Next Article in Journal
The Small Metal-Binding Protein SmbP Simplifies the Recombinant Expression and Purification of the Antimicrobial Peptide LL-37
Next Article in Special Issue
Co-Application of Allicin and Chitosan Increases Resistance of Rosa roxburghii against Powdery Mildew and Enhances Its Yield and Quality
Previous Article in Journal
Anti-Biofilm Coatings Based on Chitosan and Lysozyme Functionalized Magnetite Nanoparticles
Previous Article in Special Issue
Prevalence, Patterns, Association with Biofilm Formation, Effects on Milk Quality and Risk Factors for Antibiotic Resistance of Staphylococci from Bulk-Tank Milk of Goat Herds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antibiotics- and Heavy Metals-Based Titanium Alloy Surface Modifications for Local Prosthetic Joint Infections

by
Jaime Esteban
1,2,*,
María Vallet-Regí
3,4 and
John J. Aguilera-Correa
2,3,*
1
Clinical Microbiology Department, Jiménez Díaz Foundation Health Research Institute, Autonomous University of Madrid, Av. Reyes Católicos 2, 28040 Madrid, Spain
2
Networking Research Centre on Infectious Diseases (CIBER-ID), 28029 Madrid, Spain
3
Department of Chemistry in Pharmaceutical Sciences, Research Institute Hospital 12 de Octubre (i+12), School of Pharmacy, Complutense University of Madrid, Pza. Ramón y Cajal s/n, 28040 Madrid, Spain
4
Networking Research Centre on Bioengineering, Biomaterials and Nanomedicine (CIBER-BBN), 28029 Madrid, Spain
*
Authors to whom correspondence should be addressed.
Submission received: 21 September 2021 / Revised: 5 October 2021 / Accepted: 13 October 2021 / Published: 19 October 2021

Abstract

:
Prosthetic joint infection (PJI) is the second most common cause of arthroplasty failure. Though infrequent, it is one of the most devastating complications since it is associated with great personal cost for the patient and a high economic burden for health systems. Due to the high number of patients that will eventually receive a prosthesis, PJI incidence is increasing exponentially. As these infections are provoked by microorganisms, mainly bacteria, and as such can develop a biofilm, which is in turn resistant to both antibiotics and the immune system, prevention is the ideal approach. However, conventional preventative strategies seem to have reached their limit. Novel prevention strategies fall within two broad categories: (1) antibiotic- and (2) heavy metal-based surface modifications of titanium alloy prostheses. This review examines research on the most relevant titanium alloy surface modifications that use antibiotics to locally prevent primary PJI.

1. Introduction

The use of arthroplasty makes it possible to replace a natural joint with artificial material or a joint prosthesis. Although, arthroplasty is highly effective and has improved the quality of life of millions of patients [1], implant-related complications can appear during the lifetime of patients [2]. One of the most important complications is prosthetic joint infection (PJI), although others may occur. This is probably the most devastating complication due to the high morbidity, mortality, and costs associated with PJI. The mean cost per patient with knee PJI of is USD 52,555 (EUR 40,542), with a range of between USD 24,980 (EUR 19,270.80) for patients with early PJI, and USD 78,111 (EUR 60,257) for those with late PJI [3]. Incidence varies from country to country, between 0.5–2%. Thus, PJI incidence is ranged between 1 and 2% in the United States, and between 0.6% and 0.72% in Nordic countries [4,5]. It is important to know this incidence could be higher in patients undergoing a primary arthroplasty with a history of a PJI in another joint showing up to a three-fold higher risk of PJI [6]. Currently, the 5-year mortality rate associated with PJI is greater than that of breast cancer, melanoma, and Hodgkin’s lymphoma [7].
The aim of this work is to review research on the most relevant titanium alloy surface modifications that use antibiotics to locally prevent primary PJI.

2. Etiopathology

Staphylococci, including Staphylococcus aureus (30–40%) and different species of coagulase-negative staphylococci (27–43%), among which S. epidermidis predominates, are the most common etiological agents associated with PJI [8,9,10,11,12]. Among Gram-negative bacteria (3–9%) [13], enterobacteria and non-fermenting Gram-negative bacilli stand out. However, there could be differences in these patterns according to the characteristics of the infection [9] or the affected joint [14,15]. Polymicrobial infections, or those caused by more than one microorganism, may occur in 10–35% of cases [2,13]. Enterococcus species, Staphylococcus, and various Gram-negative bacilli such as Enterobacteriaceae and Pseudomonas aeruginosa are often associated with these infections.
A problem of growing importance associated with bacterial infections is antibiotic resistance [16]. According to the Centers for Disease Control, approximately 2.8 million antibiotic-resistant bacterial infections take place in the United States and provoke more than 35,000 deaths every year [17]. The main bacteria related to this antibiotic resistance are (as declared by the WHO) Acinetobacter baumannii, P. aeruginosa, enterobacteria (e.g., Klebsiella pneumoniae and Enterobacter cloacae), Enterococcus faecium, S. aureus, Helicobacter pylori, Campylobacter spp., Salmonella spp., Neisseria gonorrhoeae, Streptococcus pneumoniae, Haemophilus influenzae, and Shigella spp. [16,18]. As can be seen, many of the listed bacteria are causative agents of PJI, e.g., S. aureus, P. aeruginosa, K. pneumonia, E. cloacae, and E. faecium, and for that, the antibiotic resistance is also an emerging threat for PJI and must be taken into account in the development of any preventive treatment against them.
One of the most important characteristics in all the aspects of PJI is the ability of microorganisms to form biofilms. A biofilm is a conglomerate of microbial cells of at least one species that is irreversibly attached or not on a surface or an interface, and embedded in a self-produced matrix of polymeric extracellular substances [19], where numerous complex sociomicrobiological interactions prevail [20,21,22]. It is estimated that at least 80% of chronic infections are directly related to the ability of the causative microorganism to develop a biofilm, likely including 100% of all implant-related infections [23,24]. Biofilm formation involves at least three different stages:
(1)
Attachment. Microorganisms come into contact with the surface, a process that is at least partly stochastic, driven by physical and chemical forces [25,26,27]. Furthermore, host proteins rapidly coat the surface of medical devices, facilitating specific adhesion mediated by microbial surface components recognizing adhesive matrix molecules (MSCRAMMs), which are part of the surface of many bacteria, e.g., Staphylococcus spp. [28,29].
(2)
Maturation is characterized by intercellular aggregation coupled to a variety of molecules such as proteins or, usually, exopolysaccharides of a polysaccharide nature, and structuring forces that rearrange the biofilm into three-dimensional structures of variable morphology depending on the species and with microchannels within them [28]. During this stage, one of the most important processes is the production of the exopolysaccharide matrix, whose composition is characteristic of each species, and even of each strain [28,29,30,31]. At this stage, the relatively simple structure that the pre-biofilm acquired in irreversible adhesion takes on a much more structurally complex three-dimensional organization [32]. The nutritional gradient inside the biofilm gives rise to a variety of cells with metabolic differences, including starved cells, dormant cells, viable non-cultivable cells, “persister” cells, and dead cells [27,33].
(3)
Dispersal. This is the process by which mature biofilm cells disperse to adjacent areas passively or actively [23,27]. Through this stage, the infection spreads to adjacent niches in an environment or within the host once nutrients or space has been depleted [32], where it attaches again and restarts the cycle.
The implications of biofilms in treatment and outcomes are enormous, as they confer phenotypical resistance that required the use of new surgeries and prolonged treatments. It is therefore of utmost importance to avoid bacterial colonization of implants and thus avoid the appearance of infection. Moreover, the possibility of an interaction between biofilms, cells, and implanted biomaterials is also of great importance, as the reservoir in the tissue also needs to be removed to cure patients [34,35].

3. Conventional Prevention of Prosthetic Joint Infections

Conventional prevention of PJI includes all measures developed for preventing surgical site infections (SSIs) that have appeared in official guidelines and statements [36,37]. More specific measures for the prevention of PJI have also been published recently [38,39,40,41], and the importance of these measures was considered at the 2nd International Consensus Meeting at Philadelphia as a whole chapter in the General Assembly issues [42]. Factors increasing PJI risk can be grouped into three categories: preoperative, intraoperative, and postoperative [43]. Among the preoperative factors, some well-known ones are obesity, malnutrition, diabetes mellitus, smoking, skin decolonization before surgery, and nasal decolonization. Some important intraoperative factors are surgical-site hair removal, perioperative antibiotics whose use has been successful in reducing the risk of such infections by up to 80% [44,45], and perioperative antibiotic timing [13,46], surgical site skin decolonization, intraarticular irrigation by incorporating antiseptic substances, fibrinolytic agent use, wound closure, implant surface properties, and local antibiotic delivery, since, for instance, the use of a prosthesis cemented with antibiotic-loaded polymethyl-methacrylate cement has been proposed as a potentially useful method that diminishes the risk of PJI [47,48,49,50]. However, the use of antibiotic-loaded cements is not used in all patients so far, since its use has shown a high variability between cohorts, which is translated as a problem when comparing results [51] and requires the employment of specific heat-tolerant antibiotics. Among postoperative factors, some authors consider the typical temporal patterns of C-reactive protein, erythrocyte sedimentation rate test, interleukin 6, and D-Dimer in the early postoperative period [43].
However, even taking all those risk factors into account, there are still several patients who develop PJI after surgery. Several strategies have been devised to avoid this kind of infection.

4. Local Preventive Antibiotic-Based Strategies

During prosthetic implantation, the bone and surrounding tissue must be irrigated; in addition, after implantation, the periprosthetic tissue may be left damaged, avascular, or even, necrotic. These events inherent to surgery locally reduce the concentration of the antibiotic systemically administrated and make it necessary to use a local antibiotic approach with a period of action of hours or days.
On the other hand, the foreign body reaction after the implantation gives rise to an interstitial milieu or a locus minus resistentiae, which is an immunosuppressed fibro-inflammatory zone [52]. This zone is a relatively inaccessible environment for the immune response due to the absence of normal blood supply to the periprosthetic tissue [53], which impairs the ability of lymphocytes, antibodies, and certain antibiotics to properly reach the implant surface and thus prevent and fight infection via the systemic route. For this reason, any prosthesis would be susceptible to be infected not only during the perioperative period but also throughout its whole lifetime [54]. Therefore, a local antibiotic approach with an active period of months or years is required.
The ideal antibiotic-loaded titanium alloy surface modification would require two components: a titanium alloy component and an antibiotic component. The ideal titanium alloy surface modification must not compromise its good corrosion resistance, high strength, low weight, its Young’s modulus of elasticity [55], or non-cytotoxicity. In addition, this titanium alloy surface should be a selective surface able to impair the bacterial adhesion and to favor bone tissue integration [56]. The ideal antibiotic to be loaded should be a broad-spectrum drug based on local prevalence of antibiotic resistance with no adverse local or systemic effects. Further, the ideal antibiotic-loaded titanium alloy should fulfil some market requirements such as an acceptable cost, wide availability, and be easy to manufacture and overcome regulatory issues [57].
The local prevention approach can be classified into two types according to the mechanism of action: passive and active modifications. Passive modifications are surface coatings that endow biomaterial with antibacterial (anti-adherent, bacteriostatic, and/or bactericide) properties without releasing any compound that is responsible for these properties. The active modifications do endow biomaterials with antibacterial properties through a compound released from the material. These active modifications are divided into two groups: active surfaces and coatings. The most recent antibiotic-loaded surface modifications of titanium alloys are illustrated in Figure 1.

4.1. Active Titanium Surfaces Loaded with Antibiotics

The active titanium surfaces loaded with antibiotics can be divided into two categories: nanostructured surfaces and surfaces with covalently bound antibiotics.
The most representative nanostructured titanium surface approaches are summarized in Table 1. This strategy mainly consists of growing nanoscopic carriers made of the bulk alloy and loading them with at least one antibiotic. The most widely used nanostructure is the nanotube, a hollow cylinder without one of its circular faces. Nanotubes can be manufactured using different methods such as sol–gel synthesis, template-assisted synthesis, hydrothermal synthesis, and electrical anodization [58]. Among them, an exponential trend of the use of hydrothermal synthesis and electrical anodization can be observed over last two decades due to their multiple applications [59]. The hydrothermal synthesis modifies the crystallinity of the titanium precursor [60] and allows incorporating other chemical elements into the titanium nanotubes, which enhances their photoelectrochemical [1] properties [59] and confers interesting environmental applications involved, for instance, in the recalcitrant organic pollutant degradation [61]. However, between the two, the most versatile and used in the field of Biomaterials is electrical anodization due to its easy use and thrift.
This nanostructure allows its loading using different methods, mainly simplified lyophilization or soaking.
Bacterial and cellular adhesion are complex processes arising from the interaction between surface properties, biological factors, and environmental conditions. A recent systematic review concludes that there are three reasons why the relationship between surface topography and bacterial attachment can give rise to contradictory results: (i) roughness cannot be the sole descriptor of surface topography; (ii) topographical effects are influenced by the effects of other physicochemical factors, such as surface chemistry; and (iii) different anti-adherent mechanisms may take place at different topographical scales: nanoscale and microscale [62]. The last reason can be also applied to cell attachment. Some authors assert that titanium nanotubes increase the bacterial attachment but have excellent biocompatibility properties because of their enhanced protein interaction (including adsorption and conformation) what improves cellular adhesion and tissue growth [63]. Other authors, by contrast, assert that titanium nanostructures themselves can prevent [64] or reduce bacterial adhesion [65,66] or even biofilm development [67], and also promote cell adhesion and proliferation on the alloy [66,68]. Furthermore, nanotubes composition could be involved in part of these abilities. Thus, for instance, the incorporation of fluorine would be responsible for an anti-adherent ability [65], whilst the additional incorporation of phosphorus would be responsible for better osseointegration [69].
The nanotube diameter is pivotal for the release profile [70]; that is, the larger the diameter, the faster the release. Most of the nanotube-based approaches offer a constant antibiotic release for a few hours after surgery. As a result, this type of approach only guarantees local antibiotic with an active period of hours. The main antibiotic used for loading nanotubes are gentamicin [71,72,73] and vancomycin [74,75] in monotherapy since only few studies have used them in combination [76,77]. Gentamicin is a broad spectrum antibiotic effective against both Gram-positive and Gram-negative bacteria which has a great chemical stability since it remains stable at 4 °C for 30 days and at 23 °C for 7 days [78], and a great thermal stability due to this antibiotic retain its activity even after autoclaving [79]. For its part, vancomycin is a narrow spectrum antibiotic effective against Gram-positive bacteria, the main type of bacteria related to PJIs, and has a reduced chemical stability due to its the concomitant crystalline thermal degradation at physiologic condition [80], which can cause up to a 40% decrease in its activity in 3 weeks [81].
Antibiotics covalently bound to titanium surfaces is another type of active titanium surfaces with antibiotics (Table 2). The main techniques for covalently bound of antibiotics onto titanium surfaces involve the covalent attachment of end-functionalized polymers incorporating an appropriate anchor, e.g., silane anchor, catechol anchor, and phosphor-based anchor [82]. To date, numerous antibiotics have been employed using this strategy such as daptomycin [83], ciprofloxacin [84], doxycycline [85], vancomycin [86], enoxacin [87], bacitracin [88], a new antibiotic such as SPI031 [89], and even antifungals such as caspofungin [86].

4.2. Coating Loaded with Antibiotic for Titanium Alloys

Some of the most relevant coatings loaded with antibiotics described over the last 10 years are summarized in Table 3. In this period, strategies have focused on the design of coatings instead of nanostructures and the covalent binding of antibiotics. This reorientation of local antibiotic therapies may be justified by the huge versatility the coatings offer and their compatibility with not only titanium alloys, but also with almost any material from which a biomedical implant may be made.
Different approaches of deposition of antibiotic-loaded coatings such as sol–gel, covalent immobilization, spraying, electrophoretic, polyelectrolyte, and dip coating have been used on titanium surfaces [117]. Most of the coatings described are degradable over time and are composed of synthetic or natural polymers. The antibiotic release from these degradable coatings depends on their degradation or hydrolysis and the loaded antibiotic quantity depends on both the chemical composition of the coating and the chemical structure and chemical properties of the antibiotic used. The antibiotics that have been loaded onto these coatings are vancomycin [91,96,97,101,105,110,116], aminoglycosides (mainly gentamicin [93,99,100,106,109,113] and tobramycin), tetracyclines (especially doxycycline [98,103] and tetracycline) [102], cephalexin [111], moxifloxacin [112,118], and mixtures of antibiotics such as vancomycin plus tigecycline [108]. Further studies have demonstrated that antifungals, such as fluconazole and anidulafungin, loaded in a coating are effective to prevent C. albicans infection both in vitro [114] and in vivo [115].
The most commonly used synthetic polymers are poly (lactic-co-glycolic acid) (PLGA) (polycaprolactone/polyvinyl alcohol), poly (ethylene glycol-propylene sulphide), and poly-D,L-lactide. Most have been approved by the Food Drug Administration due to their biodegradability and biocompatibility in light of a vast number of recently reviewed studies [119,120]. New strategies based on the use of inorganic [92] and organo-inorganic sol–gels have recently emerged. Some of these organo-inorganic sol–gels have been shown to degrade into non-cytotoxic monomers [112], promote osteoblast proliferation [121], and can even prevent clotting [118]. The most representative natural polymers are based on the use of polysaccharides, e.g., chitosan and hyaluronic acid, and proteins, e.g., silk fibroin and collagen, whose use as drug delivery systems has been recently reviewed [122]. One of these coatings made of natural compounds, an antibiotic-loaded autologous blood glue [113], has attracted attention due to its enormous biocompatibility. This autologous blood glue is composed of a mixture of thrombin, platelet-rich plasma, and bone marrow aspirate and could therefore be loaded with gentamicin and become an antibacterial glue [113]. Several studies have evaluated the antibacterial efficacy of hybrid coatings made of biodegradable polymer and non-biodegradable material. Among them, it is important to consider gentamicin-loaded PLGA and hydroxyapatite, which improve the osteointegration of bone surrounding the implant [99]; vancomycin-loaded gelatin and mesoporous silica nanoparticles, which can carry antibiotic more efficiently [105]; and more complex coatings composed of agarose and nanocrystalline apatite for improved osseointegration, and with mesoporous silica nanoparticles loaded with cephalexin and vascular endothelial growth factor, able to promote vascularization surrounding the implant [123]. Hydroxyapatite coatings favor osteosynthesis [94,107] and prevent the development of fibrous tissue [124] surrounding the implant.
There are two marketed products based on the antibiotic-loaded degradable coating for titanium implants: gentamicin poly (D, L-lactide) (PLLA) coating, and a fast-resorbable hydrogel coating composed of covalently linked hyaluronan and PLLA. Gentamicin PLLA coating is based on a fully resorbable PLLA matrix loaded with gentamicin sulphate which releases 80% of its antibiotic load within the first 48 h [125]. Gentamicin PLLA coating is named PROtect Coating and is only marketed coating Expert Tibial Nail (DePuy Synthes, Bettlach, Switzerland). Though its use is limited to tibial intramedullary nail, it might be theoretically used on any titanium implant. In the first prospective study, Fuchs et al. [126] demonstrated that none of the 19 patients with closed or open tibial fractures who completed the 6-month follow-up showed implant-related infections. Similar results were obtained by Metsemakers et al. [98] in a single-center case series, where they demonstrated again its capacity of preventing implant-related infections in 16 patients with complex open tibia fracture and revision cases after an 18-month follow-up, but they also reported 25% of patients showed a nonunion, and 6.25% of them was a revision case. Finally, the most recent and largest study performed by Schmidmaier et al. [127] in a multicenter study analyzed the outcome of 99 patients with fresh open or closed tibial fractures or undergoing nonunion revision surgery. After an 18-month follow-up, deep SSI or osteomyelitis was only noted in 7.2% of patients after fresh fracture and in 7.7% of patients after revision surgery.
Fast-resorbable hydrogel coating is composed of covalently linked hyaluronan and PLLA and is marketed as Defensive Antibacterial Coating (DAC) (Novagenit Srl, Mezzolombardo, Italy). DAC is the first antimicrobial hydrogel specifically designed to avoid implant-related infections in orthopaedic surgery and trauma, dentistry, and maxillofacial surgery [128,129]. Its antimicrobial ability is due to the hyaluronic-based compounds that reduce microbial adhesion and biofilm formation of both bacteria and yeasts [130]. Moreover, the DAC has demonstrated itself to be capable of entrapping several antibacterial agents at concentrations ranging from 2–10%, released locally for up to 72 h [128]. The safety and efficacy of DAC have been demonstrated by using rabbit models that revealed the capacity of the vancomycin-loaded hydrogel to prevent implant-related infection [131,132]. In a further rabbit model, vancomycin-loaded DAC-coated implants showed no detrimental effects on the bone healing and implant osteointegration [133]. In the first large multicenter randomized prospective clinical trial reported by Romanò et al. [134], a total of 380 patients were included. The patients were randomly dived into two groups which received an implant with the DAC intraoperatively loaded with antibiotics (gentamicin, vancomycin, or vancomycin plus meropenem) or without the coating (control group). Overall, 96.5% of patients were available at a mean follow-up of 14.5 ± 5.5 months. Eleven SSIs were diagnosed in the control group (6%), whilst only one was observed in the treatment group (0.6%). Any patient from the treatment group showed no local or systemic side effects related to or detectable interference with implant osteointegration. In another multicenter prospective study performed by Malizos et al. [135], 256 patients undergoing osteosynthesis surgery for a closed fracture were randomly assigned to receive the DAC loaded with antibiotics (gentamicin, vancomycin, or vancomycin plus meropenem) or to a control group without coating. Six SSIs (4.6%) were observed in the control group compared with none (0%) in the treatment group after a mean follow-up of 18.1 ± 4.5 months. As in the previous study, any patient from the treatment group showed no local or systemic side effects related to or detectable interference with implant osseointegration.
Trentinaglia et al. [136] have recently described an algorithm to calculate the cost-effectiveness of different antibacterial coating strategies applied to joint prostheses, considering both direct and indirect hospital costs. According to their model, an antibacterial coating able to decrease post-surgical infection by 80%, at a cost per patient of EUR 600, would reduce hospital costs by EUR 200 per patient if routinely applied in a population that would theoretically show an expected PJI rate of 2% [137]. At a European level, considering that approximately 2.2 million joint arthroplasties are performed per year, they speculate that a year of delay in the routine use of this kind of coating would result in 35,200 PJI cases per year with associated annual costs of approximately EUR 440 million per year [137].

4.3. The Antibiotic of Choice for Local Antibiotic-Based Therapy

The use of almost any antibiotic in clinical practice is always followed by the development of resistant organisms, and the case of antibiotic-loaded titanium surfaces is not an exception. Antimicrobial resistance is the result of three major factors: (1) the increasing frequency of antimicrobial-resistant phenotypes among microbes resulting from selective pressure exerted by the widespread use of antimicrobials; (2) globalization, which favors the rapid spread of pathogens worldwide; and (3) improper use of antibiotics [138].
The antibiotic of choice for local antibiotic-based therapy should ideally be a broad-spectrum antibiotic that is the least allergenic possible and with no local adverse effects or cytotoxicity; furthermore, these antibiotics should not interfere with osseointegration or be essential for the treatment of PJI [56]. Most of the local antibiotics of choice are broad-spectrum antibiotics used in monotherapy, concretely gentamycin, tobramycin, and vancomycin. To date, there is no antibiotic that is evolution-proof [139,140], as any antibiotic monotherapy is associated with the emergence of antibiotic resistance to that particular antibiotic. This has been described previously, for instance, when a gentamicin-loaded spacer was used in a two-stage replacement which favored the emergence of gentamicin-resistant S. aureus [141] and S. epidermidis [142]. Therefore, the best prophylactic therapy should be based on the use of at least two antibiotics from different antibiotic families, as a handful of studies have done [76,94,108,143]. The microorganisms tested are staphylococci and, to a lesser extent, Gram-negative bacteria, such as E. coli and P. aeruginosa. Given the incidence of PJI (up to 40%) [144], Gram-negative bacteria should always be prevented by the local antibiotic approach.

5. Local Preventive Heavy Metals-Based Strategies

The increasing prevalence of antibiotic resistance among bacteria resulting in the selective pressure which the widespread use of antibiotics exerts on them, the globalization, and the inadequate use of antibiotics in many different settings [138] threaten to completely impede the development of an ideal preventive antibiotic therapy for any type of infection. Given this scenario, new non-antibiotic antimicrobials are gaining increasing importance in the field of PJI prevention strategies (Table 4).
Metals have been used by the Persians, Phoenicians, Greeks, Romans, and Egyptians for their antimicrobial properties for thousands of years [145,146]. Despite the fact that the exact mechanism involved in their broad-spread antibacterial mechanism remains unknown, metals show a higher number of unspecific targets within the bacteria, unlike the antibiotic, which is directly related to a reduced not null emergence of metal resistance. These targets are attacked by metallic cations and/or reactive oxygen species generated by both cations and by metallic oxide [147]. Thereby, the main antibacterial mechanisms of metals that show an antibacterial effect per se can be grouped into four categories: (outer and/or cytoplasmatic) membrane damage, protein blocking/inactivation, protein synthesis blocking, and DNA damage [145] (Figure 2). Different strategies have incorporated heavy metals into titanium surfaces. The main heavy metals used to provide titanium alloys with antimicrobial capacity are silver, copper, and gallium. The type of surface modification used to incorporate the metal on the titanium surfaces are mainly metallurgical addition, co-sputtering, ion implantation, and coatings.
Regarding the use of these metallic-based titanium alloy surface modifications in patients, it is noteworthy that there are no comparative or prospective studies and only retrospective cases of series have been published. Only silver has been proven in humans and has shown low infection risk in clinical studies. There are two technologies marketed nowadays for incorporated silver into titanium alloys: anodization and galvanic deposition. Titanium alloy prostheses with silver incorporated by anodizing is marketed under the name Agluna® (Accentus Medical, Oxfordshire, UK). Anodizing gives rise to the formation of 5 μm diameter circular tanks in the surface of the prosthesis, containing an amorphous titania species where the bulk of the ionic silver is stored. Silver galvanic deposition into titanium alloy prostheses is marketed under the name MUTARS® (tumor system components; Implantcast GmbH, Buxtehude, Germany). Its technology consists of a 15 ± 5 μm-thick silver coating deposited by galvanic deposition on a 200 nm layer of gold that acts as a carrier and bonding layer to the prosthesis. Recently, Deng et al. [173] have pointed out that some factors might underestimate the real anti-infective effect of silver-modified prostheses in clinical studies. First, most of indications published vouch for the use of this type of prosthesis in immunocompromised patients, those with musculoskeletal tumors [174,175,176,177] and/or with a previous PJI [175,176,178], and patients who are themselves more vulnerable to developing PJI [179]. Second, the antibiotherapy is usually administered to all patients, whether or not they carry silver-modified prostheses.
The use of heavy metals for PJI prevention may just be getting started, thus new promising metallic candidates with antimicrobial capacity are yet to be employed. This is the case for metals such as nickel [180,181], cerium [182], selenium [183,184], cesium [185], yttrium [186], palladium [187,188], or superparamagnetic Fe NPs [189].

6. Limitations Associated with Local PJI Prevention

Despite all the potential benefits offered by local prevention strategies for prosthetic joint infections, each has several limitations associated with its use. The advantages and disadvantages related to each preventive approach of PJI are summarized in Table 5.
Titanium nanotubular surfaces have at least five limitations. Firstly, the low drug concentration resulting from sustained release in a non-bacteria environment consumes antibiotic reserves and increases the possibility of developing drug-resistant bacteria in the vicinity of the implant [58]. Therefore, the ideal antibiotic release of a nanotube-based approach should terminate after the infection is eliminated until the next stimulus [58]. This perspective would require the use of self-responsive nanotubes able to release antibiotics before different infection scenarios. Secondly, any metallic implant in the human body degrades due to at least four fundamental phenomena: leaching, wear, corrosion, as well as the phenomenon resulting from the synergy between the latter two, tribocorrosion. Wear studies about the properties of nanostructured titanium surfaces are scarce, and it is known that wear proprieties of nanotubular titanium surfaces have to be hypothetically different as non-nanostructured surface and these nanostructures can be damaged during the prosthesis implantation; nanostructures pulled from the surface could act as debris, able to cause an aseptic loosening [190]. Nanotube fabrication increases the surface area and hence the corrodible area. Corrosion studies of Ti-6Al-4V implants in patients showed that the detection of elevated levels of titanium and normal levels of aluminum and vanadium (relative to a control group without loosening) in the serum or urine of wearers of a prosthesis made of this alloy was associated with the existence of aseptic loosening [191,192,193,194]. Thirdly, nothing is known about the repercussions that this corrosion may have on the useful life of the implant or its osseointegration. Fourthly, the current load methods require the employment of specific equipment (vacuum ovens, agitators, etc.) and long loading times, which make it impossible to load them in the operating theatre for the time being. Fifthly, this approach has no clinical trials to support its widespread use in humans and marketing.
Regarding antibiotics covalently bound to titanium surfaces, there are also important limitations associated with this approach. Unlike nanostructured surface, the antibiotics covalently bound to surfaces are not released into the milieu, and thus can only exert their action on bacteria in direct contact with the modified surface. There is no information about the exact durability of their protection or the hypothetical effect of the release of chemically modified antibiotic on the target bacteria and its role on the emergence of antibiotic resistance. The chemical reaction needed for obtaining these surfaces makes the intra-operative antibiotic load impossible. Finally, there are no clinical trials to back up their use in humans.
Antibiotic-loaded coatings also show limitations. The main limitation is the incomplete protection of the implant, since the intramedullary component of the prosthesis and some modular components (e.g., the acetabular component and the polyethylene insert) cannot be coated. Therefore, an area of susceptibility will exist, where a bacterial infection could proliferate. There is absence of knowledge about the long-term effect that the product resulting from its degradation could exert on the useful life of the implant, its osseointegration, or even, the patient coagulation profile. Although it is the only approach with clinical trials, few antibiotics loaded in such coatings have been used so far.
Heavy metals into titanium surfaces are also associated with some limitations. First, the price of these modified implants is high because they are indicated for a very low number of specific patients [173]. Second, the heavy metals are linked to both local and systemic toxicity. The main side effects of local toxicity are the immunosuppressive effect [195] and the poor or impaired osteointegration that has been reported by both in vitro [196] and in vivo [197] studies. The main systemic side effect related to a titanium alloy surface modified with heavy metals has been described for silver. Argyria, a disease caused by a high silver concentration in the human body, has been reported in up to 23% patients that underwent megaendoprostheses for infection or resection of malignant tumors [198]. In this cohort, no neurological, renal, or hepatic symptoms of silver poisoning were found, and neither a relationship between argyria and the size of the implant or levels of serum silver [198]. Therefore, more studies about the silver intoxication caused by silver-coated implants need to be performed.
Therefore, toxicity is the first concern pertaining to these modifications. With a silver coating, the elevated silver concentration in the blood or in organs has been proven by Gosheger et al. [34], while there were no detectable clinical side effects in this study. The silver ion concentration was lower than the reported harmful concentration, which could be an explanation. Argyria, a disease caused by physiologic silver cation overload, was reported in nearly 22% patients who have received silver-coated prostheses [67]. Therefore, the release of silver ions to the human body after implantation of silver-coated prostheses should be investigated [52]. Impaired osteointegration, which is a special concern for arthroplasty, was generally tested in in vitro co-culture models [68].
Other limitations include the selection of antimicrobial compound. For preventive use, narrow-spectrum antibiotics that cover most potential pathogens are recommended for chemoprophylaxis [36,37]. However, because some antibiotics, such as beta-lactams, can degrade with different factors, such as time or temperature, more stable antibiotics (for example, vancomycin, gentamicin, quinolones) are chosen in many studies. Another important problem not directly related to the biomaterial is the increasing burden of infections caused by antibiotic-resistant microorganisms [8]. The problem of antimicrobial resistance is currently considered one of the most important menaces facing modern medicine [199]. The recent appearance of multidrug-resistant microorganisms has become an extremely important problem with implications for all aspects of medical practice. In orthopaedic surgery, the increasing number of multidrug-resistant organisms, especially Gram-negative organisms, has been described in PJI [8]. This type of infections caused by these microorganisms implies a poor outcome in many cases [200,201,202]. Even silver or copper as heavy metals representants can give rise to heavy metal-resistant Gram-negative bacteria (mainly E. coli and P. aeruginosa) [203,204], one of the bacterial groups related with PJI that is increasing its incidence [205].
In this scenario, the selection of the antimicrobials necessary to prevent PJI infections should consider the existence of multidrug-resistant bacteria [206], which emphasizes the need to select a mixture of at least two antibiotics for preventing PJIs or even using more than one of the preventive approaches described here, e.g., an antibiotic-loaded and heavy metal-dopped surface modification, but also drives the search for new strategies based on the use of iodine-doped titanium alloys [207], antimicrobial peptides [208], and bacteriophages [209,210,211], among others.

7. Conclusions

Research into the development of locally antibiotic therapy approaches is broad and varied, though this review could mark the beginning of a promising journey towards the development of prostheses capable of complete PJI prevention. Despite the numerous preclinical studies that have been conducted, such as those using in vivo models, the move from bench to bedside continues to be hindered by at least two factors, including the low incidence of PJIs and the costs of clinical trials needed to demonstrate the efficacy of these approaches in human beings; indeed, these costs are so high that only large pharmaceutical companies can afford such an investment. These factors may be responsible for the fact that existing multicenter prospective clinical trials are poorly well-structured and often show contradictory or inconclusive results [212]. Thus, the only way patients can benefit from these promising approaches is by improving collaboration between governments, regulatory agencies, industry leaders, and health care payers [213].
Our review highlights that a trend from the antibiotic-loaded surface modifications of the bulk material to the biodegradable antibiotic-load coating can be observed since only two types of these coatings have come to be used in humans. Among heavy metals, silver-modified titanium surfaces are supported by numerous in vitro studies and clinical trials, though other metals such as copper or gallium might stand up as potential future candidates. Furthermore, there is no uniform way of evaluating the efficacy of such approaches. For that, we consider that at least cytotoxicity and cell proliferation should be evaluated in vitro, and that all be tested by using in vivo models. Due to the increasingly threatening emergence of antibiotic resistance, it would therefore be recommendable to use at least two antibiotics or heavy metals for functionalizing the titanium surfaces or antimicrobial substances whose antibacterial mechanisms do not lead to the development of resistant bacterial mutants. Finally, any of the PJI prevention approaches reviewed here are exempt of limitations, many of which should be elucidated by specifically designed studies.

Author Contributions

Conceptualization, J.E. and J.J.A.-C.; writing—original draft preparation, J.E. and J.J.A.-C.; writing—review and editing, J.E., M.V.-R., and J.J.A.-C.; project administration, J.E. and M.V.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the European Research Council through an ERC-2015-AdG-694160 (VERDI) grant.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. Some data are not publicly available since some articles are not open access.

Acknowledgments

We wish to acknowledge Oliver Shaw for his help in reviewing the manuscripts for language-related aspects. Figure 1 and Figure 2 have been created with BioRender.com.

Conflicts of Interest

J.E. received travel grants from Pfizer and conference fees from Biomérieux and Heraeus. These funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Hawker, G.A.; Badley, E.M.; Borkhoff, C.M.; Croxford, R.; Davis, A.M.; Dunn, S.; Gignac, M.A.; Jaglal, S.B.; Kreder, H.J.; Sale, J.E.M. Which patients are most likely to benefit from total joint arthroplasty? Arthritis Rheum. 2013, 65, 1243–1252. [Google Scholar] [CrossRef] [PubMed]
  2. Trampuz, A.; Zimmerli, W. Prosthetic joint infections: Update in diagnosis and treatment. Swiss Med. Wkly. 2005, 135, 243–251. [Google Scholar] [PubMed]
  3. Garrido-Gómez, J.; Arrabal-Polo, M.A.; Girón-Prieto, M.S.; Cabello-Salas, J.; Torres-Barroso, J.; Parra-Ruiz, J. Descriptive analysis of the economic costs of periprosthetic joint infection of the knee for the public health system of Andalusia. J. Arthroplasty 2013, 28, 1057–1060. [Google Scholar] [CrossRef] [PubMed]
  4. Dale, H.; Fenstad, A.M.; Hallan, G.; Havelin, L.I.; Furnes, O.; Overgaard, S.; Pedersen, A.B.; Kärrholm, J.; Garellick, G.; Pulkkinen, P.; et al. Increasing risk of prosthetic joint infection after total hip arthroplasty. Acta Orthop. 2012, 83, 449–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kurtz, S.M.; Lau, E.; Watson, H.; Schmier, J.K.; Parvizi, J. Economic burden of periprosthetic joint infection in the United States. J. Arthroplasty 2012, 27, 61–65. [Google Scholar] [CrossRef] [PubMed]
  6. Chalmers, B.P.; Weston, J.T.; Osmon, D.R.; Hanssen, A.D.; Berry, D.J.; Abdel, M.P. Prior hip or knee prosthetic joint infection in another joint increases risk three-fold of prosthetic joint infection after primary total knee arthroplasty: A matched control study. Bone Jt. J. 2019, 101, 91–97. [Google Scholar] [CrossRef]
  7. DeKeyser, G.J.; Anderson, M.B.; Meeks, H.D.; Pelt, C.E.; Peters, C.L.; Gililland, J.M. Socioeconomic status may not be a risk factor for periprosthetic joint infection. J. Arthroplasty 2020, 35, 1900–1905. [Google Scholar] [CrossRef]
  8. Benito, N.; Franco, M.; Ribera, A.; Soriano, A.; Rodriguez-Pardo, D.; Sorlí, L.; Fresco, G.; Fernández-Sampedro, M.; del Toro, D.M.; Guío, L.; et al. Time trends in the aetiology of prosthetic joint infections: A multicentre cohort study. Clin. Microbiol. Infect. 2016, 22, 732.e1–732.e8. [Google Scholar] [CrossRef] [Green Version]
  9. Benito, N.; Mur, I.; Ribera, A.; Soriano, A.; Rodríguez-Pardo, D.; Sorlí, L.; Cobo, J.; Fernández-Sampedro, M.; del Toro, M.D.; Guío, L.; et al. The different microbial etiology of prosthetic joint infections according to route of acquisition and time after prosthesis implantation, including the role of multidrug-resistant organisms. J. Clin. Med. 2019, 8, 673. [Google Scholar] [CrossRef] [Green Version]
  10. Villa, J.M.; Pannu, T.S.; Theeb, I.; Buttaro, M.A.; Oñativia, J.I.; Carbo, L.; Rienzi, D.H.; Fregeiro, J.I.; Kornilov, N.N.; Bozhkova, S.A.; et al. International organism profile of periprosthetic total hip and knee infections. J. Arthroplasty 2021, 36, 274–278. [Google Scholar] [CrossRef]
  11. Iqbal, F.; Shafiq, B.; Zamir, M.; Noor, S.; Memon, N.; Memon, N.; Dina, T.K. Micro-organisms and risk factors associated with prosthetic joint infection following primary total knee replacement-our experience in Pakistan. Int. Orthop. 2020, 44, 283–289. [Google Scholar] [CrossRef]
  12. Yu, Y.; Kong, Y.; Ye, J.; Wang, A.; Si, W. Microbiological pattern of prosthetic hip and knee infections: A high-volume, single-centre experience in China. J. Med. Microbiol. 2021, 70. [Google Scholar] [CrossRef]
  13. Tande, A.J.; Patel, R. Prosthetic joint infection. Clin. Microbiol. Rev. 2014, 27, 302–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Tsai, Y.; Chang, C.-H.; Lin, Y.-C.; Lee, S.-H.; Hsieh, P.-H.; Chang, Y. Different microbiological profiles between hip and knee prosthetic joint infections. J. Orthop. Surg. Hong Kong 2019, 27. [Google Scholar] [CrossRef] [Green Version]
  15. Paxton, E.S.; Green, A.; Krueger, V.S. Periprosthetic infections of the shoulder: Diagnosis and management. J. Am. Acad. Orthop. Surg. 2019, 27, e935–e944. [Google Scholar] [CrossRef]
  16. Bloom, D.E.; Cadarette, D. Infectious disease threats in the twenty-first century: Strengthening the global response. Front. Immunol. 2019, 10, 549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Centers for Disease Control and Prevention (U.S.). Antibiotic Resistance Threats in the United States, 2019; Centers for Disease Control and Prevention (U.S.): Atlanta, GA, USA, 2019.
  18. Tacconelli, E. Global priority list of antibiotic-resistant bacteria to guide research, discovery, and development of new antibiotics. World Health Organ. 2017, 27, 318–327. [Google Scholar]
  19. Costerton, J.W.; Lewandowski, Z.; Caldwell, D.E.; Korber, D.R.; Lappin-Scott, H.M. Microbial biofilms. Annu. Rev. Microbiol. 1995, 49, 711–745. [Google Scholar] [CrossRef]
  20. Diggle, S.P. Microbial communication and virulence: Lessons from evolutionary theory. Microbiology 2010, 156, 3503–3512. [Google Scholar] [CrossRef] [Green Version]
  21. Nadell, C.D.; Xavier, J.B.; Foster, K.R. The sociobiology of biofilms. FEMS Microbiol. Rev. 2009, 33, 206–224. [Google Scholar] [CrossRef] [Green Version]
  22. West, S.A.; Griffin, A.S.; Gardner, A.; Diggle, S.P. Social evolution theory for microorganisms. Nat. Rev. Microbiol. 2006, 4, 597–607. [Google Scholar] [CrossRef]
  23. Monroe, D. Looking for chinks in the armor of bacterial biofilms. PLoS Biol. 2007, 5, e307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Høiby, N.; Ciofu, O.; Johansen, H.K.; Song, Z.; Moser, C.; Jensen, P.Ø.; Molin, S.; Givskov, M.; Tolker-Nielsen, T.; Bjarnsholt, T. The clinical impact of bacterial biofilms. Int. J. Oral Sci. 2011, 3, 55–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Donlan, R.M. Biofilms: Microbial life on surfaces. Emerg. Infect. Dis. 2002, 8, 881–890. [Google Scholar] [CrossRef]
  26. Kostakioti, M.; Hadjifrangiskou, M.; Hultgren, S.J. Bacterial biofilms: Development, dispersal, and therapeutic strategies in the dawn of the postantibiotic era. Cold Spring Harb. Perspect. Med. 2013, 3, a010306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Hall-Stoodley, L.; Costerton, J.W.; Stoodley, P. Bacterial biofilms: From the natural environment to infectious diseases. Nat. Rev. Microbiol. 2004, 2, 95–108. [Google Scholar] [CrossRef]
  28. Otto, M. Staphylococcal Biofilms. Microbiol. Spectr. 2018, 6. [Google Scholar] [CrossRef] [PubMed]
  29. Otto, M. Staphylococcus Epidermidis—The “accidental” pathogen. Nat. Rev. Microbiol. 2009, 7, 555–567. [Google Scholar] [CrossRef] [Green Version]
  30. Laverty, G.; Gorman, S.P.; Gilmore, B.F. Biomolecular mechanisms of Pseudomonas aeruginosa and Escherichia coli biofilm formation. Pathog. Basel Switz. 2014, 3, 596–632. [Google Scholar] [CrossRef] [Green Version]
  31. Büttner, H.; Mack, D.; Rohde, H. Structural basis of staphylococcus epidermidis biofilm formation: Mechanisms and molecular interactions. Front. Cell. Infect. Microbiol. 2015, 5, 14. [Google Scholar] [CrossRef] [Green Version]
  32. McConoughey, S.J.; Howlin, R.; Granger, J.F.; Manring, M.M.; Calhoun, J.H.; Shirtliff, M.; Kathju, S.; Stoodley, P. Biofilms in periprosthetic orthopedic infections. Future Microbiol. 2014, 9, 987–1007. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563–575. [Google Scholar] [CrossRef] [PubMed]
  34. Valour, F.; Trouillet-Assant, S.; Rasigade, J.-P.; Lustig, S.; Chanard, E.; Meugnier, H.; Tigaud, S.; Vandenesch, F.; Etienne, J.; Ferry, T.; et al. Staphylococcus epidermidis in orthopedic device infections: The role of bacterial internalization in human osteoblasts and biofilm formation. PLoS ONE 2013, 8, e67240. [Google Scholar] [CrossRef] [Green Version]
  35. Ierano, C.; Stewardson, A.J.; Peel, T. Prosthetic Joint Infections; Peel, T., Ed.; Springer International Publishing: Berlin, Germany, 2018; ISBN 978-3-319-65249-8. [Google Scholar]
  36. Berríos-Torres, S.I.; Umscheid, C.A.; Bratzler, D.W.; Leas, B.; Stone, E.C.; Kelz, R.R.; Reinke, C.E.; Morgan, S.; Solomkin, J.S.; Mazuski, J.E.; et al. Centers for disease control and prevention guideline for the prevention of surgical site infection, 2017. JAMA Surg. 2017, 152, 784–791. [Google Scholar] [CrossRef] [PubMed]
  37. World Health Organization. Global Guidelines for the Prevention of Surgical Site Infection; WHO: Geneva, Switzerland, 2016; ISBN 978-92-4-154988-2. [Google Scholar]
  38. Levy, D.M.; Wetters, N.G.; Levine, B.R. Prevention of periprosthetic joint infections of the hip and knee. Am. J. Orthop. Belle Mead NJ 2016, 45, E299–E307. [Google Scholar]
  39. Alamanda, V.K.; Springer, B.D. The prevention of infection. Bone Jt. J. 2019, 101, 3–9. [Google Scholar] [CrossRef]
  40. Berbari, E.; Segreti, J.; Parvizi, J.; Berríos-Torres, S.I. Future research opportunities in peri-prosthetic joint infection prevention. Surg. Infect. 2017, 18, 409–412. [Google Scholar] [CrossRef]
  41. Goswami, K.; Stevenson, K.L.; Parvizi, J. Intraoperative and postoperative infection prevention. J. Arthroplasty 2020, 35, S2–S8. [Google Scholar] [CrossRef] [Green Version]
  42. General Assembly. Available online: https://icmphilly.com/general-assembly/ (accessed on 26 April 2021).
  43. Iannotti, F.; Prati, P.; Fidanza, A.; Iorio, R.; Ferretti, A.; Pèrez Prieto, D.; Kort, N.; Violante, B.; Pipino, G.; Schiavone Panni, A.; et al. Prevention of Periprosthetic joint infection (PJI): A clinical practice protocol in high-risk patients. Trop. Med. Infect. Dis. 2020, 5, 186. [Google Scholar] [CrossRef]
  44. Jämsen, E.; Furnes, O.; Engesaeter, L.B.; Konttinen, Y.T.; Odgaard, A.; Stefánsdóttir, A.; Lidgren, L. Prevention of deep infection in joint replacement surgery. Acta Orthop. 2010, 81, 660–666. [Google Scholar] [CrossRef]
  45. Siddiqi, A.; Forte, S.A.; Docter, S.; Bryant, D.; Sheth, N.P.; Chen, A.F. Perioperative antibiotic prophylaxis in total joint arthroplasty: A systematic review and meta-analysis. J. Bone Jt. Surg. Am. 2019, 101, 828–842. [Google Scholar] [CrossRef]
  46. Del Pozo, J.L.; Patel, R. Clinical practice. Infection associated with prosthetic joints. N. Engl. J. Med. 2009, 361, 787–794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Fillingham, Y.; Greenwald, A.S.; Greiner, J.; Oshkukov, S.; Parsa, A.; Porteous, A.; Squire, M.W. Hip and knee section, prevention, local antimicrobials: Proceedings of international consensus on orthopedic infections. J. Arthroplasty 2019, 34, S289–S292. [Google Scholar] [CrossRef]
  48. Baeza, J.; Cury, M.B.; Fleischman, A.; Ferrando, A.; Fuertes, M.; Goswami, K.; Lidgren, L.; Linke, P.; Manrique, J.; Makar, G.; et al. General assembly, prevention, local antimicrobials: Proceedings of international consensus on orthopedic infections. J. Arthroplasty 2019, 34, S75–S84. [Google Scholar] [CrossRef]
  49. Schiavone Panni, A.; Corona, K.; Giulianelli, M.; Mazzitelli, G.; del Regno, C.; Vasso, M. Antibiotic-loaded bone cement reduces risk of infections in primary total knee arthroplasty? A systematic review. Knee Surg. Sports Traumatol. Arthrosc. Off. J. ESSKA 2016, 24, 3168–3174. [Google Scholar] [CrossRef]
  50. Schmitt, D.R.; Killen, C.; Murphy, M.; Perry, M.; Romano, J.; Brown, N. The impact of antibiotic-loaded bone cement on antibiotic resistance in periprosthetic knee infections. Clin. Orthop. Surg. 2020, 12, 318–323. [Google Scholar] [CrossRef] [PubMed]
  51. Wall, V.; Nguyen, T.-H.; Nguyen, N.; Tran, P.A. Controlling antibiotic release from polymethylmethacrylate bone cement. Biomedicines 2021, 9, 26. [Google Scholar] [CrossRef] [PubMed]
  52. García-Gareta, E.; Davidson, C.; Levin, A.; Coathup, M.J.; Blunn, G.W. Biofilm formation in total hip arthroplasty: Prevention and treatment. RSC Adv. 2016, 6, 80244–80261. [Google Scholar] [CrossRef] [Green Version]
  53. Rakow, A.; Perka, C.; Trampuz, A.; Renz, N. Origin and Characteristics of Haematogenous Periprosthetic Joint Infection. Clinical Microbiology and Infection 2019, 25, 845–850. [Google Scholar] [CrossRef]
  54. Zimmerli, W.; Trampuz, A.; Ochsner, P.E. Prosthetic-joint infections. N. Engl. J. Med. 2004, 351, 1645–1654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Geetha, M.; Singh, A.K.; Asokamani, R.; Gogia, A.K. Ti based biomaterials, the ultimate choice for orthopaedic implants—A review. Prog. Mater. Sci. 2009, 54, 397–425. [Google Scholar] [CrossRef]
  56. Campoccia, D.; Montanaro, L.; Speziale, P.; Arciola, C.R. Antibiotic-loaded biomaterials and the risks for the spread of antibiotic resistance following their prophylactic and therapeutic clinical use. Biomaterials 2010, 31, 6363–6377. [Google Scholar] [CrossRef]
  57. Romanò, C.L.; Petrosillo, N.; Argento, G.; Sconfienza, L.M.; Treglia, G.; Alavi, A.; Glaudemans, A.W.J.M.; Gheysens, O.; Maes, A.; Lauri, C.; et al. The role of imaging techniques to define a peri-prosthetic hip and knee joint infection: Multidisciplinary consensus statements. J. Clin. Med. 2020, 9, 2548. [Google Scholar] [CrossRef] [PubMed]
  58. Li, Y.; Yang, Y.; Li, R.; Tang, X.; Guo, D.; Qing, Y.; Qin, Y. Enhanced antibacterial properties of orthopedic implants by titanium nanotube surface modification: A review of current techniques. Int. J. Nanomed. 2019, 14, 7217–7236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Fu, Y.; Mo, A. A review on the electrochemically self-organized titania nanotube arrays: Synthesis, modifications, and biomedical applications. Nanoscale Res. Lett. 2018, 13, 187. [Google Scholar] [CrossRef]
  60. López Zavala, M.Á.; Lozano Morales, S.A.; Ávila-Santos, M. Synthesis of stable TiO2 nanotubes: Effect of hydrothermal treatment, acid washing and annealing temperature. Heliyon 2017, 3, e00456. [Google Scholar] [CrossRef]
  61. Xie, Y. Photoelectrochemical application of nanotubular titania photoanode. Electrochim. Acta 2006, 51, 3399–3406. [Google Scholar] [CrossRef]
  62. Cheng, Y.; Feng, G.; Moraru, C.I. Micro- and nanotopography sensitive bacterial attachment mechanisms: A review. Front. Microbiol. 2019, 10, 191. [Google Scholar] [CrossRef] [Green Version]
  63. Puckett, S.D.; Taylor, E.; Raimondo, T.; Webster, T.J. The relationship between the nanostructure of titanium surfaces and bacterial attachment. Biomaterials 2010, 31, 706–713. [Google Scholar] [CrossRef] [PubMed]
  64. Bartlet, K.; Movafaghi, S.; Dasi, L.P.; Kota, A.K.; Popat, K.C. Antibacterial activity on superhydrophobic titania nanotube arrays. Colloids Surf. B Biointerfaces 2018, 166, 179–186. [Google Scholar] [CrossRef]
  65. Arenas, M.A.; Pérez-Jorge, C.; Conde, A.; Matykina, E.; Hernández-López, J.M.; Pérez-Tanoira, R.; de Damborenea, J.J.; Gómez-Barrena, E.; Esteba, J. Doped TiO2 anodic layers of enhanced antibacterial properties. Colloids Surf. B Biointerfaces 2013, 105, 106–112. [Google Scholar] [CrossRef]
  66. Aguilera-Correa, J.-J.; Mediero, A.; Conesa-Buendía, F.-M.; Conde, A.; Arenas, M.-Á.; de-Damborenea, J.-J.; Esteban, J. Microbiological and cellular evaluation of a fluorine-phosphorus-doped titanium alloy, a novel antibacterial and osteostimulatory biomaterial with potential applications in orthopedic surgery. Appl. Environ. Microbiol. 2018, 85, e02271-18. [Google Scholar] [CrossRef] [Green Version]
  67. Perez-Jorge, C.; Arenas, M.-A.; Conde, A.; Hernández-Lopez, J.-M.; de Damborenea, J.-J.; Fisher, S.; Hunt, A.M.A.; Esteban, J.; James, G. Bacterial and fungal biofilm formation on anodized titanium alloys with fluorine. J. Mater. Sci. Mater. Med. 2017, 28, 8. [Google Scholar] [CrossRef]
  68. Lozano, D.; Hernández-López, J.M.; Esbrit, P.; Arenas, M.A.; Gómez-Barrena, E.; de Damborenea, J.; Esteban, J.; Pérez-Jorge, C.; Pérez-Tanoira, R.; Conde, A. Influence of the nanostructure of F-doped TiO2 films on osteoblast growth and function: Influence of the nanostructure of F-doped TiO2 films. J. Biomed. Mater. Res. A 2015, 103, 1985–1990. [Google Scholar] [CrossRef]
  69. Aguilera-Correa, J.-J.; Auñón, Á.; Boiza-Sánchez, M.; Mahillo-Fernández, I.; Mediero, A.; Eguibar-Blázquez, D.; Conde, A.; Arenas, M.-Á.; de-Damborenea, J.-J.; Cordero-Ampuero, J.; et al. Urine aluminum concentration as a possible implant biomarker of Pseudomonas aeruginosa Infection using a fluorine- and phosphorus-doped Ti-6Al-4V alloy with osseointegration capacity. ACS Omega 2019, 4, 11815–11823. [Google Scholar] [CrossRef] [Green Version]
  70. Ercan, B.; Taylor, E.; Alpaslan, E.; Webster, T.J. Diameter of titanium nanotubes influences anti-bacterial efficacy. Nanotechnology 2011, 22, 295102. [Google Scholar] [CrossRef]
  71. Lin, W.; Tan, H.; Duan, Z.; Yue, B.; Ma, R.; He, G.; Tang, T. Inhibited bacterial biofilm formation and improved osteogenic activity on gentamicin-loaded titania nanotubes with various diameters. Int. J. Nanomed. 2014, 9, 1215–1230. [Google Scholar] [CrossRef] [Green Version]
  72. Feng, W.; Geng, Z.; Li, Z.; Cui, Z.; Zhu, S.; Liang, Y.; Liu, Y.; Wang, R.; Yang, X. Controlled release behaviour and antibacterial effects of antibiotic-loaded titania nanotubes. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 62, 105–112. [Google Scholar] [CrossRef]
  73. Liu, D.; He, C.; Liu, Z.; Xu, W. Gentamicin coating of nanotubular anodized titanium implant reduces implant-related osteomyelitis and enhances bone biocompatibility in rabbits. Int. J. Nanomed. 2017, 12, 5461–5471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yuan, Z.; Huang, S.; Lan, S.; Xiong, H.; Tao, B.; Ding, Y.; Liu, Y.; Liu, P.; Cai, K. Surface Engineering of titanium implants with enzyme-triggered antibacterial properties and enhanced osseointegration in vivo. J. Mater. Chem. B 2018, 6, 8090–8104. [Google Scholar] [CrossRef]
  75. Zhang, H.; Wang, G.; Liu, P.; Tong, D.; Ding, C.; Zhang, Z.; Xie, Y.; Tang, H.; Ji, F. Vancomycin-loaded titanium coatings with an interconnected micro-patterned structure for prophylaxis of infections: An in vivo study. RSC Adv. 2018, 8, 9223–9231. [Google Scholar] [CrossRef] [Green Version]
  76. Auñón, Á.; Esteban, J.; Doadrio, A.L.; Boiza-Sánchez, M.; Mediero, A.; Eguibar-Blázquez, D.; Cordero-Ampuero, J.; Conde, A.; Arenas, M.; de-Damborenea, J.; et al. Staphylococcus aureus prosthetic joint infection is prevented by a fluorine- and phosphorus-doped nanostructured Ti–6Al–4V alloy loaded with gentamicin and vancomycin. J. Orthop. Res. 2020, 38, 588–597. [Google Scholar] [CrossRef]
  77. Aguilera-Correa, J.-J.; Doadrio, A.L.; Conde, A.; Arenas, M.-A.; de-Damborenea, J.-J.; Vallet-Regí, M.; Esteban, J. Antibiotic release from F-doped nanotubular oxide layer on TI6AL4V alloy to decrease bacterial viability. J. Mater. Sci. Mater. Med. 2018, 29, 118. [Google Scholar] [CrossRef]
  78. Xu, Q.A.; Trissel, L.A.; Saenz, C.A.; Ingram, D.S. Stability of gentamicin sulfate and tobramycin sulfate in autodose infusion system bags. Int. J. Pharm. Compd. 2002, 6, 152–154. [Google Scholar]
  79. Mullins, N.D.; Deadman, B.J.; Moynihan, H.A.; McCarthy, F.O.; Lawrence, S.E.; Thompson, J.; Maguire, A.R. The impact of storage conditions upon gentamicin coated antimicrobial implants. J. Pharm. Anal. 2016, 6, 374–381. [Google Scholar] [CrossRef] [Green Version]
  80. Melichercik, P.; Klapkova, E.; Landor, I.; Judl, T.; Sibek, M.; Jahoda, D. The effect of vancomycin degradation products in the topical treatment of osteomyelitis. Bratisl. Lek. Listy 2014, 115, 796–799. [Google Scholar] [CrossRef] [Green Version]
  81. Mousset, B.; Benoit, M.A.; Delloye, C.; Bouillet, R.; Gillard, J. Biodegradable implants for potential use in bone infection. An in vitro study of antibiotic-loaded calcium sulphate. Int. Orthop. 1995, 19, 157–161. [Google Scholar] [CrossRef] [PubMed]
  82. Chouirfa, H.; Bouloussa, H.; Migonney, V.; Falentin-Daudré, C. Review of titanium surface modification techniques and coatings for antibacterial applications. Acta Biomater. 2019, 83, 37–54. [Google Scholar] [CrossRef]
  83. Chen, C.-W.; Hsu, C.-Y.; Lai, S.-M.; Syu, W.-J.; Wang, T.-Y.; Lai, P.-S. Metal nanobullets for multidrug resistant bacteria and biofilms. Adv. Drug Deliv. Rev. 2014, 78, 88–104. [Google Scholar] [CrossRef]
  84. Badar, M.; Rahim, M.I.; Kieke, M.; Ebel, T.; Rohde, M.; Hauser, H.; Behrens, P.; Mueller, P.P. Controlled drug release from antibiotic-loaded layered double hydroxide coatings on porous titanium implants in a mouse model: Antibiotic-loaded layered double hydroxide coatinGS. J. Biomed. Mater. Res. A 2015, 103, 2141–2149. [Google Scholar] [CrossRef]
  85. Walter, M.S.; Frank, M.J.; Satué, M.; Monjo, M.; Rønold, H.J.; Lyngstadaas, S.P.; Haugen, H.J. Bioactive implant surface with electrochemically bound doxycycline promotes bone formation markers in vitro and in vivo. Dent. Mater. 2014, 30, 200–214. [Google Scholar] [CrossRef]
  86. Kucharíková, S.; Gerits, E.; de Brucker, K.; Braem, A.; Ceh, K.; Majdič, G.; Španič, T.; Pogorevc, E.; Verstraeten, N.; Tournu, H.; et al. Covalent immobilization of antimicrobial agents on titanium prevents Staphylococcus aureus and Candida albicans colonization and biofilm formation. J. Antimicrob. Chemother. 2016, 71, 936–945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Nie, B.; Long, T.; Ao, H.; Zhou, J.; Tang, T.; Yue, B. Covalent immobilization of enoxacin onto titanium implant surfaces for inhibiting multiple bacterial species infection and In Vivo methicillin-resistant staphylococcus aureus infection prophylaxis. Antimicrob. Agents Chemother. 2017, 61, e01766-16. [Google Scholar] [CrossRef] [Green Version]
  88. Nie, B.; Ao, H.; Long, T.; Zhou, J.; Tang, T.; Yue, B. Immobilizing bacitracin on titanium for prophylaxis of infections and for improving osteoinductivity: An in vivo study. Colloids Surf. B Biointerfaces 2017, 150, 183–191. [Google Scholar] [CrossRef]
  89. Gerits, E.; Kucharíková, S.; van Dijck, P.; Erdtmann, M.; Krona, A.; Lövenklev, M.; Fröhlich, M.; Dovgan, B.; Impellizzeri, F.; Braem, A.; et al. Antibacterial activity of a new broad-spectrum antibiotic covalently bound to titanium surfaces. J. Orthop. Res. 2016, 34, 2191–2198. [Google Scholar] [CrossRef]
  90. Chen, C.-P.; Wickstrom, E. Self-protecting bactericidal titanium alloy surface formed by covalent bonding of daptomycin bisphosphonates. Bioconjug. Chem. 2010, 21, 1978–1986. [Google Scholar] [CrossRef] [Green Version]
  91. Lawson, M.C.; Hoth, K.C.; Deforest, C.A.; Bowman, C.N.; Anseth, K.S. Inhibition of Staphylococcus epidermidis biofilms using polymerizable vancomycin derivatives. Clin. Orthop. 2010, 468, 2081–2091. [Google Scholar] [CrossRef] [Green Version]
  92. Mohorcič, M.; Jerman, I.; Zorko, M.; Butinar, L.; Orel, B.; Jerala, R.; Friedrich, J. Surface with antimicrobial activity obtained through silane coating with covalently bound polymyxin B. J. Mater. Sci. Mater. Med. 2010, 21, 2775–2782. [Google Scholar] [CrossRef]
  93. Moskowitz, J.S.; Blaisse, M.R.; Samuel, R.E.; Hsu, H.-P.; Harris, M.B.; Martin, S.D.; Lee, J.C.; Spector, M.; Hammond, P.T. The effectiveness of the controlled release of gentamicin from polyelectrolyte multilayers in the treatment of Staphylococcus aureus infection in a rabbit bone model. Biomaterials 2010, 31, 6019–6030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Alt, V.; Kirchhof, K.; Seim, F.; Hrubesch, I.; Lips, K.S.; Mannel, H.; Domann, E.; Schnettler, R. Rifampicin–fosfomycin coating for cementless endoprostheses: Antimicrobial effects against methicillin-sensitive Staphylococcus aureus (MSSA) and methicillin-resistant Staphylococcus aureus (MRSA). Acta Biomater. 2014, 10, 4518–4524. [Google Scholar] [CrossRef] [PubMed]
  95. Mattioli-Belmonte, M.; Cometa, S.; Ferretti, C.; Iatta, R.; Trapani, A.; Ceci, E.; Falconi, M.; de Giglio, E. Characterization and cytocompatibility of an antibiotic/chitosan/cyclodextrins nanocoating on titanium implants. Carbohydr. Polym. 2014, 110, 173–182. [Google Scholar] [CrossRef] [PubMed]
  96. Ordikhani, F.; Tamjid, E.; Simchi, A. Characterization and antibacterial performance of electrodeposited chitosan–vancomycin composite coatings for prevention of implant-associated infections. Mater. Sci. Eng. C 2014, 41, 240–248. [Google Scholar] [CrossRef]
  97. Zhang, L.; Yan, J.; Yin, Z.; Tang, C.; Guo, Y.; Li, D.; Wei, B.; Gu, Q.; Xu, Y.; Wang, L. Electrospun vancomycin-loaded coating on titanium implants for the prevention of implant-associated infections. Int. J. Nanomed. 2014, 3027. [Google Scholar] [CrossRef] [Green Version]
  98. Metsemakers, W.-J.; Emanuel, N.; Cohen, O.; Reichart, M.; Potapova, I.; Schmid, T.; Segal, D.; Riool, M.; Kwakman, P.H.S.; de Boer, L.; et al. A doxycycline-loaded polymer-lipid encapsulation matrix coating for the prevention of implant-related osteomyelitis due to doxycycline-resistant methicillin-resistant Staphylococcus Aureus. J. Control. Release 2015, 209, 47–56. [Google Scholar] [CrossRef]
  99. Neut, D.; Dijkstra, R.; Thompson, J.; Kavanagh, C.; van der Mei, H.; Busscher, H. A biodegradable gentamicin-hydroxyapatite-coating for infection prophylaxis in cementless hip prostheses. Eur. Cell. Mater. 2015, 29, 42–56. [Google Scholar] [CrossRef]
  100. Diefenbeck, M.; Schrader, C.; Gras, F.; Mückley, T.; Schmidt, J.; Zankovych, S.; Bossert, J.; Jandt, K.D.; Völpel, A.; Sigusch, B.W.; et al. Gentamicin coating of plasma chemical oxidized titanium alloy prevents implant-related osteomyelitis in rats. Biomaterials 2016, 101, 156–164. [Google Scholar] [CrossRef]
  101. Jennings, J.A.; Beenken, K.E.; Skinner, R.A.; Meeker, D.G.; Smeltzer, M.S.; Haggard, W.O.; Troxel, K.S. Antibiotic-loaded phosphatidylcholine inhibits staphylococcal bone infection. World J. Orthop. 2016, 7, 467. [Google Scholar] [CrossRef]
  102. Ma, K.; Cai, X.; Zhou, Y.; Wang, Y.; Jiang, T. In vitro and in vivo evaluation of tetracycline loaded chitosan-gelatin nanosphere coatings for titanium surface functionalization. Macromol. Biosci. 2017, 17, 1600130. [Google Scholar] [CrossRef]
  103. Song, W.; Seta, J.; Chen, L.; Bergum, C.; Zhou, Z.; Kanneganti, P.; Kast, R.E.; Auner, G.W.; Shen, M.; Markel, D.C.; et al. Doxycycline-loaded coaxial nanofiber coating of titanium implants enhances osseointegration and inhibits Staphylococcus Aureus infection. Biomed. Mater. 2017, 12, 045008. [Google Scholar] [CrossRef]
  104. Cheng, T.; Qu, H.; Zhang, G.; Zhang, X. Osteogenic and antibacterial properties of vancomycin-laden mesoporous bioglass/PLGA composite scaffolds for bone regeneration in infected bone defects. Artif. Cells Nanomed. Biotechnol. 2017, 46, 1935–1947. [Google Scholar] [CrossRef] [Green Version]
  105. Zhou, X.; Weng, W.; Chen, B.; Feng, W.; Wang, W.; Nie, W.; Chen, L.; Mo, X.; Su, J.; He, C. Mesoporous silica nanoparticles/gelatin porous composite scaffolds with localized and sustained release of vancomycin for treatment of infected bone defects. J. Mater. Chem. B 2018, 6, 740–752. [Google Scholar] [CrossRef]
  106. Grohmann, S.; Menne, M.; Hesse, D.; Bischoff, S.; Schiffner, R.; Diefenbeck, M.; Liefeith, K. Biomimetic multilayer coatings deliver gentamicin and reduce implant-related osteomyelitis in rats. Biomed. Eng. Biomed. Tech. 2019, 64, 383–395. [Google Scholar] [CrossRef]
  107. Janson, O.; Sörensen, J.H.; Strømme, M.; Engqvist, H.; Procter, P.; Welch, K. Evaluation of an alkali-treated and hydroxyapatite-coated orthopedic implant loaded with tobramycin. J. Biomater. Appl. 2019, 34, 699–720. [Google Scholar] [CrossRef] [PubMed]
  108. Stavrakis, A.I.; Zhu, S.; Loftin, A.H.; Weixian, X.; Niska, J.; Hegde, V.; Segura, T.; Bernthal, N.M. Controlled release of vancomycin and tigecycline from an orthopaedic implant coating prevents Staphylococcus Aureus infection in an open fracture animal model. BioMed Res. Int. 2019, 2019, 1–9. [Google Scholar] [CrossRef] [Green Version]
  109. Thompson, K.; Petkov, S.; Zeiter, S.; Sprecher, C.M.; Richards, R.G.; Moriarty, T.F.; Eijer, H. Intraoperative loading of calcium phosphate-coated implants with gentamicin prevents experimental Staphylococcus Aureus Infection in Vivo. PLoS ONE 2019, 14, e0210402. [Google Scholar] [CrossRef] [PubMed]
  110. Zhang, B.; Braun, B.M.; Skelly, J.D.; Ayers, D.C.; Song, J. Significant suppression of Staphylococcus aureus colonization on intramedullary Ti6Al4V implants surface-grafted with vancomycin-bearing polymer brushes. ACS Appl. Mater. Interfaces 2019, 11, 28641–28647. [Google Scholar] [CrossRef]
  111. Paris, J.L.; Vallet-Regí, M. Mesoporous silica nanoparticles for co-delivery of drugs and nucleic acids in oncology: A review. Pharmaceutics 2020, 12, 526. [Google Scholar] [CrossRef] [PubMed]
  112. Aguilera-Correa, J.J.; Garcia-Casas, A.; Mediero, A.; Romera, D.; Mulero, F.; Cuevas-López, I.; Jiménez-Morales, A.; Esteban, J. A new antibiotic-loaded sol-gel can prevent bacterial prosthetic joint infection: From in vitro studies to an in vivo model. Front. Microbiol. 2019, 10, 2935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Ramalhete, R.; Brown, R.; Blunn, G.; Skinner, J.; Coathup, M.; Graney, I.; Sanghani-Kerai, A. A novel antimicrobial coating to prevent periprosthetic joint infection. Bone Jt. Res. 2020, 9, 848–856. [Google Scholar] [CrossRef]
  114. Romera, D.; Toirac, B.; Aguilera-Correa, J.-J.; García-Casas, A.; Mediero, A.; Jiménez-Morales, A.; Esteban, J. A biodegradable antifungal-loaded sol–gel coating for the prevention and local treatment of yeast prosthetic-joint infections. Materials 2020, 13, 3144. [Google Scholar] [CrossRef]
  115. Garlito-Díaz, H.; Esteban, J.; Mediero, A.; Carias-Cálix, R.A.; Toirac, B.; Mulero, F.; Faus-Rodrigo, V.; Jiménez-Morales, A.; Calvo, E.; Aguilera-Correa, J.J. A new antifungal-loaded sol-gel can prevent candida albicans prosthetic joint infection. Antibiotics 2021, 10, 711. [Google Scholar] [CrossRef]
  116. Ujcic, A.; Krejcikova, S.; Nevoralova, M.; Zhigunov, A.; Dybal, J.; Krulis, Z.; Fulin, P.; Nyc, O.; Slouf, M. Thermoplastic starch composites with titanium dioxide and vancomycin antibiotic: Preparation, morphology, thermomechanical properties, and antimicrobial susceptibility testing. Front. Mater. 2020, 7, 9. [Google Scholar] [CrossRef] [Green Version]
  117. 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]
  118. Aguilera-Correa, J.J.; Vidal-Laso, R.; Carias-Cálix, R.A.; Toirac, B.; García-Casas, A.; Velasco-Rodríguez, D.; Llamas-Sillero, P.; Jiménez-Morales, A.; Esteban, J. A new antibiotic-loaded sol-gel can prevent bacterial intravenous catheter-related infections. Materials 2020, 13, 2946. [Google Scholar] [CrossRef]
  119. da Silva, A.C.; Córdoba de Torresi, S.I. Advances in conducting, biodegradable and biocompatible copolymers for biomedical applications. Front. Mater. 2019, 6, 98. [Google Scholar] [CrossRef] [Green Version]
  120. Ramot, Y.; Haim-Zada, M.; Domb, A.J.; Nyska, A. Biocompatibility and safety of PLA and its copolymers. Adv. Drug Deliv. Rev. 2016, 107, 153–162. [Google Scholar] [CrossRef]
  121. Garcia-Casas, A.; Aguilera-Correa, J.J.; Mediero, A.; Esteban, J.; Jimenez-Morales, A. Functionalization of sol-gel coatings with organophosphorus compounds for prosthetic devices. Colloids Surf. B Biointerfaces 2019, 181, 973–980. [Google Scholar] [CrossRef]
  122. Tong, X.; Pan, W.; Su, T.; Zhang, M.; Dong, W.; Qi, X. Recent advances in natural polymer-based drug delivery systems. React. Funct. Polym. 2020, 148, 104501. [Google Scholar] [CrossRef]
  123. Paris, J.L.; Lafuente-Gómez, N.; Cabañas, M.V.; Román, J.; Peña, J.; Vallet-Regí, M. Fabrication of a nanoparticle-containing 3D porous bone scaffold with proangiogenic and antibacterial properties. Acta Biomater. 2019, 86, 441–449. [Google Scholar] [CrossRef]
  124. Aebli, N.; Krebs, J.; Schwenke, D.; Stich, H.; Schawalder, P.; Theis, J.C. Degradation of hydroxyapatite coating on a well-functioning femoral component. J. Bone Jt. Surg. Br. 2003, 85, 499–503. [Google Scholar] [CrossRef] [Green Version]
  125. Schmidmaier, G.; Wildemann, B.; Stemberger, A.; Haas, N.P.; Raschke, M. Biodegradable poly(D,L-lactide) coating of implants for continuous release of growth factors. J. Biomed. Mater. Res. 2001, 58, 449–455. [Google Scholar] [CrossRef]
  126. Fuchs, T.; Stange, R.; Schmidmaier, G.; Raschke, M.J. The use of gentamicin-coated nails in the tibia: Preliminary results of a prospective study. Arch. Orthop. Trauma Surg. 2011, 131, 1419–1425. [Google Scholar] [CrossRef] [Green Version]
  127. Schmidmaier, G.; Kerstan, M.; Schwabe, P.; Südkamp, N.; Raschke, M. Clinical experiences in the use of a gentamicin-coated titanium nail in tibia fractures. Injury 2017, 48, 2235–2241. [Google Scholar] [CrossRef]
  128. Drago, L.; Boot, W.; Dimas, K.; Malizos, K.; Hänsch, G.M.; Stuyck, J.; Gawlitta, D.; Romanò, C.L. Does implant coating with antibacterial-loaded hydrogel reduce bacterial colonization and biofilm formation in vitro? Clin. Orthop. 2014, 472, 3311–3323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Romanò, C.L.; de Vecchi, E.; Bortolin, M.; Morelli, I.; Drago, L. Hyaluronic acid and its composites as a local antimicrobial/antiadhesive barrier. J. Bone Jt. Infect. 2017, 2, 63–72. [Google Scholar] [CrossRef] [PubMed]
  130. Ardizzoni, A.; Neglia, R.G.; Baschieri, M.C.; Cermelli, C.; Caratozzolo, M.; Righi, E.; Palmieri, B.; Blasi, E. Influence of hyaluronic acid on bacterial and fungal species, including clinically relevant opportunistic pathogens. J. Mater. Sci. Mater. Med. 2011, 22, 2329–2338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Giavaresi, G.; Meani, E.; Sartori, M.; Ferrari, A.; Bellini, D.; Sacchetta, A.C.; Meraner, J.; Sambri, A.; Vocale, C.; Sambri, V.; et al. Efficacy of antibacterial-loaded coating in an in vivo model of acutely highly contaminated implant. Int. Orthop. 2014, 38, 1505–1512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Boot, W.; Vogely, H.C.; Jiao, C.; Nikkels, P.G.; Pouran, B.; van Rijen, M.H.; Ekkelenkamp, M.B.; Hänsch, G.M.; Dhert, W.J.; Gawlitta, D. Prophylaxis of implant-related infections by local release of vancomycin from a hydrogel in rabbits. Eur. Cell. Mater. 2020, 39, 108–120. [Google Scholar] [CrossRef]
  133. Boot, W.; Gawlitta, D.; Nikkels, P.G.J.; Pouran, B.; van Rijen, M.H.P.; Dhert, W.J.A.; Vogely, H.C. Hyaluronic acid-based hydrogel coating does not affect bone apposition at the implant surface in a rabbit model. Clin. Orthop. 2017, 475, 1911–1919. [Google Scholar] [CrossRef] [Green Version]
  134. Romanò, C.L.; Malizos, K.; Capuano, N.; Mezzoprete, R.; D’Arienzo, M.; van der Straeten, C.; Scarponi, S.; Drago, L. Does an antibiotic-loaded hydrogel coating reduce early post-surgical infection after joint arthroplasty? J. Bone Jt. Infect. 2016, 1, 34–41. [Google Scholar] [CrossRef] [Green Version]
  135. Malizos, K.; Blauth, M.; Danita, A.; Capuano, N.; Mezzoprete, R.; Logoluso, N.; Drago, L.; Romanò, C.L. Fast-resorbable antibiotic-loaded hydrogel coating to reduce post-surgical infection after internal osteosynthesis: A multicenter randomized controlled trial. J. Orthop. Traumatol. Off. J. Ital. Soc. Orthop. Traumatol. 2017, 18, 159–169. [Google Scholar] [CrossRef] [Green Version]
  136. Trentinaglia, M.T.; van der Straeten, C.; Morelli, I.; Logoluso, N.; Drago, L.; Romanò, C.L. Economic evaluation of antibacterial coatings on healthcare costs in first year following total joint arthroplasty. J. Arthroplasty 2018, 33, 1656–1662. [Google Scholar] [CrossRef] [PubMed]
  137. Romanò, C.L.; Tsuchiya, H.; Morelli, I.; Battaglia, A.G.; Drago, L. Antibacterial coating of implants: Are we missing something? Bone Jt. Res. 2019, 8, 199–206. [Google Scholar] [CrossRef] [PubMed]
  138. Michael, C.A.; Dominey-Howes, D.; Labbate, M. The antimicrobial resistance crisis: Causes, consequences, and management. Front. Public Health 2014, 2, 145. [Google Scholar] [CrossRef] [PubMed]
  139. Raymond, B. Five rules for resistance management in the antibiotic apocalypse, a road map for integrated microbial management. Evol. Appl. 2019, 12, 1079–1091. [Google Scholar] [CrossRef]
  140. Bell, G.; MacLean, C. The search for ‘evolution-proof’ antibiotics. Trends Microbiol. 2018, 26, 471–483. [Google Scholar] [CrossRef] [PubMed]
  141. Ma, D.; Shanks, R.M.Q.; Davis, C.M.; Craft, D.W.; Wood, T.K.; Hamlin, B.R.; Urish, K.L. Viable bacteria persist on antibiotic spacers following two-stage revision for periprosthetic joint infection. J. Orthop. Res. Off. Publ. Orthop. Res. Soc. 2018, 36, 452–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Schmolders, J.; Hischebeth, G.T.; Friedrich, M.J.; Randau, T.M.; Wimmer, M.D.; Kohlhof, H.; Molitor, E.; Gravius, S. Evidence of MRSE on a gentamicin and vancomycin impregnated polymethyl-methacrylate (PMMA) bone cement spacer after two-stage exchange arthroplasty due to periprosthetic joint infection of the knee. BMC Infect. Dis. 2014, 14, 144. [Google Scholar] [CrossRef]
  143. Aguilera-Correa, J.-J.; Conde, A.; Arenas, M.-A.; de-Damborenea, J.-J.; Marin, M.; Doadrio, A.L.; Esteban, J. Bactericidal activity of the Ti–13Nb–13Zr alloy against different species of bacteria related with implant infection. Biomed. Mater. 2017, 12, 045022. [Google Scholar] [CrossRef]
  144. da Silva, R.B.; Salles, M.J. Outcomes and risk factors in prosthetic joint infections by multidrug-resistant gram-negative bacteria: A retrospective cohort study. Antibiotics 2021, 10, 340. [Google Scholar] [CrossRef]
  145. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial activity of metals: Mechanisms, molecular targets and applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef] [PubMed]
  146. Alexander, J.W. History of the medical use of silver. Surg. Infect. 2009, 10, 289–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Djurišić, A.B.; Leung, Y.H.; Ng, A.M.C.; Xu, X.Y.; Lee, P.K.H.; Degger, N.; Wu, R.S.S. Toxicity of metal oxide nanoparticles: Mechanisms, characterization, and avoiding experimental artefacts. Small 2015, 11, 26–44. [Google Scholar] [CrossRef] [PubMed]
  148. Shirai, T.; Tsuchiya, H.; Shimizu, T.; Ohtani, K.; Zen, Y.; Tomita, K. Prevention of Pin tract infection with titanium-copper alloys. J. Biomed. Mater. Res. B Appl. Biomater. 2009, 91B, 373–380. [Google Scholar] [CrossRef] [Green Version]
  149. Pérez-Tanoira, R.; Pérez-Jorge, C.; Endrino, J.L.; Gómez-Barrena, E.; Horwat, D.; Pierson, J.F.; Esteban, J. Antibacterial properties of biomedical surfaces containing micrometric silver islands. J. Phys. Conf. Ser. 2010, 252, 012015. [Google Scholar] [CrossRef]
  150. Cao, H.; Liu, X.; Meng, F.; Chu, P.K. Biological actions of silver nanoparticles embedded in titanium controlled by micro-galvanic effects. Biomaterials 2011, 32, 693–705. [Google Scholar] [CrossRef]
  151. Secinti, K.D.; Özalp, H.; Attar, A.; Sargon, M.F. Nanoparticle silver ion coatings inhibit biofilm formation on titanium implants. J. Clin. Neurosci. 2011, 18, 391–395. [Google Scholar] [CrossRef]
  152. Ionita, D.; Grecu, M.; Ungureanu, C.; Demetrescu, I. Antimicrobial activity of the surface coatings on TiAlZr implant biomaterial. J. Biosci. Bioeng. 2011, 112, 630–634. [Google Scholar] [CrossRef] [PubMed]
  153. Zhang, E.; Li, F.; Wang, H.; Liu, J.; Wang, C.; Li, M.; Yang, K. A new antibacterial titanium–copper sintered alloy: Preparation and antibacterial property. Mater. Sci. Eng. C 2013, 33, 4280–4287. [Google Scholar] [CrossRef]
  154. Zhao, C.; Feng, B.; Li, Y.; Tan, J.; Lu, X.; Weng, J. Preparation and antibacterial activity of titanium nanotubes loaded with Ag nanoparticles in the dark and under the UV light. Appl. Surf. Sci. 2013, 280, 8–14. [Google Scholar] [CrossRef]
  155. Saidin, S.; Chevallier, P.; Abdul Kadir, M.R.; Hermawan, H.; Mantovani, D. Polydopamine as an intermediate layer for silver and hydroxyapatite immobilisation on metallic biomaterials surface. Mater. Sci. Eng. C 2013, 33, 4715–4724. [Google Scholar] [CrossRef] [PubMed]
  156. De Giglio, E.; Cafagna, D.; Cometa, S.; Allegretta, A.; Pedico, A.; Giannossa, L.C.; Sabbatini, L.; Mattioli-Belmonte, M.; Iatta, R. An innovative, easily fabricated, silver nanoparticle-based titanium implant coating: Development and analytical characterization. Anal. Bioanal. Chem. 2013, 405, 805–816. [Google Scholar] [CrossRef] [PubMed]
  157. Liu, J.; Li, F.; Liu, C.; Wang, H.; Ren, B.; Yang, K.; Zhang, E. Effect of Cu Content on the antibacterial activity of titanium–copper sintered alloys. Mater. Sci. Eng. C 2014, 35, 392–400. [Google Scholar] [CrossRef] [PubMed]
  158. Ren, L.; Ma, Z.; Li, M.; Zhang, Y.; Liu, W.; Liao, Z.; Yang, K. Antibacterial properties of Ti–6Al–4V–XCu alloys. J. Mater. Sci. Technol. 2014, 30, 699–705. [Google Scholar] [CrossRef]
  159. Xie, C.-M.; Lu, X.; Wang, K.-F.; Meng, F.-Z.; Jiang, O.; Zhang, H.-P.; Zhi, W.; Fang, L.-M. Silver nanoparticles and growth factors incorporated hydroxyapatite coatings on metallic implant surfaces for enhancement of osteoinductivity and antibacterial properties. ACS Appl. Mater. Interfaces 2014, 6, 8580–8589. [Google Scholar] [CrossRef]
  160. Rodríguez-Cano, A.; Pacha-Olivenza, M.-Á.; Babiano, R.; Cintas, P.; González-Martín, M.-L. Non-covalent derivatization of aminosilanized titanium alloy implants. Surf. Coat. Technol. 2014, 245, 66–73. [Google Scholar] [CrossRef]
  161. Chen, M.; Zhang, E.; Zhang, L. Microstructure, mechanical properties, bio-corrosion properties and antibacterial properties of Ti–Ag sintered alloys. Mater. Sci. Eng. C 2016, 62, 350–360. [Google Scholar] [CrossRef]
  162. Chen, M.; Yang, L.; Zhang, L.; Han, Y.; Lu, Z.; Qin, G.; Zhang, E. Effect of nano/micro-ag compound particles on the bio-corrosion, antibacterial properties and cell biocompatibility of Ti-Ag alloys. Mater. Sci. Eng. C 2017, 75, 906–917. [Google Scholar] [CrossRef]
  163. Yamanoglu, R.; Efendi, E.; Kolayli, F.; Uzuner, H.; Daoud, I. Production and mechanical properties of Ti–5Al–2.5Fe– x Cu alloys for biomedical applications. Biomed. Mater. 2018, 13, 025013. [Google Scholar] [CrossRef] [PubMed]
  164. Lei, Z.; Zhang, H.; Zhang, E.; You, J.; Ma, X.; Bai, X. Antibacterial activities and biocompatibilities of Ti-Ag alloys prepared by spark plasma sintering and acid etching. Mater. Sci. Eng. C 2018, 92, 121–131. [Google Scholar] [CrossRef] [PubMed]
  165. Peng, C.; Zhang, S.; Sun, Z.; Ren, L.; Yang, K. Effect of annealing temperature on mechanical and antibacterial properties of Cu-bearing titanium alloy and its preliminary study of antibacterial mechanism. Mater. Sci. Eng. C 2018, 93, 495–504. [Google Scholar] [CrossRef]
  166. 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 2019, 100, 38–47. [Google Scholar] [CrossRef]
  167. Cochis, A.; Azzimonti, B.; Chiesa, R.; Rimondini, L.; Gasik, M. Metallurgical gallium additions to titanium alloys demonstrate a strong time-increasing antibacterial activity without any cellular toxicity. ACS Biomater. Sci. Eng. 2019, 5, 2815–2820. [Google Scholar] [CrossRef] [Green Version]
  168. Tao, S.C.; Xu, J.L.; Yuan, L.; Luo, J.M.; Zheng, Y.F. Microstructure, mechanical properties and antibacterial properties of the microwave sintered porous Ti–3Cu alloys. J. Alloys Compd. 2020, 812, 152142. [Google Scholar] [CrossRef]
  169. Bolzoni, L.; Alqattan, M.; Peters, L.; Alshammari, Y.; Yang, F. Ternary Ti alloys functionalised with antibacterial activity. Sci. Rep. 2020, 10, 22201. [Google Scholar] [CrossRef]
  170. Shi, A.; Zhu, C.; Fu, S.; Wang, R.; Qin, G.; Chen, D.; Zhang, E. What controls the antibacterial activity of Ti-Ag alloy, Ag ion or Ti2Ag particles? Mater. Sci. Eng. C 2020, 109, 110548. [Google Scholar] [CrossRef]
  171. Yang, J.; Qin, H.; Chai, Y.; Zhang, P.; Chen, Y.; Yang, K.; Qin, M.; Zhang, Y.; Xia, H.; Ren, L.; et al. Molecular mechanisms of osteogenesis and antibacterial activity of Cu-bearing Ti alloy in a bone defect model with infection in vivo. J. Orthop. Transl. 2021, 27, 77–89. [Google Scholar] [CrossRef]
  172. Zhuang, Y.; Ren, L.; Zhang, S.; Wei, X.; Yang, K.; Dai, K. Antibacterial effect of a copper-containing titanium alloy against implant-associated infection induced by methicillin-resistant Staphylococcus aureus. Acta Biomater. 2021, 119, 472–484. [Google Scholar] [CrossRef] [PubMed]
  173. Deng, W.; Shao, H.; Li, H.; Zhou, Y. Is surface modification effective to prevent periprosthetic joint infection? A systematic review of preclinical and clinical studies. Orthop. Traumatol. Surg. Res. 2019, 105, 967–974. [Google Scholar] [CrossRef] [PubMed]
  174. Hardes, J.; von Eiff, C.; Streitbuerger, A.; Balke, M.; Budny, T.; Henrichs, M.P.; Hauschild, G.; Ahrens, H. Reduction of periprosthetic infection with silver-coated megaprostheses in patients with bone sarcoma: Silver-coated prostheses in sarcoma patients. J. Surg. Oncol. 2010, 101, 389–395. [Google Scholar] [CrossRef]
  175. Hussmann, B.; Johann, I.; Kauther, M.D.; Landgraeber, S.; Jäger, M.; Lendemans, S. Measurement of the silver ion concentration in wound fluids after implantation of silver-coated megaprostheses: Correlation with the clinical outcome. BioMed. Res. Int. 2013, 2013, 1–11. [Google Scholar] [CrossRef] [Green Version]
  176. Wafa, H.; Grimer, R.J.; Reddy, K.; Jeys, L.; Abudu, A.; Carter, S.R.; Tillman, R.M. Retrospective Evaluation of the Incidence of Early Periprosthetic Infection with Silver-Treated Endoprostheses in High-Risk Patients: Case-Control Study. Bone Jt. J. 2015, 97, 252–257. [Google Scholar] [CrossRef]
  177. Schmolders, J.; Koob, S.; Schepers, P.; Pennekamp, P.H.; Gravius, S.; Wirtz, D.C.; Placzek, R.; Strauss, A.C. Lower limb reconstruction in tumor patients using modular silver-coated megaprostheses with regard to perimegaprosthetic joint infection: A case series, including 100 patients and review of the literature. Arch. Orthop. Trauma Surg. 2017, 137, 149–153. [Google Scholar] [CrossRef] [PubMed]
  178. Zajonz, D.; Birke, U.; Ghanem, M.; Prietzel, T.; Josten, C.; Roth, A.; Fakler, J.K.M. Silver-coated modular megaendoprostheses in salvage revision arthroplasty after periimplant infection with extensive bone loss—A pilot study of 34 patients. BMC Musculoskelet. Disord. 2017, 18, 383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Tan, T.L.; Maltenfort, M.G.; Chen, A.F.; Shahi, A.; Higuera, C.A.; Siqueira, M.; Parvizi, J. Development and evaluation of a preoperative risk calculator for periprosthetic joint infection following total joint arthroplasty. J. Bone Jt. Surg. 2018, 100, 777–785. [Google Scholar] [CrossRef] [PubMed]
  180. Jeyaraj Pandian, C.; Palanivel, R.; Dhanasekaran, S. Screening antimicrobial activity of nickel nanoparticles synthesized using Ocimum sanctum leaf extract. J. Nanoparticles 2016, 2016, 1–13. [Google Scholar] [CrossRef] [Green Version]
  181. Ahghari, M.R.; Soltaninejad, V.; Maleki, A. Synthesis of nickel nanoparticles by a green and convenient method as a magnetic mirror with antibacterial activities. Sci. Rep. 2020, 10, 12627. [Google Scholar] [CrossRef]
  182. Bellio, P.; Luzi, C.; Mancini, A.; Cracchiolo, S.; Passacantando, M.; Di Pietro, L.; Perilli, M.; Amicosante, G.; Santucci, S.; Celenza, G. Cerium oxide nanoparticles as potential antibiotic adjuvant. Effects of CeO2 nanoparticles on bacterial outer membrane permeability. Biochim. Biophys. Acta BBA—Biomembr. 2018, 1860, 2428–2435. [Google Scholar] [CrossRef]
  183. Vahdati, M.; Tohidi Moghadam, T. Synthesis and characterization of selenium nanoparticles-lysozyme nanohybrid system with synergistic antibacterial properties. Sci. Rep. 2020, 10, 510. [Google Scholar] [CrossRef] [Green Version]
  184. Geoffrion, L.D.; Hesabizadeh, T.; Medina-Cruz, D.; Kusper, M.; Taylor, P.; Vernet-Crua, A.; Chen, J.; Ajo, A.; Webster, T.J.; Guisbiers, G. Naked selenium nanoparticles for antibacterial and anticancer treatments. ACS Omega 2020, 5, 2660–2669. [Google Scholar] [CrossRef]
  185. Kang, S.-M.; Jang, S.-C.; Heo, N.S.; Oh, S.Y.; Cho, H.-J.; Rethinasabapathy, M.; Vilian, A.T.E.; Han, Y.-K.; Roh, C.; Huh, Y.S. Cesium-induced inhibition of bacterial growth of pseudomonas aeruginosa PAO1 and their possible potential applications for bioremediation of wastewater. J. Hazard. Mater. 2017, 338, 323–333. [Google Scholar] [CrossRef]
  186. Banin, E.; Friedman, A.; Gedanken, A. Lellouche antibacterial and antibiofilm properties of yttrium fluoride nanoparticles. Int. J. Nanomed. 2012, 7, 5611. [Google Scholar] [CrossRef] [Green Version]
  187. Adams, C.P.; Walker, K.A.; Obare, S.O.; Docherty, K.M. Size-dependent antimicrobial effects of novel palladium nanoparticles. PLoS ONE 2014, 9, e85981. [Google Scholar] [CrossRef] [Green Version]
  188. Mohana, S.; Sumathi, S. Multi-functional biological effects of palladium nanoparticles synthesized using agaricus bisporus. J. Clust. Sci. 2020, 31, 391–400. [Google Scholar] [CrossRef]
  189. Xu, C.; Akakuru, O.U.; Zheng, J.; Wu, A. Applications of iron oxide-based magnetic nanoparticles in the diagnosis and treatment of bacterial infections. Front. Bioeng. Biotechnol. 2019, 7, 141. [Google Scholar] [CrossRef]
  190. Ren, K.; Dusad, A.; Zhang, Y.; Wang, D. Therapeutic intervention for wear debris-induced aseptic implant loosening. Acta Pharm. Sin. B 2013, 3, 76–85. [Google Scholar] [CrossRef] [Green Version]
  191. Jacobs, J.J.; Gilbert, J.L.; Urban, R.M. Corrosion of metal orthopaedic implants. J. Bone Jt. Surg. Am. 1998, 80, 268–282. [Google Scholar] [CrossRef]
  192. Jacobs, J.J.; Skipor, A.K.; Black, J.; Urban, R.M.; Galante, J.O. Release and excretion of metal in patients who have a total hip-replacement component made of titanium-base alloy. J. Bone Jt. Surg. Am. 1991, 73, 1475–1486. [Google Scholar] [CrossRef]
  193. Jacobs, J.J.; Silverton, C.; Hallab, N.J.; Skipor, A.K.; Patterson, L.; Black, J.; Galante, J.O. Metal release and excretion from cementless titanium alloy total knee replacements. Clin. Orthop. 1999, 358, 173–180. [Google Scholar] [CrossRef]
  194. Jacobs, J.J.; Skipor, A.K.; Patterson, L.M.; Hallab, N.J.; Paprosky, W.G.; Black, J.; Galante, J.O. Metal release in patients who have had a primary total hip arthroplasty. A prospective, controlled, longitudinal study. J. Bone Jt. Surg. 1998, 80, 1447–1458. [Google Scholar] [CrossRef] [PubMed]
  195. AshaRani, P.V.; Low Kah Mun, G.; Hande, M.P.; Valiyaveettil, S. Cytotoxicity and genotoxicity of silver nanoparticles in human cells. ACS Nano 2009, 3, 279–290. [Google Scholar] [CrossRef] [PubMed]
  196. Pauksch, L.; Hartmann, S.; Rohnke, M.; Szalay, G.; Alt, V.; Schnettler, R.; Lips, K.S. Biocompatibility of silver nanoparticles and silver ions in primary human mesenchymal stem cells and osteoblasts. Acta Biomater. 2014, 10, 439–449. [Google Scholar] [CrossRef] [PubMed]
  197. De Jong, W.H.; van der Ven, L.T.M.; Sleijffers, A.; Park, M.V.D.Z.; Jansen, E.H.J.M.; van Loveren, H.; Vandebriel, R.J. Systemic and immunotoxicity of silver nanoparticles in an intravenous 28 days repeated dose toxicity study in rats. Biomaterials 2013, 34, 8333–8343. [Google Scholar] [CrossRef] [Green Version]
  198. Glehr, M.; Leithner, A.; Friesenbichler, J.; Goessler, W.; Avian, A.; Andreou, D.; Maurer-Ertl, W.; Windhager, R.; Tunn, P.-U. Argyria following the use of silver-coated megaprostheses: No association between the development of local argyria and elevated silver levels. Bone Jt. J. 2013, 95, 988–992. [Google Scholar] [CrossRef]
  199. O’Neill, J. Antimicrobial Resistance: Tackling a Crisis for the Health and Wealth of Nations/the Review on Antimicrobial Resistance Chaired by Jim O’Neill. Available online: https://wellcomecollection.org/works/rdpck35v (accessed on 20 May 2021).
  200. Lora-Tamayo, J.; Murillo, O.; Iribarren, J.A.; Soriano, A.; Sánchez-Somolinos, M.; Baraia-Etxaburu, J.M.; Rico, A.; Palomino, J.; Rodríguez-Pardo, D.; Horcajada, J.P.; et al. A large multicenter study of methicillin-susceptible and methicillin-resistant Staphylococcus aureus prosthetic joint infections managed with implant retention. Clin. Infect. Dis. Off. Publ. Infect. Dis. Soc. Am. 2013, 56, 182–194. [Google Scholar] [CrossRef] [Green Version]
  201. Pfang, B.G.; García-Cañete, J.; García-Lasheras, J.; Blanco, A.; Auñón, Á.; Parron-Cambero, R.; Macías-Valcayo, A.; Esteban, J. Orthopedic implant-associated infection by multidrug resistant enterobacteriaceae. J. Clin. Med. 2019, 8, 220. [Google Scholar] [CrossRef] [Green Version]
  202. Papadopoulos, A.; Ribera, A.; Mavrogenis, A.F.; Rodriguez-Pardo, D.; Bonnet, E.; Salles, M.J.; del Toro, D.M.; Nguyen, S.; Blanco-García, A.; Skaliczki, G.; et al. Multidrug-resistant and extensively drug-resistant gram-negative prosthetic joint infections: Role of surgery and impact of colistin administration. Int. J. Antimicrob. Agents 2019, 53, 294–301. [Google Scholar] [CrossRef]
  203. Randall, C.P.; Gupta, A.; Jackson, N.; Busse, D.; O’Neill, A.J. Silver resistance in gram-negative bacteria: A dissection of endogenous and exogenous mechanisms. J. Antimicrob. Chemother. 2015, 70, 1037–1046. [Google Scholar] [CrossRef] [Green Version]
  204. Bondarczuk, K.; Piotrowska-Seget, Z. Molecular basis of active copper resistance mechanisms in gram-negative bacteria. Cell Biol. Toxicol. 2013, 29, 397–405. [Google Scholar] [CrossRef] [Green Version]
  205. Runner, R.P.; Mener, A.; Roberson, J.R.; Bradbury, T.L.; Guild, G.N.; Boden, S.D.; Erens, G.A. Prosthetic joint infection trends at a dedicated orthopaedics specialty hospital. Adv. Orthop. 2019, 2019, 1–9. [Google Scholar] [CrossRef] [PubMed]
  206. Perez-Jorge, C.; Gomez-Barrena, E.; Horcajada, J.-P.; Puig-Verdie, L.; Esteban, J. Drug treatments for prosthetic joint infections in the era of multidrug resistance. Expert Opin. Pharmacother. 2016, 17, 1233–1246. [Google Scholar] [CrossRef] [PubMed]
  207. Tsuchiya, H.; Shirai, T.; Nishida, H.; Murakami, H.; Kabata, T.; Yamamoto, N.; Watanabe, K.; Nakase, J. Innovative antimicrobial coating of titanium implants with iodine. J. Orthop. Sci. 2012, 17, 595–604. [Google Scholar] [CrossRef] [PubMed]
  208. Browne, K.; Chakraborty, S.; Chen, R.; Willcox, M.D.; Black, D.S.; Walsh, W.R.; Kumar, N. A new era of antibiotics: The clinical potential of antimicrobial peptides. Int. J. Mol. Sci. 2020, 21, E7047. [Google Scholar] [CrossRef] [PubMed]
  209. Morris, J.L.; Letson, H.L.; Elliott, L.; Grant, A.L.; Wilkinson, M.; Hazratwala, K.; McEwen, P. Evaluation of bacteriophage as an adjunct therapy for treatment of peri-prosthetic joint infection caused by Staphylococcus aureus. PLoS ONE 2019, 14, e0226574. [Google Scholar] [CrossRef] [Green Version]
  210. Van Belleghem, J.D.; Manasherob, R.; Miȩdzybrodzki, R.; Rogóż, P.; Górski, A.; Suh, G.A.; Bollyky, P.L.; Amanatullah, D.F. The rationale for using bacteriophage to treat and prevent periprosthetic joint infections. Front. Microbiol. 2020, 11, 591021. [Google Scholar] [CrossRef]
  211. Doub, J.B.; Ng, V.Y.; Wilson, E.; Corsini, L.; Chan, B.K. Successful treatment of a recalcitrant Staphylococcus epidermidis prosthetic knee infection with intraoperative bacteriophage therapy. Pharm. Basel Switz. 2021, 14, 231. [Google Scholar] [CrossRef]
  212. Campoccia, D.; Montanaro, L.; Arciola, C.R. A review of the biomaterials technologies for infection-resistant surfaces. Biomaterials 2013, 34, 8533–8554. [Google Scholar] [CrossRef]
  213. Romanò, C.L.; Scarponi, S.; Gallazzi, E.; Romanò, D.; Drago, L. Antibacterial coating of implants in orthopaedics and trauma: A classification proposal in an evolving panorama. J. Orthop. Surg. 2015, 10, 157. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Different local antibiotic therapy strategies. (a) Antibiotic-loaded nanotubes. (b) Antibiotic covalently bound to titanium alloy. (c) Antibiotic-loaded coating. Yellow represents live bacteria. Red represents dead bacteria.
Figure 1. Different local antibiotic therapy strategies. (a) Antibiotic-loaded nanotubes. (b) Antibiotic covalently bound to titanium alloy. (c) Antibiotic-loaded coating. Yellow represents live bacteria. Red represents dead bacteria.
Antibiotics 10 01270 g001
Figure 2. Main antibacterial mechanisms of heavy metals. PL: peptidoglycan layer. CM: cytoplasmatic membrane. C: cytoplasm.
Figure 2. Main antibacterial mechanisms of heavy metals. PL: peptidoglycan layer. CM: cytoplasmatic membrane. C: cytoplasm.
Antibiotics 10 01270 g002
Table 1. Some of the most relevant studies based on titanium nanotubes loaded with antibiotics.
Table 1. Some of the most relevant studies based on titanium nanotubes loaded with antibiotics.
YearType of Surface ModificationBacteria EvaluatedBacterial StateCytotoxicity (%)Level StudyCell Lines/Animal Used In VivoReference
2014Gentamicin-loaded nanotubes with different diametersSA, SEBiofilmNDIn vitrohBMMS cells[71]
2016Chitosan-coated gentamicin-loaded nanotubesSAPlanktonic20In vitroMG-63 osteoblasts[72]
2017Gentamicin-loaded nanotubes made with anodizationSABiofilmNDIn vivo‒/rabbit[73]
2018Chitosan-hyaluronic acid-coated vancomycin-loaded nanotubesSAPlanktonic/Biofilm0In vitro/in vivoPrimary osteoblasts/rat[74]
Vancomycin-loaded micro-patterningMRSABiofilmNDIn vivo‒/rabbit[75]
Gentamicin and/or vancomycin F-dopped nanotubesSA, SE, ECPlanktonicNDIn vitro‒/‒[66]
2019Gentamicin plus vancomycin F- and P-dopped bottle-shaped nanotubesSABiofilm0In vitro/in vivoMC3T3-E1 osteoblasts/rabbit[76]
Abbreviation: SA: S. aureus; SE: S. epidermidis, EC: E. coli; MRSA: Methicillin-resistant S. aureus; ND: Not determined. hBMMS cells: Human marrow-derived mesenchymal stem cells.
Table 2. Some of the most relevant studies based on antibiotic covalently bound to titanium surfaces.
Table 2. Some of the most relevant studies based on antibiotic covalently bound to titanium surfaces.
YearAntibiotic Covalently BoundBacteria EvaluatedBacterial StateCytotoxicity (%)Level StudyCell Lines/Animal Used In VivoReference
2010DaptomycinSABiofilmNDIn vitro‒/‒[90]
2014Doxycycline0− <40In vitro/in vivoMC3T3-E1 osteoblasts/rabbit[85]
2015CiprofloxacinPABiofilm0In vitro/in vivoNIH3T3 fibroblasts/mouse[84]
2016Vancomycin/caspofunginSA, CABiofilm0In vitro/in vivohME cells/rat[86]
SPI031SA, PABiofilm0In vitro/in vivohBMMS cells, hME cells/mouse[89]
EnoxacinMRSA, SE, ECPlanktonic, Biofilm0In vitro/in vivohBMMS cells/rat[87]
2017BacitracinSABiofilmNDIn vivo‒/rat[88]
Abbreviations: SA: S. aureus; SE: S. epidermidis, EC: E. coli; PA: P. aeruginosa; MRSA: methicillin-resistant S. aureus; ND: Not determined. hBMMS cells: Human marrow-derived mesenchymal stem cells. hME cells: human microvascular endothelial cells.
Table 3. Some of the most relevant studies based on antibiotic loaded coating for titanium implants.
Table 3. Some of the most relevant studies based on antibiotic loaded coating for titanium implants.
YearType of CoatingEvaluated BacteriaBacterial StateCytotoxicity (%)Level StudyCell Lines/Animal Used In VivoReference
2010Vancomycin-loaded PMMASEBiofilmNDIn vitro‒/‒[91]
Inorganic sol–gel with Polymyxin B covalently boundECPlanktonicNDIn vitro‒/‒[92]
Gentamicin-loaded polyelectrolyte multilayerSAPlanktonic, Biofilm0–80In vitro/in vivoMC3T3-E1 osteoblasts/rabbit[93]
2014Rifampicin and fosfomycin-loaded Hydroxyapatite coatingMSSA, MRSABiofilmNDIn vivo‒/rabbit[94]
Ciprofloxacin-loaded chitosan-nanoparticles coatingSAPlanktonic<30In vitroMG63 osteoblast-like cells[95]
Chitosan–vancomycin composite coatingsSAPlanktonic0In vitroMG63 osteoblast-like cells[96]
Vancomycin-loaded PLGA-coatingSAPlanktonic/Biofilm0In vitroMC3T3-E1 osteoblasts/rabbit[97]
2015Doxycycline-loaded polymer-lipid encapsulation matrix coatingMSSA, MRSAPlanktonic, BiofilmNDIn vitro/in vivo‒/mouse[98]
2015PLGA-gentamicin-hydroxyapatite-coatingSA, SEPlanktonic, BiofilmNDIn vitro/in vivo‒/rabbit[99]
2016Gentamicin-derivates coatingSABiofilmNDIn vivo‒/rats[100]
2016Vancomycin-loaded phosphatidyl-cholineSABiofilmNDIn vivo‒/rabbit[101]
2016Tetracycline loaded chitosan-gelatin nanosphere coatingSABiofilm>90In vitro/in vivoMC3T3-E1 osteoblasts/rabbit[102]
2017Doxycycline-loaded coaxial PCL-PVA nanofiber coatingSABiofilmNDIn vivo‒/rat[103]
Tobramycin-loaded PDLLA coatingSABiofilmNDIn vivo‒/rabbit[1]
2018Vancomycin-loaded mesoporous bioglass-PLGA coatingSAPlanktonic, Biofilm0In vitrohBMMS cells[104]
Vancomycin-loaded mesoporous silica nanoparticles-containing gelatin coatingSABiofilm0In vitrohBMMS cells[105]
2019Gentamicin-loaded polyelectrolyte multilayerSA, SEPlanktonic, Biofilm<5In vitro/in vivoMC3T3-E1 osteoblast/rats[106]
Tobramycin-loaded hydroxyapatite coatingSAPlanktonic, BiofilmNDIn vitro/in vivoEndothelial cells, primary osteoblasts/rabbit[107]
Vancomycin plus tigecycline-loaded PEG-PPS coatingSABiofilmNDIn vivo‒/mouse[108]
Gentamicin-loaded calcium phosphate-based coatingSABiofilmNDIn vivo‒/rat[109]
Vancomycin-loaded polymethacrylate coatingSAPlanktonic/BiofilmNDIn vitro/in vivo‒/mouse[110]
2020Cephalexin- and VEGF-loaded agarose-nanocrystalline apatite coatingSAPlanktonic0In vitroMC3T3-E1 osteoblast[111]
Moxifloxacin-loaded organic-inorganic sol–gelSA, SE, ECPlanktonic, Biofilm0In vitro/in vivoMC3T3-E1 osteoblasts/mouse[112]
Gentamicin loaded autologous blood gluePAPlanktonic, Biofilm0In vitrohBMMS cells[113]
Fluconazole/anidulafungin-loaded organic-inorganic sol–gelCA, CPPlanktonic, Biofilm0In vitroMC3T3-E1 osteoblasts[114]
Anidulafungin-loaded organic-inorganic sol–gelCABiofilm-In vivo‒/mouse[115]
Vancomycin-loaded starch coatingSAPlanktonicNDIn vitro‒/‒[116]
Abbreviations: PLGA: poly(lactic-co-glycolic acid); PCL-PVA: polycaprolactone/polyvinyl alcohol; PEG-PPS: poly(ethylene glycol-bl-propylene sulfide); PDLLA: poly (D, L-lactide); SA: S. aureus; SE: S. epidermidis, EC: E. coli; PA: P. aeruginosa; MRSA: methicillin-resistant S. aureus; MSSA: Methicillin-susceptible S. aureus; CA: Candida albicans; CP: Candida parapsilosis; ND: Not determined. hBMMS cells: human bone marrow mesenchymal stem cells.
Table 4. Some of the most relevant studies based on heavy metals incorporation for titanium implants.
Table 4. Some of the most relevant studies based on heavy metals incorporation for titanium implants.
YearType of Surface ModificationIncorporated MetalMetal IncorporationBacteria EvaluatedBacterial StateCytotoxicity (%)Level StudyCell Lines/Animal Used In VivoReference
2009Metallurgical additionCuForgeSA, ECPlanktonic/biofilmCtyocompatibleIn vitro/ in vivoV79 cell line/rabbits[148]
2010Co-sputteringCu-Mn-O, Ag-Mn-Oternary and quaternary oxidesSA, SEPlanktonic-In vitro-[149]
Single step silver plasma immersion ion implantationAgNanoparticlesSA, ECPlanktonicCytocompatibleIn vitroMG63 human osteoblast-like cells[150]
2011 TiO2-chitosan/heparin coatingAgNanoparticlesSABiofilm-In vivo-[151]
Hydroxyapatite coatingAgNanoparticlesECPlanktonic-In vitro-[152]
2013Metallurgical additionCuPowder metallurgySA, ECPlanktonic-In vitro-[153]
Titanium nanotubularAgNanoparticle loadingSA, ECPlanktonic-In vitro-[154]
Polydopamine-modified alloy surfaceAgSilver inonic inmobilizationECPlanktonic-In vitro-[155]
Poly(ethylene glycol diacrylate)-co-acrylic acid coatingAgNanoparticlesSA, EC, PAPlanktonicCytocompatibleIn vitroMG63 human osteoblast-like cells[156]
2014Metallurgical additionCuPowder metallurgySA, ECPlanktonic-In vitro-[157]
Metallurgical additionCuCasting with post-treatmentSA, ECPlanktonicCytocompatibleIn vitroL929 cell line[158]
BMP-2/heparinchitosan-hydroxyapatite coatingAgNanoparticlesSE, ECPlanktonicCytocompatibleIn vitroMC3T3-E1 cells, BMS cells[159]
Aminosilanized titanium alloyAgNanoparticlesSAPlanktonic-In vitro-[160]
2016Metallurgical additionAgSinteringSAPlanktonic-In vitro-[161]
2017Metallurgical additionAgSintering, casting, casting with appropiate post-treatment w/o surface tretamentSAPlanktonicCytocompatibleIn vitroMC3T3-E1 cells[162]
2018Metallurgical additionCuPowder metallurgySA, ECPlanktonicCytocompatibleIn vitroHeLa cells[163]
Metallurgical additionAgSpark plasma sintering and acid etchingSAPlanktonicCytocompatibleIn vitroMC3T3-E1 cells[164]
Metallurgical additionCuCasting with post-treatmentSAPlanktonic-In vitro-[165]
2019Metallurgical additionCuSinteringSABiofilm-In vivo-[166]
Metallurgical additionGaPowder metallurgyMRSAPlanktonic/biofilmCytocompatibleIn vitroATCC CRL-11372 and ATCC HTB-96[167]
2020Metallurgical additionCuMicrowave sinteringSA, ECPlanktonic-In vitro-[168]
Metallurgical additionCuPowder metallurgyECPlanktonic-In vitro-[169]
Metallurgical additionAgCasting with appropiate post-treatment w/o surface tretamentSAPlanktonicCytocompatibleIn vitroMC3T3-E1 cells[170]
2021Metallurgical additionCuAs-castSABiofilm-In vitro/in vivoMouse[171]
Metallurgical additionCuAs-castMRSAPlanktonic/biofilmCytocompatibleIn vitro/in vivoMC3T3-E1 cells/rat[172]
Abbreviations: BMP-2: bone morphology protein-2; BMS: bone marrow stromal cells.
Table 5. Some of the most important advantages and disadvantages related to each preventive approach of PJI.
Table 5. Some of the most important advantages and disadvantages related to each preventive approach of PJI.
Preventive Approach of PJIAdvantagesDisadvantages
Antibiotic-based strategies
Nanostructured titanium surfacesPossibility of increasing the osteointegration of the titanium surfacesReduced durability of antibiotic protection
Unknown biomechanical stability
Loaded antibiotic can act against both bacteria directly adhered on the titanium surface and bacteria near but not in contact with itUnknown effects on the useful life of the implant, osteointegration, and coagulation profile
Impossibility of intra-operative antibiotic load
No clinical trials to support their use
Antibiotics covalently bound to titanium surfacesLong durability of antibiotic protection, up to months or yearsLoaded antibiotic can only act against bacteria directly adhered on the titanium surface
Unknown durability of antibiotic protection
Impossibility of intra-operative antibiotic load
No clinical trials to support their use
Coatings loaded with antibiotic for titanium alloysPossibility of increasing the osteointegration of the titanium surfacesIncomplete surface protection
Loaded antibiotic can act against both bacteria directly adhered on the titanium surface and bacteria near but not in contact with itUnknown effects on the useful life of the implant, osteointegration, and coagulation profile
Possibility of intra-operative antibiotic load
Clinical trials to support their useClinical trials that support their use has been carried out with few antibiotics
Heavy metals-based strategiesBroad spectrum antimicrobial effect (beyond antibacterial effect)Local and systemic toxicity supported by clinical trials
Loaded metals can act against both microorganisms directly adhered on the titanium surface and those near but not in contact with it
Long durability
Clinical trials to support their use
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Esteban, J.; Vallet-Regí, M.; Aguilera-Correa, J.J. Antibiotics- and Heavy Metals-Based Titanium Alloy Surface Modifications for Local Prosthetic Joint Infections. Antibiotics 2021, 10, 1270. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10101270

AMA Style

Esteban J, Vallet-Regí M, Aguilera-Correa JJ. Antibiotics- and Heavy Metals-Based Titanium Alloy Surface Modifications for Local Prosthetic Joint Infections. Antibiotics. 2021; 10(10):1270. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10101270

Chicago/Turabian Style

Esteban, Jaime, María Vallet-Regí, and John J. Aguilera-Correa. 2021. "Antibiotics- and Heavy Metals-Based Titanium Alloy Surface Modifications for Local Prosthetic Joint Infections" Antibiotics 10, no. 10: 1270. https://0-doi-org.brum.beds.ac.uk/10.3390/antibiotics10101270

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