, Q.; Jia, Z.; Xu, X.; Shi, Y.; Cheng, Y.; Zheng, Y. Polydopamine-induced nanocomposite Ag/CaP coatings on the surface of titania nanotubes for antibacterial and osteointegration functions. J. Mater. Chem. B 2015, 45, 8796–8805. [Google Scholar] [CrossRef] [PubMed]
  19. Lung, C.Y.K.; Khan, A.S.; Zeeshan, R.; Akhtar, S.; Chaudhry, A.A.; Matinlinna, J.P. An antibacterial porous calcium phosphate bilayer functional coatings on titanium dental implants. Ceram. Int. 2023, 49, 2401–2409. [Google Scholar] [CrossRef]
  20. Andrade del Olmo, J.; Pérez-Álvarez, L.; Sáez Martínez, V.; Benito Cid, S.; Ruiz-Rubio, L.; Pérez González, R.; Vilas-Vilela, J.L.; Alonso, J.M. Multifunctional antibacterial chitosan-based hydrogel coatings on Ti6Al4V biomaterial for biomedical implant applications. Int. J. Biol. Macromol. 2023, 231, 123328. [Google Scholar] [CrossRef] [PubMed]
  21. Zhao, Z.; Wan, Y.; Yu, M.; Wang, H.; Cai, Y.; Liu, C.; Zhang, D. Biocompability evaluation of micro textures coated with zinc oxide on Ti-6Al-4V treated by nanosecond laser. Surf. Coat. Technol. 2021, 422, 127453. [Google Scholar] [CrossRef]
  22. Hashim, N.C.; Nordin, D. Nano-Hydroxyapatite Powder for Biomedical Implant Coating. Mater. Today 2019, 19, 1562–1571. [Google Scholar] [CrossRef]
  23. Mishra, S.; Chowdhary, R. PEEK materials as an alternative to titanium in dental implants: A systematic review. Clin. Implant Dent. Relat. Res. 2019, 21, 208–222. [Google Scholar] [CrossRef]
  24. Zou, R.; Bi, L.; Huang, Y.; Wang, Y.; Wang, Y.; Li, L.; Liu, J.; Feng, L.; Jiang, X.; Deng, B. A biocompatible silicon nitride dental implant material prepared by digital light processing technology. J. Mech. Behav. Biomed. 2023, 141, 105756. [Google Scholar] [CrossRef]
  25. Jayasree, R.; Raghava, K.; Sadhasivam, M.; Srinivas, P.V.V.; Vijay, R.; Pradeep, K.G.; Rao, T.N.; Chakravarty, D. Bi-layered metal-ceramic component for dental implants by spark plasma sintering. Mater. Lett. 2023, 344, 134403. [Google Scholar] [CrossRef]
  26. Zagorsky, V.A. Dental implantation. Materials and components. Symb. Sci. 2016, 9, 132–136. (In Russian) [Google Scholar]
  27. Jae-Hyun, L.; Kim, J.C.; Kim, H.-Y.; Yeo, I.-S.L. Influence of Connections and Surfaces of Dental Implants on Marginal Bone Loss: A Retrospective Study Over 7 to 19 Years. Int. J. Oral Maxillofac. Implant. 2020, 35, 1195–1202. [Google Scholar]
  28. Canullo, L.; Peñarrocha, D.; Clementini, M.; Iannello, G.; Micarelli, C. Impact of plasma of argon cleaning treatment on implant abutments in patients with a history of periodontal disease and thin biotype: Radiographic results at 24-month follow-up of a RCT. Clin. Oral. Implant. Res. 2015, 26, 8–14. [Google Scholar] [CrossRef] [PubMed]
  29. Ding, L.; Zhang, P.; Wang, X.; Kasugai, S. A doxycycline-treated hydroxyapatite implant surface attenuates the progression of peri-implantitis: A radiographic and histological study in mice. Clin. Implant Dent. Relat. Res. 2019, 21, 154–159. [Google Scholar] [CrossRef] [PubMed]
  30. Fraser, D.; Mendonca, G.; Sartori, E.; Funkenbusch, P.; Ercoli, C.; Meirelles, L. Bone response to porous tantalum implants in a gap-healing model. Clin. Oral. Implant. Res. 2019, 30, 156–168. [Google Scholar] [CrossRef] [PubMed]
  31. Anchieta, R.B.; Baldassarri, M.; Guastaldi, F.; Tovar, N.; Janal, M.N.; Gottlow, J.; Dard, M.; Jimbo, R.; Coelho, P.G. Mechanical property assessment of bone healing around a titanium-zirconium alloy dental implant. Clin. Implant Dent. Relat. Res. 2014, 16, 913–919. [Google Scholar] [CrossRef] [PubMed]
  32. Alves, S.A.; Rossi, A.L.; Ribeiro, A.R.; Toptan, F.; Pinto, A.M.; Shokuhfar, T.; Celis, J.-P.; Rocha, L.A. Improved tribocorrosion performance of bio-functionalized TiO2 nanotubes under two-cycle sliding actions in artificial saliva. J. Mech. Behav. Biomed. 2018, 80, 143–154. [Google Scholar] [CrossRef] [PubMed]
  33. Oh, J.-S.; Seo, Y.-S.; Lee, G.-J.; You, J.-S.; Kim, S.-G. A comparative study with biphasic calcium phosphate and deproteinized bovine bone in maxillary sinus augmentation: A prospective randomized controlled clinical trial. Int. J. Oral Maxillofac. Implant. 2019, 34, 233–242. [Google Scholar] [CrossRef]
  34. Jiang, Q.-H.; Gong, X.; Wang, X.-X.; He, F.-M. Osteogenesis of rat mesenchymal stem cells and osteoblastic cells on strontium-doped nanohydroxyapatite-coated titanium surfaces. Int. J. Oral Maxillofac. Implant. 2015, 30, 461–471. [Google Scholar] [CrossRef]
  35. Hunziker, E.B.; Spiegl-Habegger, M.; Rudolf, S.; Liu, Y.; Gu, Z.; Lippuner, K.; Shintani, N.; Enggist, L. A novel experimental dental implant permits quantitative grading of surface-property effects on osseointegration. Int. J. Oral Maxillofac. Implant. 2018, 33, 967–978. [Google Scholar] [CrossRef]
  36. Fardjahromi, M.A.; Ejeian, F.; Razmjou, A.; Vesey, G.; Mukhopadhyay, S.C.; Derakhshan, A.; Warkiani, M.E. Enhancing osteoregenerative potential of biphasic calcium phosphates by using bioinspired ZIF8 coating. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 123, 111972. [Google Scholar] [CrossRef] [PubMed]
  37. Bryington, M.; Mendonça, G.; Nares, S.; Cooper, L.F. Osteoblastic and cytokine gene expression of implant-adherent cells in humans. Clin. Oral. Implant. Res. 2014, 25, 52–58. [Google Scholar] [CrossRef]
  38. Zhou, C.; Ge, Z.; Song, L.; Yan, J.; Lang, X.; Zhang, Y.; He, F. Strontium-modified titanium substrate promotes osteogenic differentiation of MSCs and implant osseointegration via upregulating CDH2. Clin. Oral. Implant. Res. 2023, 34, 297–311. [Google Scholar] [CrossRef]
  39. Mehl, C.; Gaßling, V.; Schultz-Langerhans, S.; Açil, Y.; Bähr, T.; Wiltfang, J.; Kern, M. Influence of four different abutment materials and the adhesive joint of two-piece abutments on cervical implant bone and soft tissue. Int. J. Oral Maxillofac. Implant. 2016, 31, 1264–1272. [Google Scholar] [CrossRef]
  40. Chappuis, V.; Cavusoglu, Y.; Gruber, R.; Kuchler, U.; Buser, D.; Bosshardt, D.D. Osseointegration of Zirconia in the Presence of Multinucleated Giant Cells. Clin. Implant Dent. Relat. Res. 2016, 18, 686–698. [Google Scholar] [CrossRef] [PubMed]
  41. Zavan, B.; Ferroni, L.; Gardin, C.; Sivolella, S.; Piattelli, A.; Mijiritsky, E. Release of VEGF from dental implant improves osteogenetic process: Preliminary in vitro tests. Materials 2017, 10, 1052. [Google Scholar] [CrossRef] [PubMed]
  42. Gross, M.D. Occlusion in implant dentistry. A review of the literature of prosthetic determinants and current concepts. Aust. Dent. J. 2008, 53 (Suppl. 1), S60–S68. [Google Scholar] [CrossRef]
  43. Mishra, S.K.; Chowdhary, R.; Chrcanovic, B.R.; Brånemark, P.-I. Osseoperception in Dental Implants: A Systematic Review. J. Prosthodont. 2016, 25, 185–195. [Google Scholar] [CrossRef]
  44. González-Gil, D.; Dib-Zaitun, I.; Flores-Fraile, J.; López-Marcos, J. Active Tactile Sensibility in Implant Prosthesis vs. Complete Dentures: A Psychophysical Study. J. Clin. Med. 2022, 11, 6819. [Google Scholar] [CrossRef]
  45. Toledano-Serrabona, J.; Sánchez-Garcés, M.Á.; Gay-Escoda, C.; Valmaseda-Castellón, E.; Camps-Font, O.; Verdeguer, P.; Molmeneu, M.; Gil, F.J. Mechanical properties and corrosion behavior of Ti6Al4V particles obtained by implantoplasty: An in vitro study. Part II. Materials 2021, 14, 6519. [Google Scholar] [CrossRef]
  46. Sikora, C.L.; Alfaro, M.F.; Yuan, J.C.-C.; Barao, V.A.; Sukotjo, C.; Mathew, M.T. Wear and Corrosion Interactions at the Titanium/Zirconia Interface: Dental Implant Application. J. Prosthodont. 2018, 27, 842–852. [Google Scholar] [CrossRef]
  47. Revathi, A.; Borrás, A.D.; Muñoz, A.I.; Richard, C.; Manivasagam, G. Degradation mechanisms and future challenges of titanium and its alloys for dental implant applications in oral environment. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 76, 1354–1368. [Google Scholar] [CrossRef]
  48. Weller, J.; Vasudevan, P.; Kreikemeyer, B.; Ekat, K.; Jackszis, M.; Springer, A.; Chatzivasileiou, K.; Lang, H. The role of bacterial corrosion on recolonization of titanium implant surfaces: An in vitro study. Clin. Implant Dent. Relat. Res. 2022, 24, 664–675. [Google Scholar] [CrossRef]
  49. Silva, M.D.; Walton, T.R.; Alrabeah, G.O.; Layton, D.M.; Petridis, H. Comparison of corrosion products from implant and various golad-based abutment couplings: The effect of gold plating. J. Oral. Implant. 2021, 47, 370–379. [Google Scholar] [CrossRef]
  50. Eliaz, N. Corrosion of metallic biomaterials: A review. Materials 2019, 12, 407. [Google Scholar] [CrossRef]
  51. Trezubov, V.N.; Mishnev, L.M.; Zhulev, E.N.; Trezubov, V.V. Orthopedic Dentistry. Applied Materials Science, 7th ed.; MED Press-inform: Moscow, Russia, 2017; 328p. [Google Scholar]
  52. Grigoriev, S.; Sotova, C.; Vereschaka, A.; Uglov, V.; Cherenda, N. Modifying Coatings for Medical Implants Made of Titanium Alloys. Metals 2023, 13, 718. [Google Scholar] [CrossRef]
  53. Falanga, A.; Laheurte, P.; Vahabi, H.; Tran, N.; Khamseh, S.; Saeidi, H.; Khodadadi, M.; Zarrintaj, P.; Saeb, M.R.; Mozafari, M. Niobium-treated titanium implants with improved cellular and molecular activities at the tissue-implant interface. Materials 2019, 12, 3861. [Google Scholar] [CrossRef] [PubMed]
  54. Fintová, S.; Dlhý, P.; Mertová, K.; Chlup, Z.; Duchek, M.; Procházka, R.; Hutař, P. Fatigue properties of UFG Ti grade 2 dental implant vs. conventionally tested smooth specimens. J. Mech. Behav. Biomed. 2021, 123, 104715. [Google Scholar] [CrossRef] [PubMed]
  55. Monetta, T.; Acquesta, A.; Bellucci, F. Evaluation of roughness and electrochemical behavior of titanium in biological environment. Metall. Ital. 2014, 106, 13–21. [Google Scholar]
  56. Laurindo, C.A.H.; Torres, R.D.; Mali, S.A.; Gilbert, J.L.; Soares, P. Incorporation of Ca and P on anodized titanium surface: Effect of high current density. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 37, 223–231. [Google Scholar] [CrossRef] [PubMed]
  57. Babuska, V.; Palan, J.; Dobra, J.K.; Kulda, V.; Duchek, M.; Cerny, J.; Hrusak, D. Proliferation of osteoblasts on laser-modified nanostructured titanium surfaces. Materials 2018, 11, 1827. [Google Scholar] [CrossRef] [PubMed]
  58. Fintová, S.; Kuběna, I.; Palán, J.; Mertová, K.; Duchek, M.; Hutař, P.; Pastorek, F.; Kunz, L. Influence of sandblasting and acid etching on fatigue properties of ultra-fine grained Ti grade 4 for dental implants. J. Mech. Behav. Biomed. 2020, 111, 104016. [Google Scholar] [CrossRef] [PubMed]
  59. Wennerberg, A.; Jimbo, R.; Stübinger, S.; Obrecht, M.; Dard, M.; Berner, S. Nanostructures and hydrophilicity influence osseointegration: A biomechanical study in the rabbit tibia. Clin. Oral. Implant. Res. 2014, 25, 1041–1050. [Google Scholar] [CrossRef]
  60. Jacobs, N.; Seghi, R.; Johnston, W.M.; Yilmaz, B. Displacement and performance of abutments in narrow-diameter implants with different internal connections. J. Prosthet. Dent. 2022, 127, 100–106. [Google Scholar] [CrossRef] [PubMed]
  61. Catauro, M.; Bollino, F.; Papale, F.; Pacifico, S. Modulation of indomethacin release from ZrO2/PCL hybrid multilayers synthesized via solegel dip coating. J. Drug Deliv. Sci. Technol. 2015, 26, 10–16. [Google Scholar] [CrossRef]
  62. Moghaddam, N.S.; Skoracki, R.; Miller, M.; Elahinia, M.; Dean, D. Three Dimensional Printing of Stiffness-tuned, Nitinol Skeletal Fixation Hardware with an Example of Mandibular Segmental Defect Repair. Procedia CIRP 2016, 49, 45–50. [Google Scholar] [CrossRef]
  63. Mastrangelo, F.; Quaresima, R.; Abundo, R.; Spagnuolo, G.; Marenzi, G. Esthetic and physical changes of innovative titanium surface properties obtained with laser technology. Materials 2020, 13, 1066. [Google Scholar] [CrossRef]
  64. Szmukler-Moncler, S.; Blus, C.; Schwarz, D.M.; Orrù, G. Characterization of a macro-and micro-textured titanium grade 5 alloy surface obtained by etching only without sandblasting. Materials 2020, 13, 5074. [Google Scholar] [CrossRef]
  65. Ghensi, P.; Bettio, E.; Maniglio, D.; Bonomi, E.; Piccoli, F.; Gross, S.; Caciagli, P.; Segata, N.; Nollo, G.; Tessarolo, F. Dental implants with anti-biofilm properties: A pilot study for developing a new sericin-based coating. Materials 2019, 12, 2429. [Google Scholar] [CrossRef]
  66. Di Giulio, M.; Traini, T.; Sinjari, B.; Nostro, A.; Caputi, S.; Cellini, L. Porphyromonas gingivalis biofilm formation in different titanium surfaces, an in vitro study. Clin. Oral. Implant. Res. 2016, 27, 918–925. [Google Scholar] [CrossRef]
  67. Sancar, B.; Dayı, E. Evaluation of metal concentrations in hair and nails after dental implant placement. J. Prosthet. Dent. 2022, 128, 625–631. [Google Scholar] [CrossRef]
  68. More than Solid—Roxolid®. Reducing Invasiveness. Straumann® Roxolid®. Available online: https://www.straumann.com/content/dam/media-center/straumann/en/documents/interactive-sales-presentation/490.137-en_low.pdf (accessed on 20 August 2023).
  69. El Chaar, E.; Zhang, L.; Zhou, Y.; Sandgren, R.; Fricain, J.-C.; Dard, M.; Pippenger, B.; Catros, S. Osseointegration of superhydrophilic implants placed in defect grafted bones. Int. J. Oral Maxillofac. Implant. 2019, 34, 443–450. [Google Scholar] [CrossRef] [PubMed]
  70. Lotz, E.M.; Olivares-Navarrete, R.; Hyzy, S.L.; Berner, S.; Schwartz, Z.; Boyan, B.D. Comparable responses of osteoblast lineage cells to microstructured hydrophilic titanium–zirconium and microstructured hydrophilic titanium. Clin. Oral. Implant. Res. 2017, 28, e51–e59. [Google Scholar] [CrossRef] [PubMed]
  71. Xiao, W.; Chen, Y.; Chu, C.; Dard, M.M.; Man, Y. Influence of implant location on titanium-zirconium alloy narrow-diameter implants: A 1-year prospective study in smoking and nonsmoking populations. J. Prosthet. Dent. 2022, 128, 159–166. [Google Scholar] [CrossRef] [PubMed]
  72. Mishra, S.K.; Chowdhary, R. Evolution of dental implants through the work of per-ingvar branemark: A systematic review. Indian J. Dent. Res. 2020, 31, 930–956. [Google Scholar] [PubMed]
  73. Caparrós, C.; Ortiz-Hernandez, M.; Molmeneu, M.; Punset, M.; Calero, J.A.; Aparicio, C.; Fernández-Fairén, M.; Perez, R.; Gil, F.J. Bioactive macroporous titanium implants highly interconnected. J. Mater. Sci. Mater. Med. 2016, 27, 151. [Google Scholar] [CrossRef] [PubMed]
  74. Liens, A.; Etiemble, A.; Rivory, P.; Balvay, S.; Pelletier, J.-M.; Cardinal, S.; Fabrègue, D.; Kato, H.; Steyer, P.; Munhoz, T.; et al. On the potential of Bulk Metallic Glasses for dental implantology: Case study on Ti40Zr10Cu36Pd14. Materials 2018, 11, 249. [Google Scholar] [CrossRef] [PubMed]
  75. Li, P.; Ma, X.; Tong, T.; Wang, Y. Microstructural and mechanical properties of β-type Ti–Mo–Nb biomedical alloys with low elastic modulus. J. Alloys Compd. 2020, 815, 152412. [Google Scholar] [CrossRef]
  76. Ozan, S.; Lin, J.; Li, Y.; Ipek, R.; Wen, C. Development of Ti–Nb–Zr alloys with high elastic admissible strain for temporary orthopedic devices. Acta Biomater. 2015, 20, 176–187. [Google Scholar] [CrossRef]
  77. Boehlert, C.J.; Cowen, C.J.; Quast, J.P.; Akahori, T.; Niinomi, M. Fatigue and wear evaluation of Ti-Al-Nb alloys for biomedical applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2008, 28, 323–330. [Google Scholar] [CrossRef]
  78. Boehlert, C.J.; Cowen, C.J.; Jaeger, C.R.; Niinomi, M.; Akahori, T. Tensile and fatigue evaluation of Ti–15Al–33Nb (at.%) and Ti–21Al–29Nb (at.%) alloys for biomedical applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2005, 25, 263–275. [Google Scholar] [CrossRef]
  79. Hussain, S.A.; Panchal, M.; Allamraju, K.V.; Rajak, U.; Verma, T.N.; Brindhadevi, K. Optimization of wear behavior of heat-treated Ti-6Al-7Nb biomedical alloy by response surface methodology. Environ. Res. 2023, 231, 116193. [Google Scholar] [CrossRef] [PubMed]
  80. Kulesza, S.; Bramowicz, M.; Czaja, P.; Jabłoński, R.; Kropiwnicki, J.; Charkiewicz, M. Application of atomic force microscopy for studies of fractal and functional properties of biomaterials. Acta Phys. Pol. A 2016, 130, 1013–1015. [Google Scholar] [CrossRef]
  81. Msweli, N.P.; Akinwamide, S.O.; Olubambi, P.A.; Obadele, B.A. Microstructure and biocorrosion studies of spark plasma sintered yttria stabilized zirconia reinforced Ti6Al7Nb alloy in Hanks’ solution. Mater. Chem. Phys. 2023, 293, 126940. [Google Scholar] [CrossRef]
  82. Schulze, C.; Weinmann, M.; Schweigel, C.; Keßler, O.; Bader, R. Mechanical properties of a newly additive manufactured implant material based on Ti-42Nb. Materials 2018, 11, 124. [Google Scholar] [CrossRef] [PubMed]
  83. Huang, H.-H.; Tung, K.-L.; Lin, Y.-Y. Treating Ti containing dental orthodontic wires with nitrogen plasma immersion ion implantation to reduce the metal ions release and bacterial adhesion. Mater. Technol. 2015, 30, B73–B79. [Google Scholar] [CrossRef]
  84. Kalita, D.; Rogal, Ł.; Czeppe, T.; Wójcik, A.; Kolano-Burian, A.; Zackiewicz, P.; Kania, B.; Dutkiewicz, J. Microstructure and Mechanical Properties of Ti-Nb Alloys Prepared by Mechanical Alloying and Spark Plasma Sintering. J. Mater. Eng. Perform. 2020, 29, 1445–1452. [Google Scholar] [CrossRef]
  85. Hoppe, V.; Szymczyk-Ziółkowska, P.; Rusińska, M.; Dybała, B.; Poradowski, D.; Janeczek, M. Assessment of mechanical, chemical, and biological properties of Ti-Nb-Zr alloy for medical applications. Materials 2021, 14, 126. [Google Scholar] [CrossRef]
  86. Pérez-Pevida, E.; Brizuela-Velasco, A.; Chávarri-Prado, D.; Jiménez-Garrudo, A.; Sánchez-Lasheras, F.; Solaberrieta-Méndez, E.; Diéguez-Pereira, M.; Fernández-González, F.J.; Dehesa-Ibarra, B.; Monticelli, F. Biomechanical Consequences of the Elastic Properties of Dental Implant Alloys on the Supporting Bone: Finite Element Analysis. BioMed Res. Int. 2016, 2016, 1850401. [Google Scholar] [CrossRef]
  87. Hussein, M.A.; Suryanarayana, C.; Al-Aqeeli, N. Fabrication of nano-grained Ti-Nb-Zr biomaterials using spark plasma sintering. Mater. Design 2015, 87, 693–700. [Google Scholar] [CrossRef]
  88. Cordeiro, J.M.; Nagay, B.E.; Ribeiro, A.L.R.; Cruz, N.C.; Rangel, E.C.; Fais, L.M.G.; Vaz, L.G.; Barão, V.A.R. Functionalization of an experimental Ti-Nb-Zr-Ta alloy with a biomimetic coating produced by plasma electrolytic oxidation. J. Alloys Compd. 2019, 770, 1038–1048. [Google Scholar] [CrossRef]
  89. Lin, Z.; Wang, L.; Xue, X.; Lu, W.; Qin, J.; Zhang, D. Microstructure evolution and mechanical properties of a Ti–35Nb–3Zr–2Ta biomedical alloy processed by equal channel angular pressing (ECAP). Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 4551–4561. [Google Scholar] [CrossRef]
  90. Meng, X.; Wang, X.; Guo, Y.; Ma, S.; Luo, W.; Xiang, X.; Zhao, J.; Zhou, Y. Biocompatibility evaluation of a newly developed Ti-Nb-Zr-Ta-Si alloy implant. J. Biomater. Tissue Eng. 2016, 6, 861–869. [Google Scholar] [CrossRef]
  91. Angelescu, R.M.; Rəducanu, D.; Cojocaru, V.D.; Angelescu, M.L.; Buţu, M.; Cincə, I.; Dan, I. Microstructural and mechanical evaluation of a Ti-Nb-Ta alloy. Univ. Politeh. Buchar. Sci. Bull. Ser. B 2015, 77, 221–228. [Google Scholar]
  92. Mutlu, I.; Yeniyol, S.; Oktay, E. Production and Precipitation Hardening of Beta-Type Ti-35Nb-10Cu Alloy Foam for Implant Applications. J. Mater. Eng. Perform. 2016, 25, 1586–1593. [Google Scholar] [CrossRef]
  93. Alberta, L.A.; Vishnu, J.; Hariharan, A.; Pilz, S.; Gebert, A.; Calin, M. Novel low modulus beta-type Ti–Nb alloys by gallium and copper minor additions for antibacterial implant applications. J. Mater. Res. Technol. 2022, 20, 3306–3322. [Google Scholar] [CrossRef]
  94. Alberta, L.A.; Vishnu, J.; Douest, Y.; Perrin, K.; Trunfio-Sfarghiu, A.-M.; Courtois, N.; Gebert, A.; Ter-Ovanessian, B.; Calin, M. Tribocorrosion behavior of β-type Ti-Nb-Ga alloys in a physiological solution. Tribol. Int. 2023, 181, 108325. [Google Scholar] [CrossRef]
  95. Alberta, L.A.; Fortouna, Y.; Vishnu, J.; Pilz, S.; Gebert, A.; Lekka, C.; Nielsch, K.; Calin, M. Effects of Ga on the structural, mechanical and electronic properties of β-Ti-45Nb alloy by experiments and ab initio calculations. J. Mech. Behav. Biomed. 2023, 140, 105728. [Google Scholar] [CrossRef] [PubMed]
  96. Cardoso, G.C.; Kuroda, P.A.B.; Grandini, C.R. Influence of Nb addition on the structure, microstructure, Vickers microhardness, and Young’s modulus of new β Ti-xNb-5Mo alloys system. J. Mater. Res. Technol. 2023, 25, 3061–3070. [Google Scholar] [CrossRef]
  97. Wang, Z.; Li, Y.; Huang, W.; Chen, X.; He, H. Micro-abrasion–corrosion behaviour of a biomedical Ti–25Nb–3Mo–3Zr–2Sn alloy in simulated physiological fluid. J. Mech. Behav. Biomed. 2016, 63, 361–374. [Google Scholar] [CrossRef]
  98. Li, H.; Cai, Q.; Li, S.; Xu, H. Effects of Mo equivalent on the phase constituent, microstructure and compressive mechanical properties of Ti–Nb–Mo–Ta alloys prepared by powder metallurgy. J. Mater. Res. Technol. 2022, 16, 588–598. [Google Scholar] [CrossRef]
  99. Koch, M.; Burkovski, A.; Zulla, M.; Rosiwal, S.; Geißdörfer, W.; Dittmar, R.; Grobecker-Karl, T. Pilot study on the use of a laser-structured double diamond electrode (Dde) for biofilm removal from dental implant surfaces. J. Clin. Med. 2020, 9, 3036. [Google Scholar] [CrossRef]
  100. Latimer, J.M.; Roll, K.L.; Daubert, D.M.; Zhang, H.; Shalev, T.; Wolff, L.F.; Kotsakis, G.A. Clinical performance of hydrophilic, titanium-zirconium dental implants in patients with well-controlled and poorly controlled type 2 diabetes: One-year results of a dual-center cohort study. J. Periodontol. 2022, 93, 745–757. [Google Scholar] [CrossRef]
  101. Medvedev, A.E.; Molotnikov, A.; Lapovok, R.; Zeller, R.; Berner, S.; Habersetzer, P.; Dalla Torre, F. Microstructure and mechanical properties of Ti-15Zr alloy used as dental implant material. J. Mech. Behav. Biomed. 2016, 62, 384–398. [Google Scholar] [CrossRef]
  102. Cordeiro, J.M.; Faverani, L.P.; Grandini, C.R.; Rangel, E.C.; da Cruz, N.C.; Nociti, F.H., Jr.; Almeida, A.B.; Vicente, F.B.; Morais, B.R.G.; Barão, V.A.R.; et al. Characterization of chemically treated Ti-Zr system alloys for dental implant application. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 92, 849–861. [Google Scholar] [CrossRef] [PubMed]
  103. Brizuela-Velasco, A.; Pérez-Pevida, E.; Jiménez-Garrudo, A.; Gil-Mur, F.J.; Manero, J.M.; Punset-Fuste, M.; Chávarri-Prado, D.; Diéguez-Pereira, M.; Monticelli, F. Mechanical characterisation and biomechanical and biological behaviours of Ti-Zr binary-Alloy dental implants. BioMed Res. Int. 2017, 2017, 2785863. [Google Scholar] [CrossRef] [PubMed]
  104. Okazaki, Y.; Katsuda, S.-I. Biological safety evaluation and surface modification of biocompatible ti–15Zr–4Nb alloy. Materials 2021, 14, 731. [Google Scholar] [CrossRef] [PubMed]
  105. Eldabah, N.M.; Shoukry, A.; Khair-Eldeen, W.; Kobayashi, S.; Gepreel, M.A.-H. Design and characterization of low Young’s modulus Ti-Zr-Nb-based medium entropy alloys assisted by extreme learning machine for biomedical applications. J. Alloys Compd. 2023, 968, 171755. [Google Scholar] [CrossRef]
  106. Qi, P.; Li, B.; Wang, T.; Zhou, L.; Nie, Z. Microstructure and properties of a novel ternary Ti–6Zr–xFe alloy for biomedical applications. J. Alloys Compd. 2021, 854, 157119. [Google Scholar] [CrossRef]
  107. Bai, L.; Cui, C.; Wang, Q.; Bu, S.; Qi, Y. Ti–Zr–Fe–Si system amorphous alloys with excellent biocompatibility. J. Non-Cryst. Solids 2008, 354, 3935–3938. [Google Scholar] [CrossRef]
  108. Huang, Q.; Xu, S.; Ouyang, Z.; Yang, Y.; Liu, Y. Multi-scale nacre-inspired lamella-structured Ti-Ta composites with high strength and low modulus for load-bearing orthopedic and dental applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 118, 111458. [Google Scholar] [CrossRef] [PubMed]
  109. Schlee, M.; Pradies, G.; Mehmke, W.-U.; Beneytout, A.; Stamm, M.; Meda, R.G.; Kamm, T.; Poiroux, F.; Weinlich, F.; del Canto Pingarron, M.; et al. Prospective, Multicenter Evaluation of Trabecular Metal-Enhanced Titanium Dental Implants Placed in Routine Dental Practices: 1-Year Interim Report from the Development Period (2010 to 2011). Clin. Implant Dent. Relat. Res. 2015, 17, 1141–1153. [Google Scholar] [CrossRef]
  110. Aguilar, C.; San Martín, F.; Martínez, C.; Cámara, B.; Claverías, F.; Undabarrena, A.; Sancy, M.; Salinas, V.; Muñoz, L. Improving mechanical properties and antibacterial response of α/β ternary Ti–Ta alloy foams for biomedical uses. J. Mater. Res. Technol. 2023, 24, 8735–8753. [Google Scholar] [CrossRef]
  111. Mareci, D.; Chelariu, R.; Gordin, D.-M.; Ungureanu, G.; Gloriant, T. Comparative corrosion study of Ti–Ta alloys for dental applications. Acta Biomater. 2009, 5, 3625–3639. [Google Scholar] [CrossRef]
  112. Mendis, S.; Xu, W.; Tang, H.P.; Jones, L.A.; Liang, D.; Thompson, R.; Choong, P.; Brandt, M.; Qian, M. Characteristics of oxide films on Ti-(10–75)Ta alloys and their corrosion performance in an aerated Hank’s balanced salt solution. Appl. Surf. Sci. 2020, 506, 145013. [Google Scholar] [CrossRef]
  113. Kim, W.-G.; Choe, H.-C. Effects of TiN coating on the corrosion of nanostructured Ti–30Ta–xZr alloys for dental implants. Appl. Surf. Sci. 2012, 258, 1929–1934. [Google Scholar] [CrossRef]
  114. Kim, H.-J.; Jeong, Y.-H.; Choe, H.-C.; Brantley, W.A. Surface morphology of TiN-coated nanotubular Ti–25Ta–xZr alloys for dental implants prepared by RF sputtering. Thin Solid Films 2013, 549, 131–134. [Google Scholar] [CrossRef]
  115. Stenlund, P.; Omar, O.; Brohede, U.; Norgren, S.; Norlindh, B.; Johansson, A.; Lausmaa, J.; Thomsen, P.; Palmquist, A. Bone response to a novel Ti–Ta–Nb–Zr alloy. Acta Biomater. 2015, 20, 165–175. [Google Scholar] [CrossRef]
  116. Nguyen, P.M.H.; Won, D.-H.; Kim, B.-S.; Jang, Y.-S.; Nguyen, T.-D.T.; Lee, M.-H.; Bae, T.-S. The effect of two-step surface modification for Ti-Ta-Mo-Zr alloys on bone regeneration: An evaluation using calvarial defect on rat model. Appl. Surf. Sci. 2018, 442, 630–639. [Google Scholar] [CrossRef]
  117. Fowler, L.; Masia, N.; Cornish, L.A.; Chown, L.H.; Engqvist, H.; Norgren, S.; Öhman-Mägi, C. Development of antibacterial Ti-Cux alloys for dental applications: Effects of ageing for alloys with up to 10 wt% Cu. Materials 2019, 12, 4017. [Google Scholar] [CrossRef]
  118. Wang, X.; Dong, H.; Liu, J.; Qin, G.; Chen, D.; Zhang, E. In vivo antibacterial property of Ti-Cu sintered alloy implant. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 100, 38–47. [Google Scholar] [CrossRef]
  119. Pina, V.G.; Amigó, V.; Muñoz, A.I. Microstructural, electrochemical and tribo-electrochemical characterisation of titanium-copper biomedical alloys. Corros. Sci. 2016, 109, 115–125. [Google Scholar] [CrossRef]
  120. Alshammari, Y.; Yang, F.; Bolzoni, L. Fabrication and characterisation of low-cost powder metallurgy Ti-xCu-2.5Al alloys produced for biomedical applications. J. Mech. Behav. Biomed. 2022, 126, 105022. [Google Scholar] [CrossRef]
  121. Pandey, A.K.; Gautam, R.K.; Behera, C.K. Corrosion and wear behavior of Ti–5Cu-xNb biomedical alloy in simulated body fluid for dental implant applications. J. Mech. Behav. Biomed. 2023, 137, 105533. [Google Scholar] [CrossRef] [PubMed]
  122. Xu, W.; Lu, X.; Wang, L.N.; Shi, Z.M.; Lv, S.M.; Qian, M.; Qu, X.H. Mechanical properties, in vitro corrosion resistance and biocompatibility of metal injection molded Ti-12Mo alloy for dental applications. J. Mech. Behav. Biomed. 2018, 88, 534–547. [Google Scholar] [CrossRef]
  123. Mohan, P.; Rajak, D.K.; Pruncu, C.I.; Behera, A.; Amigó-Borrás, V.; Elshalakany, A.B. Influence of β-phase stability in elemental blended Ti-Mo and Ti-Mo-Zr alloys. Micron 2021, 142, 102992. [Google Scholar] [CrossRef] [PubMed]
  124. Asl, M.S.; Delbari, S.A.; Azadbeh, M.; Namini, A.S.; Mehrabian, M.; Nguyen, V.-H.; Van Le, Q.; Shokouhimehr, M.; Mohammadi, M. Nanoindentational and conventional mechanical properties of spark plasma sintered Ti–Mo alloys. J. Mater. Res. Technol. 2020, 9, 10647–10658. [Google Scholar]
  125. Mohan, P.; Elshalakany, A.B.; Osman, T.A.; Amigo, V.; Mohamed, A. Effect of Fe content, sintering temperature and powder processing on the microstructure, fracture and mechanical behaviours of Ti-Mo-Zr-Fe alloys. J. Alloys Compd. 2017, 729, 1215–1225. [Google Scholar] [CrossRef]
  126. Tejeda-Ochoa, A.; Kametani, N.; Carreño-Gallardo, C.; Ledezma-Sillas, J.E.; Adachi, N.; Todaka, Y.; Herrera-Ramirez, J.M. Formation of a metastable fcc phase and high Mg solubility in the Ti-Mg system by mechanical alloying. Powder Technol. 2020, 374, 348–352. [Google Scholar] [CrossRef]
  127. Cai, X.; Ding, S.; Li, Z.; Zhang, X.; Wen, K.; Xu, L.; Zhang, Y.; Peng, Y.; Shen, T. Simultaneous sintering of low-melting-point Mg with high-melting-point Ti via a novel one-step high-pressure solid-phase sintering strategy. J. Alloys Compd. 2021, 858, 158344. [Google Scholar] [CrossRef]
  128. Liu, Y.; Li, K.; Luo, T.; Song, M.; Wu, H.; Xiao, J.; Tan, Y.; Cheng, M.; Chen, B.; Niu, X.; et al. Powder metallurgical low-modulus Ti–Mg alloys for biomedical applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2015, 56, 241–250. [Google Scholar] [CrossRef] [PubMed]
  129. Balog, M.; Hassan Ibrahim, A.M.; Krizik, P.; Bajana, O.; Klimova, A.; Catic, A.; Schauperl, Z. Bioactive Ti + Mg composites fabricated by powder metallurgy: The relation between the microstructure and mechanical properties. J. Mech. Behav. Biomed. 2019, 90, 45–53. [Google Scholar] [CrossRef] [PubMed]
  130. Pradeep, N.B.; Rajath Hegde, M.M.; Rajendrachari, S.; Surendranathan, A.O. Investigation of microstructure and mechanical properties of microwave consolidated TiMgSr alloy prepared by high energy ball milling. Powder Technol. 2022, 408, 117715. [Google Scholar] [CrossRef]
  131. Niu, J.; Guo, Y.; Li, K.; Liu, W.; Dan, Z.; Sun, Z.; Chang, H.; Zhou, L. Improved mechanical, bio-corrosion properties and in vitro cell responses of Ti-Fe alloys as candidate dental implants. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 122, 111917. [Google Scholar] [CrossRef]
  132. Sjafrizal, T.; Kenta, D.; Dehghan-Manshadi, A.; Xiao, W.; Dargusch, M.S. Metastable Ti-Fe-Ge alloys with high elastic admissible strain. Materialia 2022, 21, 101304. [Google Scholar] [CrossRef]
  133. Haruna, T. Development of Ti alloys for dental devices with corrosion resistance to fluoride solutions. Zairyo-to-Kankyo/Corros. Eng. 2014, 63, 309–315. [Google Scholar] [CrossRef]
  134. Wang, Z.B.; Hu, H.X.; Zheng, Y.G.; Ke, W.; Qiao, Y.X. Comparison of the corrosion behavior of pure titanium and its alloys in fluoride-containing sulfuric acid. Corros. Sci. 2016, 103, 50–65. [Google Scholar] [CrossRef]
  135. Wang, X.-L.; Zhou, Q.; Yang, K.; Zou, C.-H.; Wang, L. Performance of surface on ultrafine grained Ti-0.2Pd in simulated body fluid. Appl. Surf. Sci. 2018, 434, 957–966. [Google Scholar] [CrossRef]
  136. Hua, F.; Mon, K.; Pasupathi, P.; Gordon, G.; Shoesmith, D. A review of corrosion of titanium grade 7 and other titanium alloys in nuclear waste repository environments. Corrosion 2005, 61, 987–1003. [Google Scholar] [CrossRef]
  137. Steinemann, S.G.; Perren, S.M. Titanium alloys as metallic biomaterials. In Proceedings of the Fifth World Conference On Titanium, Munich, Germany, 10–14 September 1984; Volume 2, pp. 1327–1334. [Google Scholar]
  138. Sidelnikov, A.I. Comparative characteristics of materials of the titanium group used in the production of modern dental implants. InfoDENT 2000, 5, 10–12. [Google Scholar]
  139. Mirza, A.; King, A.; Troakes, C.; Exley, C. Aluminium in brain tissue in familial Alzheimer’s disease. J. Trace Elem. Med. Biol. 2017, 40, 30–36. [Google Scholar] [CrossRef] [PubMed]
  140. Cremasco, A.; Messias, A.D.; Esposito, A.R.; Duek, E.A.R.; Caram, R. Effects of alloying elements on the cytotoxic response of titanium alloys. Mater. Sci. Eng. C Mater. Biol. Appl. 2011, 31, 833–839. [Google Scholar] [CrossRef]
  141. Lin, C.-W.; Ju, C.-P.; Lin, J.-H.C. A comparison of the fatigue behavior of cast Ti–7.5Mo with c.p. titanium, Ti–6Al–4V and Ti–13Nb–13Zr alloys. Biomaterials 2005, 26, 2899–2907. [Google Scholar] [CrossRef] [PubMed]
  142. Cardoso, G.C.; de Almeida, G.S.; Corrêa, D.O.G.; Zambuzzi, W.F.; Buzalaf, M.A.R.; Correa, D.R.N.; Grandini, C.R. Preparation and characterization of novel as-cast Ti-Mo-Nb alloys for biomedical applications. Sci. Rep. 2022, 12, 11874. [Google Scholar] [CrossRef]
  143. Albrektsson, T.; Hansson, H.A.; Ivarsson, B. Interface analysis of titanium and zirconium bone implants. Biomaterials 1985, 6, 97–101. [Google Scholar] [CrossRef] [PubMed]
  144. Eisenbarth, E.; Velten, D.; Müller, M.; Thull, R.; Breme, J. Biocompatibility of β-stabilizing elements of titanium alloys. Biomaterials 2004, 25, 5705–5713. [Google Scholar] [CrossRef]
  145. Okazaki, Y.; Gotoh, E. Comparison of metal release from various metallic biomaterials in vitro. Biomaterials 2005, 26, 11–21. [Google Scholar] [CrossRef]
  146. Fikeni, L.; Annan, K.A.; Mutombo, K.; Machaka, R. Effect of Nb content on the microstructure and mechanical properties of binary Ti-Nb alloys. Mater. Today 2021, 38, 913–917. [Google Scholar] [CrossRef]
  147. Cremasco, A.; Osório, W.R.; Freire, C.M.A.; Garcia, A.; Caram, R. Electrochemical corrosion behavior of a Ti–35Nb alloy for medical prostheses. Electrochim. Acta 2008, 53, 4867–4874. [Google Scholar] [CrossRef]
  148. Souza, J.G.S.; Bertolini, M.M.; Costa, R.C.; Nagay, B.E.; Dongari-Bagtzoglou, A.; Barão, V.A.R. Targeting implant-associated infections: Titanium surface loaded with antimicrobial. iScience 2021, 24, 102008. [Google Scholar] [CrossRef]
  149. Sarraf, M.; Ghomi, E.R.; Alipour, S.; Ramakrishna, S.; Sukiman, N.L. A state-of-the-art review of the fabrication and characteristics of titanium and its alloys for biomedical applications. Bio-Design Manuf. 2022, 5, 371–395. [Google Scholar] [CrossRef]
  150. Lemire, J.; Harrison, J.; Turner, R. Antimicrobial activity of metals: Mechanisms, molecular targets and applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef] [PubMed]
  151. Zhao, Q.; Ueno, T.; Wakabayashi, N. A review in titanium-zirconium binary alloy for use in dental implants: Is there an ideal Ti-Zr composing ratio? Jpn. Dent. Sci. Rev. 2023, 59, 28–37. [Google Scholar] [CrossRef]
  152. Correa, D.R.N.; Vicente, F.B.; Donato, T.A.G.; Arana-Chavez, V.E.; Buzalaf, M.A.R.; Grandini, C.R. The effect of the solute on the structure, selected mechanical properties, and biocompatibility of Ti–Zr system alloys for dental applications. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 34, 354–359. [Google Scholar] [CrossRef]
  153. Ho, W.-F.; Chen, W.-K.; Wu, S.-C.; Hsu, H.-C. Structure, mechanical properties, and grindability of dental Ti–Zr alloys. J. Mater. Sci. Mater. Med. 2008, 19, 3179–3186. [Google Scholar] [CrossRef]
  154. Hsu, H.-C.; Wu, S.-C.; Sung, Y.-C.; Ho, W.-F. The structure and mechanical properties of as-cast Zr–Ti alloys. J. Alloys Compd. 2009, 488, 279–283. [Google Scholar] [CrossRef]
  155. Cordeiro, J.M.; Beline, T.; Ribeiro, A.L.R.; Rangel, E.C.; da Cruz, N.C.; Landers, R.; Faverani, L.P.; Vaz, L.G.; Fais, L.M.G.; Vicentei, F.B.; et al. Development of binary and ternary titanium alloys for dental implants. Dent. Mater. 2017, 33, 1244–1257. [Google Scholar] [CrossRef] [PubMed]
  156. Grandin, H.M.; Berner, S.; Dard, M. A Review of Titanium Zirconium (TiZr) Alloys for Use in Endosseous Dental Implants. Materials 2012, 5, 1348–1360. [Google Scholar] [CrossRef]
  157. Hacisalihoglu, I.; Samancioglu, A.; Yildiz, F.; Purcek, G.; Alsaran, A. Tribocorrosion properties of different type titanium alloys in simulated body fluid. Wear 2015, 332–333, 679–686. [Google Scholar] [CrossRef]
  158. Froes, F.H. Titanium Alloys: Thermal Treatment and Thermomechanical Processing. In Encyclopedia of Materials: Science and Technology, 2nd ed.; Buschow, K.H.J., Cahn, R.W., Flemings, M.C., Ilschner, B., Kramer, E.J., Mahajan, S., Veyssière, P., Eds.; Elsevier Ltd.: Exeter Devon, UK, 2001; pp. 9369–9373. [Google Scholar]
  159. Hon, Y.-H.; Wang, J.-Y.; Pan, Y.-N. Composition/Phase Structure and Properties of Titanium-Niobium Alloys. Mater. Trans. 2003, 44, 2384–2390. [Google Scholar] [CrossRef]
  160. ISO 10993; Biological Evaluation of Medical Devices. The International Organization for Standard (ISO): Geneva, Switzerland, 2010.
  161. Trillo, E.A.; Ortiz, C.; Dickerson, P.; Villa, R.; Stafford, S.W.; Murr, L.E. Evaluation of mechanical and corrosion biocompatibility of TiTa alloys. J. Mater. Sci. Mater. Med. 2001, 12, 283–292. [Google Scholar] [CrossRef] [PubMed]
  162. Zhou, Y.L.; Niinomi, M.; Akahori, T. Effects of Ta content on Young’s modulus and tensile properties of binary Ti–Ta alloys for biomedical applications. Mater. Sci. Eng. A-Struct. 2004, 371, 283–290. [Google Scholar] [CrossRef]
  163. Schildhauer, T.; Robie, B.; Muhr, G.; Köller, M. Bacterial Adherence to Tantalum Versus Commonly Used Orthopedic Metallic Implant Materials. J. Orthop. Trauma 2006, 20, 476–484. [Google Scholar] [CrossRef]
  164. Wang, X.; Ning, B.; Pei, X. Tantalum and its derivatives in orthopedic and dental implants: Osteogenesis and antibacterial properties. Colloid. Surf. B 2021, 208, 112055. [Google Scholar] [CrossRef]
  165. Zhang, W.; Liu, Y.; Wu, H.; Song, M.; Zhang, T.; Lan, X.; Yao, T. Elastic modulus of phases in Ti–Mo alloys. Mater. Charact. 2015, 106, 302–307. [Google Scholar] [CrossRef]
  166. Issariyapat, A.; Huang, J.; Teramae, T.; Kariya, S.; Bahador, A.; Visuttipitukul, P.; Umeda, J.; Alhazaa, A.; Kondoh, K. Microstructure refinement and strengthening mechanisms of additively manufactured Ti-Zr alloys prepared from pre-mixed feedstock. Addit. Manuf. 2023, 73, 103649. [Google Scholar] [CrossRef]
  167. Yang, D.; Guo, Z.; Shao, H.; Liu, X.; Ji, Y. Mechanical Properties of Porous Ti-Mo and Ti-Nb Alloys for Biomedical Application by Gelcasting. Procedia Eng. 2012, 36, 160–167. [Google Scholar] [CrossRef]
  168. Wei, C.; Luo, L.; Wu, Z.; Zhang, J.; Su, S.; Zhan, Y. New Zr–25Ti–xMo alloys for dental implant application: Properties characterization and surface analysis. J. Mech. Behav. Biomed. 2020, 111, 104017. [Google Scholar] [CrossRef] [PubMed]
  169. Ruiz Henao, P.A.; Caneiro Queija, L.; Mareque, S.; Tasende Pereira, A.; Liñares González, A.; Blanco Carrión, J. Titanium vs. ceramic single dental implants in the anterior maxilla: A 12-month randomized clinical trial. Clin. Implant Dent. Relat. Res. 2021, 32, 951–961. [Google Scholar] [CrossRef]
  170. Siddiqi, A.; Khan, A.S.; Zafar, S. Thirty years of translational research in zirconia dental implants: A systematic review of the literature. J. Oral. Implant. 2017, 43, 314–326. [Google Scholar] [CrossRef] [PubMed]
  171. Siddiqi, A.; Kieser, J.A.; De Silva, R.K.; Thomson, W.M.; Duncan, W.J. Soft and Hard Tissue Response to Zirconia versus Titanium One-Piece Implants Placed in Alveolar and Palatal Sites: A Randomized Control Trial. Clin. Implant Dent. Relat. Res. 2015, 17, 483–496. [Google Scholar] [CrossRef] [PubMed]
  172. Thoma, D.S.; Ioannidis, A.; Cathomen, E.; Hämmerle, C.H.; Hüsler, J.; Jung, R.E. Discoloration of the peri-implant mucosa caused by zirconia and titanium implants. Int. J. Periodontics Restor. 2016, 36, 38–45. [Google Scholar] [CrossRef]
  173. Chiou, L.-L.; Panariello, B.H.D.; Hamada, Y.; Gregory, R.L.; Blanchard, S.; Duarte, S. Comparison of In Vitro Biofilm Formation on Titanium and Zirconia Implants. BioMed Res. Int. 2023, 2023, 8728499. [Google Scholar] [CrossRef]
  174. Yegorov, A.A.; Drovosekov, M.N.; Aronov, A.M.; Rozhnova, O.M.; Yegorova, O.P. Comparative characteristics of materials used in dental implantation. Bull. Sib. Med. 2014, 13, 41–47. (In Russian) [Google Scholar] [CrossRef]
  175. Paraskevich, V.L. Dental Implantology: Fundamentals of Theory and Practice, 3rd ed.; Medical Information Agency LLC: Moscow, Russia, 2011; 400p. (In Russian) [Google Scholar]
  176. Korniienko, V.; Oleshko, O.; Husak, Y.; Deineka, V.; Holubnycha, V.; Mishchenko, O.; Kazek-Kęsik, A.; Jakóbik-Kolon, A.; Pshenychnyi, R.; Leśniak-Ziółkowska, K.; et al. Formation of a bacteriostatic surface on ZrNb alloy via anodization in a solution containing Cu nanoparticles. Materials 2020, 13, 3913. [Google Scholar] [CrossRef] [PubMed]
  177. Oleshko, O.; Deineka, V.V.; Husak, Y.; Korniienko, V.; Mishchenko, O.; Holubnycha, V.; Pisarek, M.; Michalska, J.; Kazek-Kesik, A.; Jakóbik-Kolon, A.; et al. Ag nanoparticle-decorated oxide coatings formed via plasma electrolytic oxidation on ZrNb alloy. Materials 2019, 12, 3742. [Google Scholar] [CrossRef] [PubMed]
  178. Kondo, R.; Nomura, N.; Suyalatu; Tsutsumi, Y.; Doi, H.; Hanawa, T. Microstructure and mechanical properties of as-cast Zr–Nb alloys. Acta Biomater. 2011, 7, 4278–4284. [Google Scholar] [CrossRef] [PubMed]
  179. Cui, W.F.; Liu, N.; Qin, G.W. Microstructures, mechanical properties and corrosion resistance of the Zr-xTi (Ag) alloys for dental implant application. Mater. Chem. Phys. 2016, 176, 161–166. [Google Scholar] [CrossRef]
  180. Akimoto, T.; Ueno, T.; Tsutsumi, Y.; Doi, H.; Hanawa, T.; Wakabayashi, N. Evaluation of corrosion resistance of implant-use Ti-Zr binary alloys with a range of compositions. J. Biomed. Mater. Res. B Appl. Biomater. 2018, 106, 73–79. [Google Scholar] [CrossRef] [PubMed]
  181. Bolat, G.; Izquierdo, J.; Santana, J.; Mareci, D.; Souto, R.M. Electrochemical characterization of ZrTi alloys for biomedical applications. Electrochim. Acta 2013, 88, 447–456. [Google Scholar] [CrossRef]
  182. Bolat, G.; Izquierdo, J.; Mareci, D.; Sutiman, D.; Souto, R.M. Electrochemical characterization of ZrTi alloys for biomedical applications. Part 2: The effect of thermal oxidation. Electrochim. Acta 2013, 106, 432–439. [Google Scholar] [CrossRef]
  183. Mishchenko, O.; Ovchynnykov, O.; Kapustian, O.; Pogorielov, M. New Zr-Ti-Nb alloy for medical application: Development, chemical and mechanical properties, and biocompatibility. Materials 2020, 13, 1306. [Google Scholar] [CrossRef]
  184. Michalska, J.; Sowa, M.; Stolarczyk, A.; Warchoł, F.; Nikiforow, K.; Pisarek, M.; Dercz, G.; Pogorielov, M.; Mishchenko, O.; Simka, W. Plasma electrolytic oxidation of Zr-Ti-Nb alloy in phosphate-formate-EDTA electrolyte. Electrochim. Acta 2022, 419, 140375. [Google Scholar] [CrossRef]
  185. Roehling, S.; Gahlert, M.; Janner, S.; Meng, B.; Woelfler, H.; Cochran, D.L. Ligature-induced peri-implant bone loss around loaded zirconia and titanium implants. Int. J. Oral Maxillofac. Implant. 2019, 34, 357–365. [Google Scholar] [CrossRef]
  186. Bahadirli, G.; Yilmaz, S.; Jones, T.; Sen, D. Influences of implant and framework materials on stress distribution: A three-dimensional finite element analysis study. Int. J. Oral Maxillofac. Implant. 2018, 33, e117–e126. [Google Scholar] [CrossRef]
  187. Sancho-Puchades, M.; Hämmerle, C.H.F.; Benic, G.I. In vitro assessment of artifacts induced by titanium, titanium-zirconium and zirconium dioxide implants in cone-beam computed tomography. Clin. Oral. Implant. Res. 2015, 26, 1222–1228. [Google Scholar] [CrossRef] [PubMed]
  188. Steiger-Ronay, V.; Krcmaric, Z.; Schmidlin, P.R.; Sahrmann, P.; Wiedemeier, D.B.; Benic, G.I. Assessment of peri-implant defects at titanium and zirconium dioxide implants by means of periapical radiographs and cone beam computed tomography: An in-vitro examination. Clin. Oral. Implant. Res. 2018, 29, 1195–1201. [Google Scholar] [CrossRef]
  189. Benic, G.I.; Thoma, D.S.; Sanz-Martin, I.; Munoz, F.; Hämmerle, C.H.F.; Cantalapiedra, A.; Fischer, J.; Jung, R.E. Guided bone regeneration at zirconia and titanium dental implants: A pilot histological investigation. Clin. Oral. Implant. Res. 2017, 28, 1592–1599. [Google Scholar] [CrossRef]
  190. Schriber, M.; Yeung, A.W.K.; Suter, V.G.A.; Buser, D.; Leung, Y.Y.; Bornstein, M.M. Cone beam computed tomography artefacts around dental implants with different materials influencing the detection of peri-implant bone defects. Clin. Oral. Implant. Res. 2020, 31, 595–606. [Google Scholar] [CrossRef]
  191. Saito, M.M.; Onuma, K.; Yamamoto, R.; Yamakoshi, Y. New insights into bioactivity of ceria-stabilized zirconia: Direct bonding to bone-like hydroxyapatite at nanoscale. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 121, 111665. [Google Scholar] [CrossRef]
  192. Palmero, P.; Fornabaio, M.; Montanaro, L.; Reveron, H.; Esnouf, C.; Chevalier, J. Towards long lasting zirconia-based composites for dental implants: Part I: Innovative synthesis, microstructural characterization and invitro stability. Biomaterials 2015, 50, 38–46. [Google Scholar] [CrossRef]
  193. Schienle, S.; Al-Ahmad, A.; Kohal, R.J.; Bernsmann, F.; Adolfsson, E.; Montanaro, L.; Palmero, P.; Fürderer, T.; Chevalier, J.; Hellwig, E.; et al. Microbial adhesion on novel yttria-stabilized tetragonal zirconia (Y-TZP) implant surfaces with nitrogen-doped hydrogenated amorphous carbon (a-C:H:N) coatings. Clin. Oral. Investig. 2016, 20, 1719–1732. [Google Scholar] [CrossRef] [PubMed]
  194. Brüll, F.C.W.E.; van Winkelhoff, A.J.; Cune, M.S. Zirconia dental implants: A clinical, radiographic, and microbiologic evaluation up to 3 years. Int. J. Oral Maxillofac. Implant. 2014, 29, 914–920. [Google Scholar]
  195. da Cruz, M.B.; Marques, J.F.; Peñarrieta-Juanito, G.M.; Costa, M.; Souza, J.C.M.; Magini, R.S.; Miranda, G.; Silva, F.S.; da Mata, A.D.S.P.; Caramês, J.M.M. Hard and soft tissue cell behavior on PEEK, Zirconia, and Titanium implant materials. Int. J. Oral Maxillofac. Implant. 2019, 34, 39–46. [Google Scholar] [CrossRef] [PubMed]
  196. Payer, M.; Heschl, A.; Koller, M.; Arnetzl, G.; Lorenzoni, M.; Jakse, N. All-ceramic restoration of zirconia two-piece implants—A randomized controlled clinical trial. Clin. Oral. Implant. Res. 2015, 26, 371–376. [Google Scholar] [CrossRef] [PubMed]
  197. Oshima, Y.; Iwasa, F.; Tachi, K.; Baba, K. Effect of nanofeatured topography on ceria-stabilized zirconia/alumina nanocomposite on osteogenesis and osseointegration. Int. J. Oral Maxillofac. Implant. 2017, 32, 81–91. [Google Scholar] [CrossRef] [PubMed]
  198. Hagiwara, Y.; Nakabayashi, S.; Ikeda, T.; Ito, R. Ceria-Stabilized Zirconia/Alumina nanocomposite for fabricating the framework of removable dental prostheses: Preliminary results from a 4-Year Follow-up. Int. J. Prosthodont. 2019, 32, 254–256. [Google Scholar] [CrossRef] [PubMed]
  199. Hagiwara, Y.; Nakajima, K. Use of ceria-stabilized zirconia/alumina nanocomposite for fabricating the frameworks of removable dental prostheses: A clinical report. J. Prosthet. Dent. 2016, 116, 166–171. [Google Scholar] [CrossRef]
  200. Lopez-Píriz, R.; Fernández, A.; Goyos-Ball, L.; Rivera, S.; Díaz, L.A.; Fernández-Domínguez, M.; Prado, C.; Moya, J.S.; Torrecillas, R. Performance of a new Al2O3/Ce-TZP ceramic nanocomposite dental implant: A pilot study in dogs. Materials 2017, 10, 614. [Google Scholar] [CrossRef]
  201. Molaei, M.; Attarzadeh, N.; Fattah-Alhosseini, A. Tailoring the biological response of zirconium implants using zirconia bioceramic coatings: A systematic review. J. Trace Elem. Med. Biol. 2021, 66, 126756. [Google Scholar] [CrossRef]
  202. Kim, H.J.; Jeong, Y.H.; Brantley, W.A.; Choe, H.C. Nanotube nucleation phenomena on Ti−25Ta−xZr alloys for implants using ATO technique. J. Nanosci. Nanotechnol. 2014, 14, 7569–7573. [Google Scholar] [CrossRef] [PubMed]
  203. Holländer, J.; Lorenz, J.; Stübinger, S.; Hölscher, W.; Heidemann, D.; Ghanaati, S.; Sader, R. Zirconia dental implants: Investigation of clinical parameters, patient satisfaction, and microbial contamination. Int. J. Oral Maxillofac. Implant. 2016, 31, 855–864. [Google Scholar] [CrossRef] [PubMed]
  204. Nascimento, C.D.; Pita, M.S.; Fernandes, F.H.N.C.; Pedrazzi, V.; de Albuquerque, R.F., Jr.; Ribeiro, R.F. Bacterial adhesion on the titanium and zirconia abutment surfaces. Clin. Oral. Implant. Res. 2014, 25, 337–343. [Google Scholar] [CrossRef] [PubMed]
  205. Öztürk, C.; Çelik, E.; Gönüldaş, F. Effect of different surface treatments on the biaxial flexural strength of zirconia ceramics. J. Prosthet. Dent. 2023, 129, 220.e1–220.e5. [Google Scholar] [CrossRef] [PubMed]
  206. Chatchai, D. The Structure of the Mechanical Properties of Composites Based on Zirconium Dioxide and Wollastonite. Dissertation for the Degree of Candidate of Technical Sciences, Tomsk Polytechnic University, Tomsk, Russia, 2021. [Google Scholar]
  207. ISO 13356:2015; Implants for Surgery—Ceramic Materials Based on Yttria-Stabilized Tetragonal Zirconia (Y-TZP). ISO: Geneva, Switzerland, 2015.
  208. Hasiak, M.; Sobieszczańska, B.; Łaszcz, A.; Biały, M.; Chęcmanowski, J.; Zatoński, T.; Bożemska, E.; Wawrzyńska, M. Production, mechanical properties and biomedical characterization of zrti-based bulk metallic glasses in comparison with 316L stainless steel and Ti6Al4V alloy. Materials 2022, 15, 252. [Google Scholar] [CrossRef] [PubMed]
  209. Piotrowski, B.; Baptista, A.A.; Patoor, E.; Bravetti, P.; Eberhardt, A.; Laheurte, P. Interaction of bone-dental implant with new ultra low modulus alloy using a numerical approach. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 38, 151–160. [Google Scholar] [CrossRef] [PubMed]
  210. Dmitrievskii, A.A.; Zhigacheva, D.G.; Zhigachev, A.O.; Ovchinnikov, P.N. Strength Properties of Zirconium Ceramics with Silica Additives Hardened with Alumina. Phys. Solid State 2021, 63, 295–299. [Google Scholar] [CrossRef]
  211. Dovgerd, A.; Sivolapov, K. Ceramic implants are the future of dental implantology. Actual. Probl. Dent. 2022, 18, 23–31. [Google Scholar] [CrossRef]
  212. Dovgerd, A.; Sivolapov, K. Features and differences of biofilm formation in the field of ceramic and titanium implants. Actual. Probl. Dent. 2023, 19, 5–11. [Google Scholar] [CrossRef]
  213. Kulakov, O.B.; Doktorov, A.A.; D’iakova, S.V.; Denisov-Nikol’skiĭ, I.; Grötz, K.A. Experimental study of osseointegration of zirconium and titanium dental implants. Morfologiia 2005, 127, 52–55. [Google Scholar]
  214. Reveron, H.; Fornabaio, M.; Palmero, P.; Fürderer, T.; Adolfsson, E.; Lughi, V.; Bonifacio, A.; Sergo, V.; Montanaro, L.; Chevalier, J. Towards long lasting zirconia-based composites for dental implants: Transformation induced plasticity and its consequence on ceramic reliability. Acta Biomater. 2017, 48, 423–432. [Google Scholar] [CrossRef]
  215. Tanaka, S.; Takaba, M.; Ishiura, Y.; Kamimura, E.; Baba, K. A 3-year follow-up of ceria-stabilized zirconia/alumina nanocomposite (Ce-TZP/A) frameworks for fixed dental prostheses. J. Prosthodont. Res. 2015, 59, 55–61. [Google Scholar] [CrossRef] [PubMed]
  216. Saulacic, N.; Erdosi, R.; Bosshardt, D.D.; Gruber, R.; Buser, D. Acid and alkaline etching of sandblasted zirconia implants: A histomorphometric study in miniature pigs. Clin. Implant. Dent. Res. 2014, 16, 313–322. [Google Scholar] [CrossRef]
  217. Fontoura, D.C.; Barros, V.M.; de Magalhães, C.S.; Vaz, R.R.; Moreira, A.N. Evaluation of vertical misfit of CAD/CAM implant-supported titanium and zirconia frameworks. Int. J. Oral Maxillofac. Implant. 2018, 33, 1027–1032. [Google Scholar] [CrossRef] [PubMed]
  218. Tsymbalov, O.V. Dental Implantation for Periodontal Diseases (Monograph); EDVI: Krasnodar, Russia, 2014. [Google Scholar]
  219. Marin, E. Forged to heal: The role of metallic cellular solids in bone tissue engineering. Mater. Today Bio 2023, 23, 100777. [Google Scholar] [CrossRef] [PubMed]
  220. Barrack, R.L.; Burak, C.; Skinner, H.B. Concerns about ceramics in THA. Clin. Orthop. Relat. R. 2004, 429, 73–79. [Google Scholar] [CrossRef] [PubMed]
  221. Grigoriev, S.N.; Volosova, M.A.; Okunkova, A.A. Advances in Electrical Discharge Machining of Insulating Ceramics. Materials 2023, 16, 5959. [Google Scholar] [CrossRef] [PubMed]
  222. Grigoriev, S.N.; Volosova, M.A.; Peretyagin, P.Y.; Seleznev, A.E.; Okunkova, A.A.; Smirnov, A. The effect of TiC additives on mechanical and electrical properties of Al2O3 ceramic. Appl. Sci. 2018, 8, 2385. [Google Scholar] [CrossRef]
  223. Vereschaka, A.S.; Grigoriev, S.N.; Sotova, E.S.; Vereschaka, A.A. Improving the efficiency of the cutting tools made of mixed ceramics by applying modifying nano-scale multilayered coatings. Adv. Mater. Res. 2013, 712–715, 391–394. [Google Scholar] [CrossRef]
  224. Grigoriev, S.N.; Vereschaka, A.A.; Vereschaka, A.S.; Kutin, A.A. Cutting tools made of layered composite ceramics with nano-scale multilayered coatings. Procedia CIRP 2012, 1, 301–306. [Google Scholar] [CrossRef]
  225. Solís, N.W.; Peretyagin, P.; Torrecillas, R.; Fernández, A.; Menéndez, J.L.; Mallada, C.; Díaz, L.A.; Moya, J.S. Electrically conductor black zirconia ceramic by SPS using graphene oxide. J. Electroceram. 2017, 38, 119–124. [Google Scholar] [CrossRef]
  226. Chopra, D.; Jayasree, A.; Guo, T.; Gulati, K.; Ivanovski, S. Advancing dental implants: Bioactive and therapeutic modifications of zirconia. Bioact. Mater. 2022, 13, 161–178. [Google Scholar] [CrossRef] [PubMed]
  227. Titanium. Available online: https://www.allmetals.ru/metals/titanium/ (accessed on 10 September 2023).
  228. Tantalum. Available online: https://www.allmetals.ru/metals/tantalum/ (accessed on 10 September 2023).
  229. Liu, Y.; Bao, C.; Wismeijer, D.; Wu, G. The physicochemical/biological properties of porous tantalum and the potential surface modification techniques to improve its clinical application in dental implantology. Mater. Sci. Eng. C Mater. Biol. Appl. 2015, 49, 323–329. [Google Scholar] [CrossRef] [PubMed]
  230. Stiehler, M.; Lind, M.; Mygind, T.; Baatrup, A.; Dolatshahi-Pirouz, A.; Li, H.; Foss, M.; Besenbacher, F.; Kassem, M.; Bünger, C. Morphology, proliferation, and osteogenic differentiation of mesenchymal stem cells cultured on titanium, tantalum, and chromium surfaces. J. Biomed. Mater. Res. A 2008, 86, 448–458. [Google Scholar] [CrossRef] [PubMed]
  231. Liu, X.; Song, X.; Zhang, P.; Zhu, Z.; Xu, X. Effects of nano tantalum implants on inducing osteoblast proliferation and differentiation. Exp. Ther. Med. 2016, 12, 3541–3544. [Google Scholar] [CrossRef] [PubMed]
  232. Zhu, Y.; Gu, Y.; Qiao, S.; Zhou, L.; Shi, J.; Lai, H. Bacterial and mammalian cells adhesion to tantalum-decorated micro-/nano-structured titanium. J. Biomed. Mater. Res. A 2017, 105, 871–878. [Google Scholar] [CrossRef]
  233. Piglionico, S.; Bousquet, J.; Fatima, N.; Renaud, M.; Collart-dutilleul, P.-Y.; Bousquet, P. Porous tantalum vs. titanium implants: Enhanced mineralized matrix formation after stem cells proliferation and differentiation. J. Clin. Med. 2020, 9, 3657. [Google Scholar] [CrossRef]
  234. Magic, M.; Zupanek, G.; Lazic, Z.; El Chaar, E. Primary Stability of Trabecular Metal Implant in Comparison to Fully Threaded Implants: In Vitro Study Simulating Immediate Implant Placement. J. Oral. Implant. 2022, 48, 584–589. [Google Scholar] [CrossRef]
  235. Fraser, D.; Funkenbusch, P.; Ercoli, C.; Meirelles, L. Biomechanical analysis of the osseointegration of porous tantalum implants. J. Prosthet. Dent. 2020, 123, 811–820. [Google Scholar] [CrossRef]
  236. Bencharit, S.; Morelli, T.; Barros, S.; Seagroves, J.T.; Kim, S.; Yu, N.; Byrd, K.; Brenes, C.; Offenbacher, S. Comparing Initial Wound Healing and Osteogenesis of Porous Tantalum Trabecular Metal and Titanium Alloy Materials. J. Oral. Implant. 2019, 45, 173–180. [Google Scholar] [CrossRef]
  237. Benoit, B.; Fouad, Z.; Laflamme, G.H.; Rouleau, D.; Laflamme, G.Y. Augmentation of tibial plateau fractures with Trabecular Metal™: A biomechanical study. J. Orthop. Surg. Res. 2009, 4, 37. [Google Scholar] [CrossRef]
  238. Gorbatyuk, D.S.; Kolesov, S.V.; Sazhnev, M.L.; Pereverzev, V.S.; Kazmin, A.I. Tantalum based implants experimental and clinical aspects of application. Bull. Traumatol. Orthop. After N. N. Priorova 2018, 25, 71–83. [Google Scholar] [CrossRef]
  239. Liu, Y.; Rath, B.; Tingart, M.; Eschweiler, J. Role of implants surface modification in osseointegration: A systematic review. J. Biomed. Mater. Res. A 2020, 108, 470–484. [Google Scholar] [CrossRef] [PubMed]
  240. Meneghini, R.M.; Meyer, C.; Buckley, C.A.; Hanssen, A.D.; Lewallen, D.G. Mechanical Stability of Novel Highly Porous Metal Acetabular Components in Revision Total Hip Arthroplasty. J. Arthroplast. 2010, 25, 337–341. [Google Scholar] [CrossRef] [PubMed]
  241. Karageorgiou, V.; Kaplan, D. Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials 2005, 26, 5474–5491. [Google Scholar] [CrossRef] [PubMed]
  242. Cachinho, S.C.P.; Correia, R.N. Titanium scaffolds for osteointegration: Mechanical, in vitro and corrosion behaviour. J. Mater. Sci. Mater. Med. 2008, 19, 451–457. [Google Scholar] [CrossRef] [PubMed]
  243. Zardiackas, L.D.; Parsell, D.E.; Dillon, L.D.; Mitchell, D.W.; Nunnery, L.A.; Poggie, R. Structure, metallurgy, and mechanical properties of a porous tantalum foam. J. Biomed. Mater. Res. B Appl. Biomater. 2001, 58, 180–187. [Google Scholar] [CrossRef]
  244. Peron, C.; Romanos, G. Immediate placement and occlusal loading of single-tooth restorations on partially threaded, titanium-tantalum combined dental implants: 1-year results. Int. J. Periodontics Restor. 2016, 36, 392–399. [Google Scholar] [CrossRef]
  245. Brauner, E.; Di Carlo, S.; Ciolfi, A.; Pompa, G.; Jamshir, S.; De Angelis, F.; Della Monaca, M.; Valentini, V. Use of porous implants for the prosthetic rehabilitation of fibula free flap reconstructed patients. J. Craniofac. Surg. 2019, 30, 1163–1169. [Google Scholar] [CrossRef]
  246. Schlee, M.; van der Schoor, W.P.; van der Schoor, A.R.M. Immediate loading of trabecular metal-enhanced titanium dental implants: Interim results from an international proof-of-principle study. Clin. Implant Dent. Relat. Res. 2015, 17, e308–e320. [Google Scholar] [CrossRef]
  247. Zhou, X.; Hu, X.; Lin, Y. Coating of Sandblasted and Acid-Etched Dental Implants with Tantalum Using Vacuum Plasma Spraying. Implant Dent. 2018, 27, 202–208. [Google Scholar] [CrossRef] [PubMed]
  248. Almeida Alves, C.F.; Cavaleiro, A.; Carvalho, S. Bioactivity response of Ta1-xOx coatings deposited by reactive DC magnetron sputtering. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 58, 110–118. [Google Scholar] [CrossRef] [PubMed]
  249. Beline, T.; da Silva, J.H.D.; Matos, A.O.; Azevedo Neto, N.F.; de Almeida, A.B.; Nociti, F.H., Jr.; Leite, D.M.G.; Rangel, E.C.; Barão, V.A.R. Tailoring the synthesis of tantalum-based thin films for biomedical application: Characterization and biological response. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 101, 111–119. [Google Scholar] [CrossRef] [PubMed]
  250. Lu, T.; Wen, J.; Qian, S.; Cao, H.; Ning, C.; Pan, X.; Jiang, X.; Liu, X.; Chu, P.K. Enhanced osteointegration on tantalum-implanted polyetheretherketone surface with bone-like elastic modulus. Biomaterials 2015, 51, 173–183. [Google Scholar] [CrossRef]
  251. Cui, J.; Zhao, L.; Zhu, W.; Wang, B.; Zhao, C.; Fang, L.; Ren, F. Antibacterial activity, corrosion resistance and wear behavior of spark plasma sintered Ta-5Cu alloy for biomedical applications. J. Mech. Behav. Biomed. 2017, 74, 315–323. [Google Scholar] [CrossRef]
  252. Hlinka, J.; Kraus, M.; Hajnys, J.; Pagac, M.; Petru, J.; Brytan, Z.; Tanski, T. Complex corrosion properties of aisi 316L steel prepared by 3D printing technology for possible implant applications. Materials 2020, 13, 1527. [Google Scholar] [CrossRef]
  253. Kayali, Y.; Aslan, O.; Karabaş, M.; Talaş, Ş. Corrosion behaviour of single and double layer hydroxyapatite coatings on 316L stainless steel by plasma spray. Prot. Met. Phys. Chem. Surf. 2016, 52, 1079–1085. [Google Scholar] [CrossRef]
  254. Pavlic, A.; Perissinotto, F.; Turco, G.; Contardo, L.; Stjepan, S. Do chlorhexidine and probiotics solutions provoke corrosion of orthodontic mini-implants? An in vitro study. Int. J. Oral Maxillofac. Implant. 2019, 34, 1379–1388. [Google Scholar] [CrossRef]
  255. Marques, R.A.; Rogero, S.O.; Terada, M.; Pieretti, E.F.; Costa, I. Localized Corrosion Resistance and Cytotoxicity Evaluation of Ferritic Stainless Steels for Use in Implantable Dental Devices with Magnetic Connections. Int. J. Electrochem. Sci. 2014, 9, 1340–1354. [Google Scholar] [CrossRef]
  256. Dos Santos, L.C.P.; Malheiros, F.C.; Guarato, A.Z. Surface parameters of as-built additive manufactured metal for intraosseous dental implants. J. Prosthet. Dent. 2020, 124, 217–222. [Google Scholar] [CrossRef] [PubMed]
  257. ISO 5832-1; Implants for Surgery—Metallic Materials—Part 1: Wrought Stainless Steel. The International Organization for Standard (ISO): Geneva, Switzerland, 2007.
  258. Aminian, A.; Shirzadi, B.; Azizi, Z.; Maedler, K.; Volkmann, E.; Hildebrand, N.; Maas, M.; Treccani, L.; Rezwan, K. Enhanced cell adhesion on bioinert ceramics mediated by the osteogenic cell membrane enzyme alkaline phosphatase. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 69, 184–194. [Google Scholar] [CrossRef] [PubMed]
  259. Bulut, B.; Erkmen, Z.E.; Kayali, E.S. Biocompatibility of hydroxyapatite-alumina and hydroxyapatite-zirconia composites including commercial inert glass (CIG) as a ternary component. J. Ceram. Sci. Technol. 2016, 7, 263–275. [Google Scholar]
  260. Badran, Z.; Struillou, X.; Hughes, F.J.; Soueidan, A.; Hoornaert, A.; Ide, M. Silicon nitride (Si3N4) implants: The future of dental implantology? J. Oral. Implant. 2017, 43, 240–244. [Google Scholar] [CrossRef] [PubMed]
  261. Purcel, G.; Surdu, V.-A.; Stoleriu, S. Influence of spark plasma sintering conditions on Si3N4 ceramics used for bone implants. Rev. Rom. Mater. 2015, 45, 160–170. [Google Scholar]
  262. Da Cruz, M.B.; Marques, J.F.; Peñarrieta-Juanito, G.M.; Costa, M.; Souza, J.C.M.; Magini, R.S.; Miranda, G.; Silva, F.S.; Caramês, J.M.M.; Da Mata, A.S.P. Bioactive-enhanced polyetheretherketone dental implant materials: Mechanical characterization and cellular responses. J. Oral. Implant. 2021, 47, 9–17. [Google Scholar] [CrossRef] [PubMed]
  263. Abyzov, A.M. Aluminum oxide and aluminum oxide ceramics (review). Part 2. Foreign manufacturers of aluminum oxide ceramics. Technologies and research in the field of alumina ceramics. Novye Ogneupory 2019, 2, 13–22. (In Russian) [Google Scholar] [CrossRef]
  264. Grigoriev, S.N.; Volosova, M.A.; Maslov, A.R.; Okunkova, A.A.; Smurov, I.Y. Highly Efficient Processing Technologies (Monograph); Mashinostroenie: Moscow, Russia, 2014; 455p. (In Russian) [Google Scholar]
  265. Sotova, E.S.; Vereshchaka, A.S.; Vereshchaka, A.A. Ceramic Cutting Tools (Monograph); MSUT “STANKIN”: Moscow, Russia, 2013; 149p. (In Russian) [Google Scholar]
  266. Grigoriev, S.N.; Hamdy, K.; Volosova, M.A.; Okunkova, A.A.; Fedorov, S.V. Electrical discharge machining of oxide and nitride ceramics: A review. Mater. Design 2021, 209, 109965. [Google Scholar] [CrossRef]
  267. Liu, X.-J.; Huang, Z.-Y.; Ge, Q.-M.; Sun, X.-W.; Huang, L.-P. Microstructure and mechanical properties of silicon nitride ceramics prepared by pressureless sintering with MgO–Al2O3–SiO2 as sintering additive. J. Eur. Ceram. Soc. 2005, 25, 3353–3359. [Google Scholar] [CrossRef]
  268. Guo, Z.; Huang, Y.; Sun, C.; He, Z.; Li, Y.; Qiu, H.; Yuan, D.; Zhang, J.; Chu, C.; Shen, B. Preparation and characterization of metastable β-type titanium alloys with favorable properties for orthopedic applications. J. Alloys Compds 2023, 949, 169839. [Google Scholar] [CrossRef]
  269. Skinner, H.B. Composite technology for total hip arthroplasty. Clin. Orthop. Relat. Res. 1988, 235, 224–236. [Google Scholar] [CrossRef]
  270. Schwitalla, A.; Müller, W.-D. PEEK Dental Implants: A Review of the Literature. J. Oral. Implant. 2013, 39, 743–749. [Google Scholar] [CrossRef]
  271. Zhang, H.; Guo, Z.; Zhang, Z.; Wu, G.; Sang, L. Biomimetic design and fabrication of PEEK and PEEK/CF cage with minimal surface structures by fused filament fabrication. J. Mater. Res. Technol. 2023, 26, 5001–5015. [Google Scholar] [CrossRef]
  272. Rakotoaridina, K.; Delrieu, J.; Pages, P.; Vergé, T.; Nasr, K.; Canceill, T. Evaluation of Poly(etheretherketone) Post’s Mechanical Strength in Comparison with Three Metal-Free Biomaterials: An In Vitro Study. Polymers 2023, 15, 3583. [Google Scholar] [CrossRef]
  273. Polyaryletherketone. Available online: https://rusplast.com/catalog/polyaryletherketone/ (accessed on 10 November 2023).
  274. Bilyalov, A.R.; Minasov, B.S.; Yakupov, R.R.; Akbashev, V.N.; Rafikova, G.A.; Bikmeev, A.T.; Chugunov, S.S.; Kireev, V.N.; Pavlov, V.N.; Krzyszkowska, Y.G. Using ceramic 3D printing for tissue engineering problems: A review. Polytrauma 2023, 1, 89–109. [Google Scholar]
  275. Webster, T.J.; Patel, A.A.; Rahaman, M.N.; Bal, B.S. Anti-infective and osteointegration properties of silicon nitride, poly(ether ether ketone), and titanium implants. Acta Biomater. 2012, 8, 4447–4454. [Google Scholar] [CrossRef] [PubMed]
  276. El Awadly, T.A.; Wu, G.; Ayad, M.; Radi, I.A.W.; Wismeijer, D.; Abo El Fetouh, H.; Osman, R.B. A histomorphometric study on treated and untreated ceramic filled PEEK implants versus titanium implants: Preclinical in vivo study. Clin. Oral. Implant. Res. 2020, 31, 246–254. [Google Scholar] [CrossRef] [PubMed]
  277. Nakahara, I.; Takao, M.; Goto, T.; Ohtsuki, C.; Hibino, S.; Sugano, N. Interfacial shear strength of bioactive-coated carbon fiber reinforced polyetheretherketone after in vivo implantation. J. Orthop. Res. 2012, 30, 1618–1625. [Google Scholar] [CrossRef]
  278. Najeeb, S.; Khurshid, Z.; Zohaib, S.; Zafar, M.S. Bioactivity and osseointegration of PEEK are inferior to those of titanium: A systematic review. J. Oral. Implant. 2016, 42, 512–516. [Google Scholar] [CrossRef]
  279. Gheisarifar, M.; Thompson, G.A.; Drago, C.; Tabatabaei, F.; Rasoulianboroujeni, M. In vitro study of surface alterations to polyetheretherketone and titanium and their effect upon human gingival fibroblasts. J. Prosthet. Dent. 2021, 125, 155–164. [Google Scholar] [CrossRef] [PubMed]
  280. Elawadly, T.; Radi, W.I.A.; El Khadem, A.; Osman, R.B. Can PEEK be an implant material? Evaluation of surface topography and wettability of filled versus unfilled peek with different surface roughness. J. Oral. Implant. 2017, 43, 456–461. [Google Scholar] [CrossRef]
  281. Sirandoni, D.; Leal, E.; Weber, B.; Noritomi, P.Y.; Fuentes, R.; Borie, E. Effect of different framework materials in implant-supported fixed mandibular prostheses: A finite element analysis. Int. J. Oral Maxillofac. Implant. 2019, 34, e107–e114. [Google Scholar] [CrossRef] [PubMed]
  282. Souza, R.C.; Santos, L.A.; Baptista, C.A.R.P.; Bicalho, L.A.; Strecker, K.; Barboza, M.J.R.; Santos, C. Subcritical crack growth and fatigue life prediction of ZrO2-Al2O3 composite ceramic. Ceramica 2014, 60, 41–51. [Google Scholar] [CrossRef]
  283. Bartolomeu, F.; Buciumeanu, M.; Costa, M.M.; Alves, N.; Gasik, M.; Silva, F.S.; Miranda, G. Multi-material Ti6Al4V & PEEK cellular structures produced by Selective Laser Melting and Hot Pressing: A tribocorrosion study targeting orthopedic applications. J. Mech. Behav. Biomed. 2019, 89, 54–64. [Google Scholar]
  284. Warreth, A.; Ibieyou, N.; O’Leary, R.B.; Cremonese, M.; Abdulrahim, M. Dental implants: An overview. Dent. Update 2017, 44, 596–620. [Google Scholar] [CrossRef]
  285. Demiyanova, A.V.; Mitrofanov, E.A.; Simakin, S.B.; Sipkin, A.M. Analysis of ion-plasma technology application in medical products proceedings to be used for maxillofacial surgery (Review). IOP Conf. Ser-Mater. Sci. 2020, 781, 012015. [Google Scholar] [CrossRef]
  286. Kulakov, A.A.; Grigoryan, A.S.; Arkhipov, A.V. Impact of surface modifications of dental implants on their integration potential. Stomatologiya 2012, 91, 75–77. (In Russian) [Google Scholar]
  287. Lang, M.S.; Roselyn Cerutis, D.; Miyamoto, T.; Nunn, M.E. Cell attachment following instrumentation with titanium and plastic instruments, diode laser, and titanium brush on titanium, titanium-zirconium, and zirconia surfaces. Int. J. Oral Maxillofac. Implant. 2016, 31, 799–806. [Google Scholar] [CrossRef]
  288. Guarnieri, R.; Rappelli, G.; Piemontese, M.; Procaccini, M.; Quaranta, A. A double-blind randomized trial comparing implants with laser-microtextured and machined collar surfaces: Microbiologic and clinical results. Int. J. Oral Maxillofac. Implant. 2016, 31, 1117–1125. [Google Scholar] [CrossRef]
  289. Xing, R.; Lyngstadaas, S.P.; Ellingsen, J.E.; Taxt-Lamolle, S.; Haugen, H.J. The influence of surface nanoroughness, texture and chemistry of TiZr implant abutment on oral biofilm accumulation. Clin. Oral. Implant. Res. 2015, 26, 649–656. [Google Scholar] [CrossRef]
  290. Guarnieri, R.; Miccoli, G.; Reda, R.; Mazzoni, A.; Di Nardo, D.; Testarelli, L. Sulcus fluid volume, IL-6, and Il-1b concentrations in periodontal and peri-implant tissues comparing machined and laser-microtextured collar/abutment surfaces during 12 weeks of healing: A split-mouth RCT. Clin. Oral. Implant. Res. 2022, 33, 94–104. [Google Scholar] [CrossRef]
  291. Camarda, A.J.; Durand, R.; Benkarim, M.; Rompré, P.H.; Guertin, G.; Ciaburro, H. Prospective randomized clinical trial evaluating the effects of two different implant collar designs on peri-implant healing and functional osseointegration after 25 years. Clin. Oral. Implant. Res. 2021, 32, 285–296. [Google Scholar] [CrossRef] [PubMed]
  292. Kim, Y.-S.; Ko, Y.; Kye, S.-B.; Yang, S.-M. Human gingival fibroblast (HGF-1) attachment and proliferation on several abutment materials with various colors. Int. J. Oral Maxillofac. Implant. 2014, 29, 969–975. [Google Scholar] [CrossRef] [PubMed]
  293. Prati, C.; Zamparini, F.; Scialabba, V.S.; Gatto, M.R.; Piattelli, A.; Montebugnoli, L.; Gandolfi, M.G. A 3-year prospective cohort study on 132 calcium phosphate-blasted implants: Flap vs. flapless technique. Int. J. Oral Maxillofac. Implant. 2016, 31, 413–423. [Google Scholar] [CrossRef] [PubMed]
  294. Corvino, E.; Pesce, P.; Mura, R.; Marcano, E.; Canullo, L. Influence of modified titanium abutment surface on peri-implant soft tissue behavior: A systematic review of in vitro studies. Int. J. Oral Maxillofac. Implant. 2020, 35, 503–519. [Google Scholar] [CrossRef]
  295. El-Din El-Helbawy, N.G.; El-Wahab El-Hatery, A.A.; Ahmed, M.H. Comparison of oxygen plasma treatment and sandblasting of titanium implant-abutment surface on bond strength and surface topography. Int. J. Oral Maxillofac. Implant. 2016, 31, 555–562. [Google Scholar] [CrossRef]
  296. Turker, N.T.; Özarslan, M.M.; Buyukkaplan, U.S.; Başar, E.K. Effect of Different Surface Treatments Applied to Short Zirconia and Titanium Abutments. Int. J. Oral Maxillofac. Implant. 2020, 35, 948–954. [Google Scholar] [CrossRef] [PubMed]
  297. Ding, Q.; Zhang, R.; Zhang, L.; Sun, Y.; Xie, Q. Effects of Different Microstructured Surfaces on the Osseointegration of CAD/CAM Zirconia Dental Implants: An Experimental Study in Rabbits. Int. J. Oral Maxillofac. Implant. 2020, 35, 1113–1121. [Google Scholar] [CrossRef]
  298. Lee, J.W.; Wen, H.B.; Gubbi, P.; Romanos, G.E. New bone formation and trabecular bone microarchitecture of highly porous tantalum compared to titanium implant threads: A pilot canine study. Clin. Oral. Implant. Res. 2018, 29, 164–174. [Google Scholar] [CrossRef]
  299. Hariharan, K.; Arumaikkannu, G. Influence of Hydroxyapatite coated Additive Manufactured polyamide substrate on Biocompatibility. Biomed. Res. 2015, 26, S15–S21. [Google Scholar]
  300. Chen, J.; Zhang, Z.; Chen, X.; Zhang, C.; Zhang, G.; Xu, Z. Design and manufacture of customized dental implants by using reverse engineering and selective laser melting technology. J. Prosthet. Dent. 2014, 112, 1088–1095. [Google Scholar] [CrossRef]
  301. Revilla-León, M.; Ceballos, L.; Martínez-Klemm, I.; Özcan, M. Discrepancy of complete-arch titanium frameworks manufactured using selective laser melting and electron beam melting additive manufacturing technologies. J. Prosthet. Dent. 2018, 120, 942–947. [Google Scholar] [CrossRef] [PubMed]
  302. Grigoriev, S.; Peretyagin, N.; Apelfeld, A.; Smirnov, A.; Rybkina, A.; Kameneva, E.; Zheltukhin, A.; Gerasimov, M.; Volosova, M.; Yanushevich, O.; et al. Investigation of the Characteristics of MAO Coatings Formed on Ti6Al4V Titanium Alloy in Electrolytes with Graphene Oxide Additives. J. Compos. Sci. 2023, 7, 142. [Google Scholar] [CrossRef]
  303. Solís Pinargote, N.W.; Smirnov, A.; Peretyagin, N.; Seleznev, A.; Peretyagin, P. Direct Ink Writing Technology (3D Printing) of Graphene-Based Ceramic Nanocomposites: A Review. Nanomaterials 2020, 10, 1300. [Google Scholar] [CrossRef] [PubMed]
  304. Grigoriev, S.; Smirnov, A.; Solis Pinargote, N.W.; Yanushevich, O.; Kriheli, N.; Kramar, O.; Pristinskiy, Y.; Peretyagin, P. Evaluation of Mechanical and Electrical Performance of Aging Resistance ZTA Composites Reinforced with Graphene Oxide Consolidated by SPS. Materials 2022, 15, 2419. [Google Scholar] [CrossRef]
  305. Smirnov, A.; Peretyagin, P.; Bartolomé, J.F. Processing and mechanical properties of new hierarchical metal-graphene flakes reinforced ceramic matrix composites. J. Eur. Ceram. Soc. 2019, 39, 3491–3497. [Google Scholar] [CrossRef]
  306. Peretyagin, N.Y.; Suminov, I.V.; Smirnov, A.; Krikheli, N.I.; Kramar, O.V.; Peretyagin, P.Y. Electrochemical deposition of ceramic-like coatings on Ti-6Al-4V parts fabricated by electron beam melting. J. Phys. Conf. Ser. 2021, 2144, 012017. [Google Scholar] [CrossRef]
  307. Smirnov, A.; Volosova, M.; Peretyagin, P.; Bartolomé, J.F. Tribological behaviour of a 3Y-TZP/Ta ceramic-metal biocomposite against ultrahigh molecular weight polyethylene (UHMWPE). Ceram. Int. 2018, 44, 1404–1410. [Google Scholar] [CrossRef]
  308. Guarnieri, R.; Serra, M.; Bava, L.; Grande, M.; Farronato, D.; Iorio-Siciliano, V. The impact of a laser-microtextured collar on crestal bone level and clinical parameters under various placement and loading protocols. Int. J. Oral Maxillofac. Implant. 2014, 29, 354–363. [Google Scholar] [CrossRef]
  309. Mastrangelo, F.; Quaresima, R.; Canullo, L.; Scarano, A.; Muzio, L.L.; Piattelli, A. Effects of novel laser dental implant microtopography on human osteoblast proliferation and bone deposition. Int. J. Oral Maxillofac. Implant. 2020, 35, 320–329. [Google Scholar] [CrossRef]
  310. Ferraris, S.; Warchomicka, F.; Iranshahi, F.; Rimondini, L.; Cochis, A.; Spriano, S. Electron beam structuring of Ti6Al4V: New insights on the metal surface properties influencing the bacterial adhesion. Materials 2020, 13, 409. [Google Scholar] [CrossRef]
  311. Hu, F.; Fan, X.; Peng, F.; Yan, X.; Song, J.; Deng, C.; Liu, M.; Zeng, D.; Ning, C. Characterization of Porous Titanium-Hydroxyapatite Composite Biological Coating on Polyetheretherketone (PEEK) by Vacuum Plasma Spraying. Coatings 2022, 12, 433. [Google Scholar] [CrossRef]
  312. Brunello, G.; Brun, P.; Gardin, C.; Ferroni, L.; Bressan, E.; Meneghello, R.; Zavan, B.; Sivolella, S. Biocompatibility and antibacterial properties of zirconium nitride coating on titanium abutments: An in vitro study. PLoS ONE 2018, 13, e0199591. [Google Scholar] [CrossRef] [PubMed]
  313. Prachar, P.; Bartakova, S.; Brezina, V.; Cvrcek, L.; Vanek, J. Cytocompatibility of implants coated with titanium nitride and zirconium nitride. Bratisl. Med. J. 2015, 116, 154–156. [Google Scholar] [CrossRef]
  314. Grigoriev, S.N.; Volosova, M.A.; Migranov, M.S.; Minin, I.V.; Shekhtman, S.R.; Suhova, N.A.; Gurin, V.D.; Pivkin, P.M. Nanostructured biocompatible Ti-TiN coating for implants with improved functional properties. Proc. SPIE 2021, 11867, 1186708. [Google Scholar]
  315. Tao, H.; Zhylinski, V.; Vereschaka, A.; Chayeuski, V.; Yuanming, H.; Milovich, F.; Sotova, C.; Seleznev, A.; Salychits, O. Comparison of the Mechanical Properties and Corrosion Resistance of the Cr-CrN, Ti-TiN, Zr-ZrN, and Mo-MoN Coatings. Coatings 2023, 13, 750. [Google Scholar] [CrossRef]
  316. Vereschaka, A.; Grigoriev, S.; Chigarev, A.; Milovich, F.; Sitnikov, N.; Andreev, N.; Sotova, C.; Bublikov, J. Development of a Model of Crack Propagation in Multilayer Hard Coatings under Conditions of Stochastic Force Impact. Materials 2021, 14, 260. [Google Scholar] [CrossRef]
  317. Grigoriev, S.; Vereschaka, A.; Milovich, F.; Tabakov, V.; Sitnikov, N.; Andreev, N.; Sviridova, T.; Bublikov, J. Investigation of multicomponent nanolayer coatings based on nitrides of Cr, Mo, Zr, Nb, and Al. Surf. Coat. Technol. 2020, 401, 126258. [Google Scholar] [CrossRef]
  318. Alves, C.F.A.; Calderón, S.V.; Dias, D.; Carvalho, S. Influence of Oxygen content on the electrochemical behavior of Ta1-xOx coatings. Electrochim. Acta 2016, 211, 385–394. [Google Scholar] [CrossRef]
  319. Liu, F.; Li, Y.; Liang, J.; Sui, W.; Bellare, A.; Kong, L. Effects of micro/nano strontium-loaded surface implants on osseointegration in ovariectomized sheep. Clin. Implant Dent. Relat. Res. 2019, 21, 377–385. [Google Scholar] [CrossRef]
  320. Yokota, S.; Nishiwaki, N.; Ueda, K.; Narushima, T.; Kawamura, H.; Takahashi, T. Evaluation of thin amorphous calcium phosphate coatings on titanium dental implants deposited using magnetron sputtering. Implant Dent. 2014, 23, 343–350. [Google Scholar] [CrossRef]
  321. Pan, Y.-H.; Yi Lin, J.C.; Chen, M.K.; Salamanca, E.; Choy, C.S.; Tsai, P.-Y.; Leu, S.-J.; Yang, K.-C.; Huang, H.-M.; Yao, W.-L.; et al. Glow discharge plasma treatment on zirconia surface to enhance osteoblastic-like cell differentiation and antimicrobial effects. Materials 2020, 13, 3771. [Google Scholar] [CrossRef] [PubMed]
  322. Khandaker, M.; Riahinezhad, S.; Li, Y.; Vaughan, M.B.; Sultana, F.; Morris, T.L.; Phinney, L.; Hossain, K. Plasma nitriding of titanium alloy: Effect of roughness, hardness, biocompatibility, and bonding with bone cement. Bio-Med. Mater. Eng. 2016, 27, 461–474. [Google Scholar] [CrossRef]
  323. Canullo, L.; Genova, T.; Wang, H.-L.; Carossa, S.; Mussano, F. Plasma of argon increases cell attachment and bacterial decontamination on different implant surfaces. Int. J. Oral Maxillofac. Implant. 2017, 32, 1315–1323. [Google Scholar] [CrossRef]
  324. Ogawa, T. Ultraviolet Photofunctionalization of Titanium Implant. Int. J. Oral Maxillofac. Implant. 2014, 29, 95–102. [Google Scholar] [CrossRef]
  325. Kim, H.S.; Lee, J.I.; Yang, S.S.; Kim, B.S.; Kim, B.C.; Lee, J. The effect of alendronate soaking and ultraviolet treatment on bone–implant interface. Clin. Oral. Implant. Res. 2017, 28, 1164–1172. [Google Scholar] [CrossRef]
  326. Jin, G.; Qin, H.; Cao, H.; Qian, S.; Zhao, Y.; Peng, X.; Zhang, X.; Liu, X.; Chu, P.K. Synergistic effects of dual Zn/Ag ion implantation in osteogenic activity and antibacterial ability of titanium. Biomaterials 2014, 35, 7699–7713. [Google Scholar] [CrossRef] [PubMed]
  327. Mei, S.; Wang, H.; Wang, W.; Tong, L.; Pan, H.; Ruan, C.; Ma, Q.; Liu, M.; Yang, H.; Zhang, L.; et al. Antibacterial effects and biocompatibility of titanium surfaces with graded silver incorporation in titania nanotubes. Biomaterials 2014, 35, 4255–4265. [Google Scholar] [CrossRef] [PubMed]
  328. Wang, X.; Lu, T.; Wen, J.; Xu, L.; Zeng, D.; Wu, Q.; Cao, L.; Lin, S.; Liu, X.; Jiang, X. Selective responses of human gingival fibroblasts and bacteria on carbon fiber reinforced polyetheretherketone with multilevel nanostructured TiO2. Biomaterials 2016, 83, 207–218. [Google Scholar] [CrossRef] [PubMed]
  329. Cao, H.; Meng, F.; Liu, X. Antimicrobial activity of tantalum oxide coatings decorated with Ag nanoparticles. J. Vac. Sci. Technol. A 2016, 34, 04C102. [Google Scholar] [CrossRef]
  330. Zhao, X.; Yang, J.; You, J. Surface modification of TiO2 coatings by Zn ion implantation for improving antibacterial activities. Bull. Mater. Sci. 2016, 39, 285–291. [Google Scholar] [CrossRef]
  331. Benalcázar Jalkh, E.B.; Parra, M.; Torroni, A.; Nayak, V.V.; Tovar, N.; Castellano, A.; Badalov, R.M.; Bonfante, E.A.; Coelho, P.G.; Witek, L. Effect of supplemental acid-etching on the early stages of osseointegration: A preclinical model. J. Mech. Behav. Biomed. 2021, 122, 104682. [Google Scholar] [CrossRef] [PubMed]
  332. Cavalcanti, Y.W.; Soare, R.V.; Leite Assis, M.A.; Zenóbio, E.G.; Girundi, F.M. Titanium Surface Roughing Treatments contribute to Higher Interaction with Salivary Proteins MG2 and Lactoferrin. J. Contemp. Dent. Pract. 2015, 16, 141–146. [Google Scholar] [CrossRef]
  333. Ren, B.; Wan, Y.; Liu, C.; Wang, H.; Yu, M.; Zhang, X.; Huang, Y. Improved osseointegration of 3D printed Ti-6Al-4V implant with a hierarchical micro/nano surface topography: An in vitro and in vivo study. Mater. Sci. Eng. C Mater. Biol. Appl. 2021, 118, 111505. [Google Scholar] [CrossRef] [PubMed]
  334. Al Mustafa, M.; Agis, H.; Müller, H.-D.; Watzek, G.; Gruber, R. In vitro adhesion of fibroblastic cells to titanium alloy discs treated with sodium hydroxide. Clin. Oral. Implant. Res. 2015, 26, 15–19. [Google Scholar] [CrossRef] [PubMed]
  335. Park, S.; Lee, S.; Song, I.-H.; Kim, S. Water glass coating on a Ti substrate to form Si-OH groups for improving cell behaviors of dental implants. Korean J. Met. Mater. 2014, 52, 123–128. [Google Scholar] [CrossRef]
  336. Lee, K.; Shin, G. RF magnetron sputtering coating of hydroxyapatite on alkali solution treated titanate nanorods. Arch. Metall. Mater. 2015, 60, 1319–1322. [Google Scholar] [CrossRef]
  337. Karl, M.; Albrektsson, T. Clinical performance of dental implants with a moderately rough (TiUnite) surface: A meta-analysis of prospective clinical studies. Int. J. Oral Maxillofac. Implant. 2017, 32, 717–734. [Google Scholar] [CrossRef]
  338. Martínez-Rus, F.; Prieto, M.; Salido, M.P.; Madrigal, C.; Özcan, M.; Pradíes, G. A clinical study assessing the influence of anodized titanium and zirconium dioxide abutments and periimplant soft tissue thickness on the optical outcome of implant-supported lithium disilicate single crowns. Int. J. Oral Maxillofac. Implant. 2017, 32, 156–163. [Google Scholar] [CrossRef]
  339. Kasatkin, V.E.; Kasatkina, I.V.; Bogdashkina, N.L.; Gerasimov, M.V.; Krit, B.L.; Grigoriev, S.N.; Suminov, I.V.; Kozlov, I.A. Influence of different modes of microarc oxidation of titanium on the electrochemical properties and surface morphology of the obtained coatings. Surf. Eng. 2023, 39, 295–306. [Google Scholar] [CrossRef]
  340. Djošić, M.; Janković, A.; Mišković-Stanković, V. Electrophoretic deposition of biocompatible and bioactive hydroxyapatite-based coatings on titanium. Materials 2021, 14, 5391. [Google Scholar] [CrossRef]
  341. Prem Ananth, K.; Joseph Nathanael, A.; Jose, S.P.; Oh, T.H.; Mangalaraj, D. A novel silica nanotube reinforced ionic incorporated hydroxyapatite composite coating on polypyrrole coated 316L SS for implant application. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 59, 1110–1124. [Google Scholar] [CrossRef] [PubMed]
  342. Mehboob, H.; Awais, M.; Khalid, H.; Ch, A.A.; Siddiqi, S.A.; Rehman, I. Polymer-assisted deposition of hydroxyapatite coatings using electrophoretic technique. Biomed. Eng. Appl. Basis Commun. 2014, 26, 1450073. [Google Scholar] [CrossRef]
  343. Zhenga, C.Y.; Niea, F.L.; Zhenga, Y.F.; Cheng, Y.; Wei, S.C.; Ruan, L.; Valiev, R.Z. Enhanced corrosion resistance and cellular behavior of ultrafine-grained bio-medical NiTi alloy with a novel SrO-SiO2-TiO2 sol-gel coating. Appl. Surf. Sci. 2011, 257, 5913–5918. [Google Scholar] [CrossRef]
  344. Mohammad, N.F.; Ahmad, R.N.; Mohd Rosli, N.L.; Abdul Manan, M.S.; Marzuki, M.; Wahi, A. Sol gel deposited hydroxyap-atite-based coating technique on porous titanium niobium for biomedical applications: A mini review. Mater. Today 2021, 41, 127–135. [Google Scholar]
  345. Radin, S.; Ducheyne, P. Controlled release of vancomycin from thin sol–gel films on titanium alloy fracture plate material. Biomaterials 2007, 28, 1721–1729. [Google Scholar] [CrossRef] [PubMed]
  346. Alavi, S.E.; Panah, N.; Page, F.; Gholami, M.; Dastfal, A.; Sharma, L.A.; Shahmabadi, H.E. Hydrogel-based therapeutic coatings for dental implants. Eur. Polym. J. 2022, 181, 111652. [Google Scholar] [CrossRef]
  347. Ali, M.; Ali, F.; Yang, B.; Abbas, A. A comprehensive account of biomedical applications of CVD diamond coatings. J. Phys. D Appl. Phys. 2021, 54, 443001. [Google Scholar] [CrossRef]
  348. Gu, M.; Lv, L.; Du, F.; Niu, T.; Chen, T.; Xia, D.; Wang, S.; Zhao, X.; Liu, J.; Liu, Y.; et al. Effects of thermal treatment on the adhesion strength and osteoinductive activity of single-layer graphene sheets on titanium substrates. Sci. Rep. 2018, 8, 8141. [Google Scholar] [CrossRef]
  349. Strąkowska, P.; Beutner, R.; Gnyba, M.; Zielinski, A.; Scharnweber, D. Electrochemically assisted deposition of hydroxyapatite on Ti6Al4V substrates covered by CVD diamond films—Coating characterization and first cell biological results. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 59, 624–635. [Google Scholar] [CrossRef]
  350. Shubladze, G.K. Comparison of implant anodized surface advantage with other types of surfaces. Med. Alph. 2014, 3, 20–25. (In Russian) [Google Scholar]
  351. Rajan, A.; Sivarajan, S.; Vallabhan, C.G.; Nair, A.S.; Jayakumar, S.; Pillai, A.S. An in vitro study to evaluate and compare the hemocompatibility of titanium and zirconia implant materials after sandblasted and acid etched surface treatment. J. Contemp. Dent. Pract. 2018, 19, 1449–1455. [Google Scholar]
  352. Kim, H.-C.; Park, S.-Y.; Han, M.-S.; Lee, Y.-M.; Ku, Y.; Rhyu, I.-C.; Seol, Y.-J. Occurrence of progressive bone loss around anodized surface implants and resorbable blasting media implants: A retrospective cohort study. J. Periodontol. 2017, 88, 329–337. [Google Scholar] [CrossRef] [PubMed]
  353. Cervino, G.; Fiorillo, L.; Iannello, G.; Santonocito, D.; Risitano, G.; Cicciù, M. Sandblasted and acid etched titanium dental implant surfaces systematic review and confocal microscopy evaluation. Materials 2019, 12, 1763. [Google Scholar] [CrossRef] [PubMed]
  354. Gnedenkov, S.V.; Sharkeev, Y.P.; Sinebryukhov, S.L.; Khrisanfova, O.A.; Legostaeva, E.V.; Zavidnaya, A.G.; Puz’, A.V.; Khlusov, I.A.; Opra, D.P. Functional coatings formed on the titanium and magnesium alloys as implant materials by plasma electrolytic oxidation technology: Fundamental principles and synthesis conditions. Corros. Rev. 2016, 34, 65–83. [Google Scholar] [CrossRef]
  355. Momesso, G.A.C.; de Souza Santos, A.M.; Fonseca e Santos, J.M.; da Cruz, N.C.; Okamoto, R.; Ricardo Barão, V.A.; Siroma, R.S.; Shibli, J.A.; Faverani, L.P. Comparison between plasma electrolytic oxidation coating and sandblasted acid-etched surface treatment: Histometric, tomographic, and expression levels of osteoclastogenic factors in osteoporotic rats. Materials 2020, 13, 1604. [Google Scholar] [CrossRef] [PubMed]
  356. Apachiteia, I.; Leonia, A.; Riemslagb, A.C.; Fratila-Apachitei, L.E.; Duszczyk, J. Enhanced fatigue performance of porous coated Ti6Al4V biomedical alloy. Appl. Surf. Sci. 2011, 257, 6941–6944. [Google Scholar] [CrossRef]
  357. Apelfeld, A.; Grigoriev, S.; Krit, B.; Ludin, V.; Suminov, I.; Chudinov, D. Improving the stability of the coating properties for group plasma electrolytic oxidation. Manuf. Lett. 2022, 33, 54–59. [Google Scholar] [CrossRef]
  358. Kusmanov, S.; Mukhacheva, T.; Tambovskiy, I.; Kusmanova, I.; Shadrin, S.; Belov, R.; Nikiforov, R.; Suminov, I.; Karasev, M.; Grigoriev, S. Possibilities of Duplex Plasma Electrolytic Treatment for Increasing the Hardness and Wear Resistance of a Commercially Pure Titanium Surface. Coatings 2023, 13, 1363. [Google Scholar] [CrossRef]
  359. Krit, B.L.; Apelfeld, A.V.; Borisov, A.M.; Morozova, N.V.; Rakoch, A.G.; Suminov, I.V.; Grigoriev, S.N. Plasma Electrolytic Modification of Zirconium and Its Alloys: Brief Review. Materials 2023, 16, 5543. [Google Scholar] [CrossRef]
  360. Grigoriev, S.; Peretyagin, N.; Apelfeld, A.; Smirnov, A.; Morozov, A.; Torskaya, E.; Volosova, M.; Yanushevich, O.; Yarygin, N.; Krikheli, N.; et al. Investigation of Tribological Characteristics of PEO Coatings Formed on Ti6Al4V Titanium Alloy in Electrolytes with Graphene Oxide Additives. Materials 2023, 16, 3928. [Google Scholar] [CrossRef]
  361. Tverdokhlebov, S.I.; Shesterikov, E.V.; Malchikhina, A.I. Formation hybrid method of multilayer biocoatings based on PVD technology. Adv. Mater. Res. 2014, 872, 241–247. [Google Scholar] [CrossRef]
  362. Jiang, H.; Zhang, T.; Zhou, W.; Lin, Z.; Liu, Z. Effect of plasma oxidation-treated TiOx film on early osseointegration. Int. J. Oral Maxillofac. Implant. 2018, 33, 1011–1018. [Google Scholar] [CrossRef] [PubMed]
  363. Velten, M.F.; Odatsu, T.; Aswath, P.B.; Kamiya, N.; Kim, H.; Varanasi, V.G. PeCVD SiOx accelerates hydroxyapatite surface formation for enhanced early osteogenic differentiation. Ceram. Trans. 2014, 251, 105–113. [Google Scholar]
  364. Novikov, S.V.; Tamazov, I.D.; Matveev, A.I.; Topoljanskij, P.A.; Topoljanskij, A.P. Optimization of the surface of titanium dental implants of grade 5 alloy by barrier glass ceramic coating. Clin. Dent. 2021, 24, 29–36. [Google Scholar] [CrossRef]
  365. Gao, F.; Li, G.; Xia, Y. Influence of hysteresis effect on properties of reactively sputtered TiAlSiN films. Appl. Surf. Sci. 2018, 431, 160–164. [Google Scholar] [CrossRef]
  366. Momozawa, A.; Katsui, H.; Ito, A.; Goto, T. Dental application of TiO2 films prepared by laser CVD. J. Jpn. Soc. Powder Metall. 2016, 63, 128–131. [Google Scholar] [CrossRef]
  367. Pou, J.; Lusquiños, F.; Comesaña, R.; Boutinguiza, M. Production of biomaterial coatings by laser-assisted processes. In Advances in Laser Materials Processing: Technology, Research and Application; Lawrence, J., Pou, J., Low, D.K.Y., Toyserkani, E., Eds.; Woodhead Publishing: Cambridge, UK, 2010; pp. 394–425. [Google Scholar]
  368. Al-Moraissi, E.A.; Oginni, F.O.; Holkom, M.A.M.; Mohamed, A.A.S.; Al-Sharani, H.M. Tissue-Engineered bone using mesenchymal stem cells versus conventional bone grafts in the regeneration of maxillary alveolar bone: A systematic review and meta-analysis. Int. J. Oral Maxillofac. Implant. 2020, 35, 79–90. [Google Scholar] [CrossRef] [PubMed]
  369. Sánchez-Garcés, M.; Alvira-González, J.; Sánchez, C.M.; Cairó, J.R.B.; del Pozo, M.R.; Gay-Escoda, C. Bone regeneration using adipose-derived stem cells with fibronectin in dehiscence-type defects associated with dental implants: An experimental study in a dog model. Int. J. Oral Maxillofac. Implant. 2017, 32, e97–e106. [Google Scholar] [CrossRef]
  370. Zanicotti, D.G.; Duncan, W.J.; Seymour, G.J.; Coates, D.E. Effect of titanium surfaces on the osteogenic differentiation of human adipose-derived stem cells. Int. J. Oral Maxillofac. Implant. 2018, 33, e77–e87. [Google Scholar] [CrossRef]
  371. Yoo, S.-Y.; Kim, S.-K.; Heo, S.-J.; Koak, J.-Y.; Lee, J.-H.; Park, J.-M. Biochemical responses of anodized titanium implants with a poly(lactide-co-glycolide)/bone morphogenic protein-2 submicron particle coating. part 1: An in vitro study. Int. J. Oral Maxillofac. Implant. 2015, 30, 512–518. [Google Scholar] [CrossRef]
  372. Knopf-Marques, H.; Barthes, J.; Lachaal, S.; Mutschler, A.; Muller, C.; Dufour, F.; Rabineau, M.; Courtial, E.-J.; Bystroňová, J.; Marquette, C.; et al. Multifunctional polymeric implant coatings based on gelatin, hyaluronic acid derivative and chain length-controlled poly(arginine). Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 104, 109898. [Google Scholar] [CrossRef] [PubMed]
  373. Horasawa, N.; Yamashita, T.; Uehara, S.; Udagawa, N. High-performance scaffolds on titanium surfaces: Osteoblast differentiation and mineralization promoted by a globular fibrinogen layer through cell-autonomous BMP signaling. Mater. Sci. Eng. C Mater. Biol. Appl. 2015, 46, 86–96. [Google Scholar] [CrossRef]
  374. de Oliveira Fernandes, G.V.; Santos, N.B.M.; de Sousa, M.F.C.; Fernandes, J.C.H. Liquid Platelet-Rich Fibrin Coating Implant Surface to Enhance Osseointegration: A Double-Blinded, Randomized Split-Mouth Trial with 1-Year Follow-up. Int. J. Oral Maxillofac. Implant. 2022, 37, 159–170. [Google Scholar] [CrossRef]
  375. Öncü, E.; Alaaddinoğlu, E.E. The effect of platelet-rich fibrin on implant stability. Int. J. Oral Maxillofac. Implant. 2015, 30, 578–582. [Google Scholar] [CrossRef]
  376. Shah, R.; Thomas, R.; Gowda, T.M.; Baron, T.K.A.; Vemanaradhya, G.G.; Bhagat, S. In Vitro Evaluation of Osteoblast Response to the Effect of Injectable Platelet-rich Fibrin Coating on Titanium Disks. J. Contemp. Dent. Pract. 2021, 22, 107–110. [Google Scholar] [CrossRef] [PubMed]
  377. Kulakov, A.A.; Kasparov, A.S.; Porfenchuk, D.A. Factors affecting osseointegration and the use of early functional loading to reduce the treatment time for dental implantation. Stomatologiya 2019, 4, 107–115. [Google Scholar]
  378. Nakazawa, M.; Yamada, M.; Wakamura, M.; Egusa, H.; Sakurai, K. Activation of osteoblastic function on titanium surface with titanium-doped hydroxyapatite nanoparticle coating: An in vitro study. Int. J. Oral Maxillofac. Implant. 2017, 32, 779–791. [Google Scholar] [CrossRef]
  379. Li, X.; Li, Y.; Liao, Y.; Li, J.; Zhang, L.; Hu, J. The Effect of Magnesium-Incorporated Hydroxyapatite Coating on Titanium Implant Fixation in Ovariectomized Rats. Int. J. Oral Maxillofac. Implant. 2014, 29, 196–202. [Google Scholar] [CrossRef]
  380. Martinez, E.F.; Ishikawa, G.J.; de Lemos, A.B.; Bezerra, F.J.B.; Sperandio, M.; Napimoga, M.H. Evaluation of a Titanium surface treated with hydroxyapatite nanocrystals on osteoblastic cell behavior: An in vitro study. Int. J. Oral Maxillofac. Implant. 2018, 33, 597–602. [Google Scholar] [CrossRef]
  381. Thiem, D.G.E.; Adam, M.; Ganz, C.; Gerber, T.; Kämmerer, P.W. The implant surface and its role in affecting the dynamic processes of bone remodeling by means of distance osteogenesis: A comparative in vivo study. Int. J. Oral Maxillofac. Implant. 2019, 34, 133–140. [Google Scholar] [CrossRef]
  382. Thoma, D.S.; Jung, U.-W.; Park, J.-Y.; Bienz, S.P.; Hüsler, J.; Jung, R.E. Bone augmentation at peri-implant dehiscence defects comparing a synthetic polyethylene glycol hydrogel matrix vs. standard guided bone regeneration techniques. Clin. Oral. Implant. Res. 2017, 28, e76–e83. [Google Scholar] [CrossRef] [PubMed]
  383. Fadeeva, I.V.; Volchenkova, V.A.; Fomina, A.A.; Mamin, G.V.; Shurtakova, D.V.; Kalita, V.I. Plasma-sprayed manganese-containing tricalcium phosphate coatings on titanium. Inorg. Mater. 2021, 57, 967–972. [Google Scholar] [CrossRef]
  384. Mayborodin, I.V.; Toder, M.S.; Shevela, A.I.; Razumakhina, M.S.; Shevela, A.A.; Patrushev, A.Y.; Ragimova, T.M.; Kuznetsova, I.V. Histological results of implantation of metal products with rough and smooth surfaces into bone tissue in an experiment. Fund Res. 2014, 7, 114–118. [Google Scholar]
  385. Gadallah, M.A.; Darwish, R.M.; Alshimy, A.M.; Gepreel, M.A.; Marei, M.K. Biomimetic Coating of Titanium Surface Using Bioactive Glass Nanoparticles. Int. J. Oral Maxillofac. Implant. 2022, 37, 86–97. [Google Scholar] [CrossRef] [PubMed]
  386. Herrero-Climent, M.; Romero Ruiz, M.M.; Calvo, P.L.; Santos, J.V.R.; Perez, R.A.; Gil Mur, F.J. Effectiveness of a new dental implant bioactive surface: Histological and histomorphometric comparative study in minipigs. Clin. Oral. Investig. 2018, 22, 1423–1432. [Google Scholar] [CrossRef]
  387. Janson, O.; Gururaj, S.; Pujari-Palmer, S.; Karlsson Ott, M.; Strømme, M.; Engqvist, H.; Welch, K. Titanium surface modification to enhance antibacterial and bioactive properties while retaining biocompatibility. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 96, 272–279. [Google Scholar] [CrossRef]
  388. Hao, J.; Chou, J.; Kuroda, S.; Otsuka, M.; Kasugai, S.; Lang, N.P. Strontium hydroxyapatite in situ gel-forming system—A new approach for minimally invasive bone augmentation. Clin. Oral. Implant. Res. 2015, 26, 581–585. [Google Scholar] [CrossRef]
  389. Badiadka, N. Studies on characteristics and corrosion behaviour of chitosan/eudragit RS100 bilayer film coated Ti-6Al-4V. Indian J. Chem. Technol. 2023, 30, 570–578. [Google Scholar]
  390. Fontelo, R.; da Costa, D.S.; Gomez-Florit, M.; Tiainen, H.; Reis, R.L.; Novoa-Carballal, R.; Pashkuleva, I. Antibacterial nanopatterned coatings for dental implants. J. Mater. Chem. B 2022, 10, 8710–8718. [Google Scholar] [CrossRef]
Figure 1. Analysis of search results for publications using the keywords “dental implant” from 1989, the year when the first patent for a titanium implant was registered, up to the present (according Scopus, ScienceDirect).
Figure 1. Analysis of search results for publications using the keywords “dental implant” from 1989, the year when the first patent for a titanium implant was registered, up to the present (according Scopus, ScienceDirect).
Materials 16 07383 g001
Figure 4. Number of scientific publications in the search results of publications for the keywords (dental implant*) from 2000 to the present (according to ScienceDirect).
Figure 4. Number of scientific publications in the search results of publications for the keywords (dental implant*) from 2000 to the present (according to ScienceDirect).
Materials 16 07383 g004
Figure 5. Dental prosthesis: 1—implant, 2—abutment, 3—crown.
Figure 5. Dental prosthesis: 1—implant, 2—abutment, 3—crown.
Materials 16 07383 g005
Figure 6. Number of scientific publications in the search results of publications for the keywords (dental implant*) from 2000 to the present (according to ScienceDirect). The number of publications (%) is presented relative to their total number for all groups of materials (SS+Ti+Zr+Ta).
Figure 6. Number of scientific publications in the search results of publications for the keywords (dental implant*) from 2000 to the present (according to ScienceDirect). The number of publications (%) is presented relative to their total number for all groups of materials (SS+Ti+Zr+Ta).
Materials 16 07383 g006
Figure 7. Hardness and elastic modulus of Ti-Mo and Ti-Mo-Zr alloys according to data [123].
Figure 7. Hardness and elastic modulus of Ti-Mo and Ti-Mo-Zr alloys according to data [123].
Materials 16 07383 g007
Table 1. Titanium alloys used for the production of dental implants.
Table 1. Titanium alloys used for the production of dental implants.
Double AlloysTriple AlloysMore Complex Alloys
Ti-AlTi-Al-V [62,63,64,65,66,67]
Ti-Al-Nb [77,78,79,80]Ti-Al-Nb + YSZ (ZrO2, stabilised Y) [81]
Ti-Nb [53,82,83,84]Ti-Nb-Zr (TNZ alloys) [76,85,86,87]Ti-Nb-Zr-Ta [88,89]
Ti-Nb-Zr-Ta-Si [90]
Ti-Nb-Ta [91]
Ti-Nb-Cu [92,93]Ti-Nb-Cu-Ga [93]
Ti-Nb-Ga [93,94,95]
Ti-Nb-Mo [75,96]Ti-Nb-Mo-Zr-Sn [97]
Ti-Nb-Mo-Ta [98]
Ti-Zr [68,69,70,71,99,100,101,102,103]Ti-Zr-Nb [104,105]Ti-Zr-Nb-Ta [104]
Ti-Zr-Fe [106]Ti-Zr-Fe-Si [107]
Ti-Ta [108,109,110,111,112]Ti-Ta-Zr [113,114]Ti-Ta-Zr-Nb [115]
Ti-Ta-Zr-Mo [116]
Ti-Cu [117,118,119]Ti-Cu-Al [120]
Ti-Cu-Nb [121]
Ti-Mo [122,123,124]Ti-Mo-Zr [123]Ti-Mo-Zr-Fe [125]
Ti-Mg [126,127,128,129]Ti-Mg-Sr [130]
Ti-Fe [131]Te-Fe-Ge [132]
Ti-Ca [133]
Ti-Pd [133,134,135]
Ti-Pt [133]
Table 2. Mechanical properties of basic materials used for dental implants [26,93,101,154,161,205,206,219,261,262,263,264,265,266,267,268,269,270,271,272,273].
Table 2. Mechanical properties of basic materials used for dental implants [26,93,101,154,161,205,206,219,261,262,263,264,265,266,267,268,269,270,271,272,273].
MaterialCompressive Strength, MPaTensile Strength, MPaModulus of Elasticity, GPaHardness HV, GPa
ZrO21300–2000150190–21010–13
Al2O32100–2500260221–40012.5–21
Si3N42200–3000<350210–30015–22
PEEK118–240100–2303.6–3.80.7
PEEK with carbon fibres280–300250–26014–180.8
cp-Ti235–353860100–1051.8–2.2
Ti-6Al-4V990–1565689110–1153.1–3.5
Table 3. Methods of surface modification of dental implants.
Table 3. Methods of surface modification of dental implants.
GroupMethodFeatures (Characteristics) of the Method
Mechanical treatmentMachining through cutting (machine stroking) [7,287,288,289,290,291,292]It is performed in order to increase the surface roughness in order to increase the osteointegration of the implant.
This method is inefficient, as the rough surface obtained by this method is much inferior to the porous one as per this indicator.
The main materials of implants belong to the class of hard-to-machine materials.
Sandblasting
[7,293,294,295,296,297]
A simple and inexpensive method.
Cell adhesion, proliferation, and the differentiation of osteoblasts are improved.
However, it is often necessary to carry out acid etching of the implant to homogenize the surface microprofile in order to remove the remaining sand particles.
Otherwise, the inhomogeneity of the surface material reduces the corrosion resistance of the implant.
Physical TechniquesAdditive technologies [25,285,298,299,300,301,302,303,304,305,306,307]It allows to create complex three-dimensional structures (e.g., trabecular, heroic), which increase the osteoconductivity of implants.
Compared to traditional methods of manufacturing complex profile products, it is possible to save resources (materials, energy, labor costs).
Laser ablation (including laser surface texturing) [288,290,308,309]Compared to machining, surface texturing reduces the operating costs of the production process (no consumables (e.g., cutting tools)).
Provides an effective way to clean the surface of the product without the use of any chemicals.
Ability to automate the texturing process with high precision.
Allows the creation of micro-caverns—cavities with a diameter of a few micrometers—with an orderly programmed structure.
Increases the osseointegration of the implant due to the creation of the precise “branched” macrostructure (topography) of the surface.
Highly energetic and expensive processing method.
Electron beam (EB) structuring/surface texturing
[310]
It differs from laser texturing in that instead of a laser beam, a sharply focused beam of electrons moving at high speed is used for technological purposes.
EL texturing shows a statistically significant reduction in bacterial adhesion in the absence of antibacterial agents (even after implant polishing).
High-energy processing method.
Plasma spraying [7,253,311]Economical and safe method.
Possibility of applying a wide variety of metal, ceramic, or plastic coatings at atmospheric pressure.
Possibility of obtaining composite bioactive substances—in particular, osseointegration enhancing and coatings (e.g., porous hydroxyapatite reinforced with titanium).
Vacuum arc coating (cathodic arc deposition; physical coating deposition (PVD)) [292,299,312,313,314,315,316,317]The formation of coatings is carried out in vacuum.
It is used for the deposition of metal, ceramic, and composite coatings on substrates made of various materials (including low-heat-resistant (up to 200 °C)).
To improve the quality of coatings, equipment and technologies are used for the filtration of the microparticle phase from the plasma stream, ion bombardment and etching of the surface are used to clean and thermoactivate it, and ion implantation is used to increase adhesion and modify the outer layer of the substrate [a number of our publications].
By varying both the composition of cathodes and reaction gases as well as coating modes, it is possible to obtain multilayer functional coatings, thereby improving the operational properties of products—in particular, increasing the corrosion resistance of implants and giving the product antibacterial properties.
Magnetron sputtering method [318,319,320]High speed of coating formation at low vacuum of 0.1–1 Pa; no overheating of the substrate.
Magnetron sputtering can be carried out both on direct and alternating currents.
It is used for the formation of protective coatings in order to increase the corrosion resistance and osteointegration of implants.
The coatings formed on the implants are, as a rule, single-layer and simple in composition films (oxide, carbide, nitride, or metallic).
Complex coatings (e.g., calcium phosphate) can be deposited wherein the composition of each coating corresponds to the composition of the target atomized in argon.
It is possible to combine the deposition of coatings with preliminary plasma-immersion ion implantation.
The coatings are characterized by high uniformity, relatively low porosity, and high adhesion to the substrate.
The compound formed after the chemical reaction is deposited not only on the substrate but also on the entire surface of the chamber (including the sputtering target, which reduces the efficiency and productivity of the process)
Glow discharge plasma treatment [321,322,323]It allows to reduce the roughness of the implant surface while increasing its wettability, which contributes to a reduction in bacterial adhesion.
Surface treatment using the glow discharge plasma is carried out in vacuum, which makes it possible to combine this operation with the application of the functional coating in the working chamber of the device.
By adjusting the plasma composition, it is possible to change the chemical composition of the surface, which leads to changes in the physical and mechanical properties of the implant (in particular, the hardness of the surface and the breaking strength of the “implant-bone” connection).
Ultraviolet light treatment (UV photofunctionalization) [324,325]Allows to restore the original surface qualities of “aged” implants and extend their service life by exposing them to ultraviolet light just before insertion into the bone.
Ion implantation [326,327,328,329,330]The introduction of functional metal ions (for example, bioactive ones: Ag, Zn) into the surface is carried out in vacuum.
This operation can be performed as an independent method of surface modification (for example, to increase osteogenic activity and antibacterial activity of implants), or it can be combined with other technological operations (for example, to increase adhesion of vacuum–arc coatings).
Chemical TechniquesChemical (including electrochemical) etching:
-
acid etching (Acid etching) [331,332,333].
-
Alkaline etching (Alkali treatment) [289,334,335,336]
It is possible to obtain different surface textures (its roughness) by adjusting the composition of etching solutions (as a rule, acid concentration), temperature, and etching time.
Acid etching significantly increases the roughness of the implant surface without changing its marginal wetting angle.
Alkaline etching can increase the hydrophilicity of the dental implant surface.
Acid etching is performed at a lower temperature and in a shorter period of time than the commonly used alkali etching treatment.
Electrochemical deposition of coatings, including oxidation (anodizing) [7,292,294,306,337,338,339]A relatively simple inexpensive method of forming coatings (usually oxide films).
Formed coatings have high porosity, which contributes to the improved osteogenesis of implants.
Allows to obtain a hydrophilic surface, which contributes to the better osteointegration of implants.
It is practically impossible to control the structure and properties of coatings in the process of their deposition.
Due to the high level of environmental contamination and danger for personnel, it requires the use of complex treatment and protection facilities.
Electrophoretic deposition of coatings [340,341,342]A relatively simple low-temperature method for forming coatings on an electrically conductive substrate.
Virtually indifferent to the shape of the coated surface, it has high productivity and is well adapted to mass production.
It is possible to control the thickness and morphology of the coating by varying the deposition voltage and time.
Possibility of producing polymer, metal–polymer, ceramic, and composite bioactive coatings (e.g., hydroxyapatite coatings with silver and lignin or chitosan coatings).
Sol-gel method
[343,344,345,346]
The synthesis process is relatively simple and allows excellent control over the composition of coatings and their application to surfaces of virtually any shape.
Allows to achieve uniform distribution of elements of multicomponent systems on the surface of various solids.
Enables the localized delivery of a wide range of drugs (in particular, antimicrobials) at a controlled rate.
Chemical Deposition of Coatings (CDC) (CVD)
[347,348,349]
The high-temperature process of formation of high-purity homogeneous solid films by means of a chemical reaction.
Possibility of coating complex implants (including internal cavities of products).
Coating is only possible on heat-resistant substrates.
Combined
Processing Techniques
Sandblasting of the surface followed by acid etching [70,309,350,351,352]Combines the advantages of both types of surface modification.
-
The SLA (Sand-blasted, Large-gri, Acidetched, translated as “sand-blasted; coarse-grained; acid etched”) surface has a coarse-grained structure; implants show better osseointegration and osteoconductivity, especially in the initial phase (after insertion), compared to the untreated surface.
-
The RBM (Resorbable Blast Media, translated as “resorbable; jet; medium”) surface is formed by sandblasting the implant surface with particles of resorbable Ca-phosphate compound followed by etching in organic low-concentrated acid. It has deeper micropores compared to SLA, which contributes to better osteoconductivity. In addition, it helps reduce osteoporosis.
However, the surface after combined treatment is hydrophobic, which, in general, somewhat complicates the osteointegration of implants.
Plasma electrolytic treatment (e.g., Plasma electrolytic oxidation) [177,302,353,354,355,356,357,358,359,360]Simple, inexpensive, and environmentally safe method.
It allows to apply various (in particular, polymeric) coatings on metal and ceramic substrates with their simultaneous cleaning in order to increase the corrosion resistance of implants.
It is possible to realize the microalloying of both the implant surface as well as of the formed coating (for example, the formation of Ca-P coatings with silver nanoparticles)
Combined treatment with different PVD methods [361]Allows for the formation of multilayer coatings, each layer of which is deposited in a layer-by-layer manner, such as in the deposition of a multilayer coating comprising an oxide sub-layer obtained via gas-thermal oxidation and a calcium phosphate layer obtained via magnetron sputtering.
Cold atmospheric plasma assisted vapor phase CVD method (PECVD) [362,363,364,365]The CVD process uses plasma to decompose the initial substances, activate the substrate surface, and achieve ion etching.
Compared to the traditional method, it is low-temperature (due to plasma amplification), so it allows to obtain coatings on non-thermal substrates.
Formed oxide coatings have superhydrophilic surfaces.
Possibility to apply thin (up to 0.5 µm) glass–ceramic coatings.
Laser chemical vapor deposition (LCVD) [366,367]CVD process that utilizes a laser source to decompose starting substances and activate the substrate surface.
Allows to obtain glass–ceramic coatings on heat-resistant substrates.
By adjusting the laser power and pressure in the chamber, it is possible to obtain ceramic-like oxide films with different microstructures.
Biochemical TechniquesStem cell culturing [368,369,370]Stem cells (mesenchymal (MSC), adipose-derived, bone marrow-derived, or embryonic (ESC)) can be successfully grown in culture media (e.g., serum-free and xenogeneic for adipose-derived MSC) for further culturing on various dental implant surfaces to enhance implant osseintegration and bone regeneration.
Protein-containing coating on implant surfaces [11,371,372,373,374,375,376]It is possible to incorporate protein (e.g., bone morphogenetic protein (BMP-2) or extracellular matrix (ECM)) into polymeric biodegradable coatings applied to implants.
It is possible to apply a liquid protein coating (e.g., platelet- or leukocyte-rich fibrin) to implant surfaces.
Used to accelerate healing of soft and hard tissues of the body.
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

Sotova, C.; Yanushevich, O.; Kriheli, N.; Grigoriev, S.; Evdokimov, V.; Kramar, O.; Nozdrina, M.; Peretyagin, N.; Undritsova, N.; Popelyshkin, E.; et al. Dental Implants: Modern Materials and Methods of Their Surface Modification. Materials 2023, 16, 7383. https://doi.org/10.3390/ma16237383

AMA Style

Sotova C, Yanushevich O, Kriheli N, Grigoriev S, Evdokimov V, Kramar O, Nozdrina M, Peretyagin N, Undritsova N, Popelyshkin E, et al. Dental Implants: Modern Materials and Methods of Their Surface Modification. Materials. 2023; 16(23):7383. https://doi.org/10.3390/ma16237383

Chicago/Turabian Style

Sotova, Catherine, Oleg Yanushevich, Natella Kriheli, Sergey Grigoriev, Vladimir Evdokimov, Olga Kramar, Margarita Nozdrina, Nikita Peretyagin, Nika Undritsova, Egor Popelyshkin, and et al. 2023. "Dental Implants: Modern Materials and Methods of Their Surface Modification" Materials 16, no. 23: 7383. https://doi.org/10.3390/ma16237383

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