Next Article in Journal
Early Diagnosis and Antibiotic Treatment Combined with Multicomponent Hemodynamic Support for Addressing a Severe Case of Lemierre’s Syndrome
Next Article in Special Issue
Current Knowledge on Biomaterials for Orthopedic Applications Modified to Reduce Bacterial Adhesive Ability
Previous Article in Journal
Levofloxacin Versus Ciprofloxacin-Based Prophylaxis during the Pre-Engraftment Phase in Allogeneic Hematopoietic Stem Cell Transplant Pediatric Recipients: A Single-Center Retrospective Matched Analysis
Previous Article in Special Issue
Collagen Hydrogels Loaded with Silver Nanoparticles and Cannabis Sativa Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Systematic Review

In Vivo Antibacterial Efficacy of Nanopatterns on Titanium Implant Surface: A Systematic Review of the Literature

1
Orthopaedic Medical Center, The Second Hospital of Jilin University, Changchun 130041, China
2
Engineering Research Centre of Molecular Diagnosis and Cell Treatment for Metabolic Bone Disease, The Second Hospital of Jilin University, Changchun 130041, China
*
Author to whom correspondence should be addressed.
Antibiotics 2021, 10(12), 1524; https://doi.org/10.3390/antibiotics10121524
Submission received: 31 October 2021 / Revised: 4 December 2021 / Accepted: 10 December 2021 / Published: 14 December 2021
(This article belongs to the Special Issue Antimicrobial Activity of Medical Materials)

Abstract

:
Background: Bionic surface nanopatterns of titanium (Ti) materials have excellent antibacterial effects in vitro for infection prevention. To date, there is a lack of knowledge about the in vivo bactericidal outcomes of the nanostructures on the Ti implant surfaces. Methods: A systematic review was performed using the PubMed, Embase, and Cochrane databases to better understand surface nanoscale patterns’ in vivo antibacterial efficacy. The inclusion criteria were preclinical studies (in vivo) reporting the antibacterial activity of nanopatterns on Ti implant surface. Ex vivo studies, studies not evaluating the antibacterial activity of nanopatterns or surfaces not modified with nanopatterns were excluded. Results: A total of five peer-reviewed articles met the inclusion criteria. The included studies suggest that the in vivo antibacterial efficacy of the nanopatterns on Ti implants’ surfaces seems poor. Conclusions: Given the small number of literature results, the variability in experimental designs, and the lack of reporting across studies, concluding the in vivo antibacterial effectiveness of nanopatterns on Ti substrates’ surfaces remains a big challenge. Surface coatings using metallic or antibiotic elements are still practical approaches for this purpose. High-quality preclinical data are still needed to investigate the in vivo antibacterial effects of the nanopatterns on the implant surface.

1. Introduction

In the past few decades, with the rapid development of material science and biomedical technology, biomedical metal materials have been widely used in the manufacture of artificial joints, bone trauma fixation, and other osseous tissue replacement or repair medical devices, where titanium (Ti) has become the preferred material due to its similarity with the biomechanical characteristics of human bones and the excellent biocompatibility [1,2]. While many researchers focus on improving the osseointegration ability of Ti implants, the occurrence of postoperative infection, as one of the most severe complications after implantation, cannot be overlooked [3,4]. The infection not only leads to the failure of the implant and the surgery but also increases the patients’ recovery period and makes an economic burden on both patients and the medical system [5]. Although surgical techniques and concepts of sterility have improved in recent years, the average postoperative infection rate is still as high as 2–5% [3,6,7].
The use of antibiotics is a common and effective way to control this issue, but the reason why implant-related infections are challenging to treat lies in the formation of bacterial biofilm on the implant surface. This structure is mainly composed of polysaccharide polymers secreted by bacteria, which can protect the bacteria and resist various antibiotics from outside the biofilm [8]. In addition, the system-administered anti-infection method can also result in a low concentration of drugs in the surgical area due to scars or fibrosis of the surrounding tissues, and thus affects the antibacterial efficiency [5,9,10]. Meanwhile, the extensive use of antibiotics brings the problem of microbial resistance. Surface modifications such as the antibiotic or metallic coating on the implant surface provides more peri-implant antibacterial activity than traditional methods, thus showing an excellent antibacterial effect [11]. However, the antibacterial ability of coating will gradually weaken with the release of antibacterial substances. Moreover, the antibacterial metal ions released from coatings have intrinsic cytotoxicity and may affect the osseointegration performance of Ti implants [12,13].
It has been shown that modifying the surface morphology of Ti materials without adding other chemical reagents can achieve long-term antibacterial effects and inhibit biofilm formation in vitro [14,15,16]. The presence of nano-protrusions on the bionic nanopatterns leads to the destruction and death of the microbes through direct contact and stress concentration [5,17]. The development of surface nanopatterns with efficient antibacterial properties may enable the better clinical application of Ti-related medical materials and address the bacterial resistance problem caused by antibiotic abuse [5]. However, the interactions between nanopatterns and bacteria can be multifaceted, and the antibacterial efficiency in vivo and the role of various factors in regulating bactericidal behavior of the nanopatterns on the Ti implant surface remain unclear [18,19].
This article summarizes and analyzes the preclinical studies of the nanoscale patterns’ antibacterial behaviors on the Ti implant surface to draw a clearer view of the in vivo antibacterial efficacy of bionic nanoscale patterns for the future Ti implant surface bactericidal modification research and applications.

2. Results

2.1. Identification and Selection of Studies

Electronic database searches identified 145 articles (Figure 1). After titles and abstracts were screened for relevance, 78 articles were deemed irrelevant per the inclusion and exclusion criteria. Of the 59 full texts of the in vivo animal studies assessed for eligibility, 5 were selected and reviewed after applying the inclusion and exclusion criteria. The 54 excluded articles did not evaluate the antibacterial activity of nanopatterns or did not modify Ti surfaces with nanopatterns. Meta-analysis was not conducted due to the scarcity and heterogeneity of the studies.

2.2. Study Characteristics

Table 1 and Table 2 provide a general description of the characteristics of the included studies. The five studies used three experimental models (rabbit [20], rat [21,22,23], and mice [24]). Sample sizes varied between 20 [22] and 45 animals [23]. All the studies performed the evaluation of the in vivo antibacterial activity of nanopatterns themselves on the surfaces of the Ti implants. Among these, one study focused on evaluating the antibacterial efficacy of nanopatterns alone on the implant surface [24]. Another evaluated the antibacterial effect of Ag and polydopamine as additional agents to the nanorods (NRDs) on the implant surface [22]. Two studies evaluated the antibacterial effect of metallic agents (Ag and Mg, respectively) as additional modifications on the NTs and NRDs, respectively, on the implant surface [22,23]. One study evaluated the antibacterial effect of fluorine (F)-incorporated NRDs on the implant surface [20]. The monitoring period ranged from 2 days [24] to 8 weeks [20]. All the studies assessed the antibacterial activity of nanopatterns alone on the surfaces. Only one study reported excellent antibacterial efficacy of nanopatterns alone on the implant surface [24]. The most frequently used shape of the implant was a cylinder, and the most used measurement method for in vivo antibacterial efficiency evaluation was bacterial culture and histopathological analysis.

2.3. Risk of Bias and Quality Assessment of the Studies

Assessment of the risk of bias according to SYRCLE (Systematic Review Center for Laboratory Animal Experimentation) guidelines was performed [25]. The included studies presented heterogeneous levels of risk of bias (Figure 2). The evaluation and scoring of the quality of the studies according to ARRIVE (Animal Research: Reporting of In Vivo Experiments) criteria [26,27] (Table S1) yielded an average score of 17.3. None of the studies reported item 19 (Replace, Reduce and Refine) or item 20 (Adverse events) in the abstract. Only one of the studies reported item 22 (Generalization/Applicability) [22] and item 5 (Reasons for animal models) [20]. One of the studies failed to report item 13 (Assignment of animals to experimental groups) [20].

3. Discussion

Modifying the surface nanopattern of materials to achieve antibacterial properties has attracted much attention in the past decade [14,15,16]. It has been shown that the surface morphology of insect wings such as dragonflies and cicadas has excellent antibacterial and antifungal properties [17,28,29]. With the presence of physical nano-protrusions on the surface of insect wings, the antibacterial properties may be attributable to the fact that when microbial cells encounter the surface protrusions, they increase the stress and deformation of the microbial cell membrane structure, leading to their destruction and ultimately leading to cell dissolution and deaths [5,17]. Investigating the surface nanostructure of insect wings and preparing bionic nanopatterns on Ti-based materials according to it has emerged as new ideas for preparing modern antibacterial implants. The modification of the surface morphology of Ti implant to obtain or improve antibacterial ability without adding other chemical reagents, such as silver (Ag) or antibiotics, has been widely reported in ex vivo studies [5,15].
However, the in vivo experimental studies that we retrieved and reviewed did not provide factual data to support this view. Only one study provided evidence that the surface nanostructure on Ti implant has anti-infective effects in vivo [24]. The remaining four articles either indicated that the nanopatterns did not exhibit antibacterial activity in vivo at all [20,22] or only showed limited antibacterial efficacy [21,23]. Moreover, the in vivo antibacterial activity mechanisms of the Ti implant surface morphology seem more complex than just physical puncturing. By modification of the surface morphology on Ti foil using TiO2 NRDs, Zhang et al. reported that the NRD arrays under irradiation with 808 near-infrared (NIR) light produced excellent antibacterial activity against Staphylococcus aureus in vivo through the combined actions of hyperthermia, reactive oxygen species (ROS), and puncturing effects, and could eradicate the attached biofilms on the implant surface in a Kumming mice model [24].
The effect of inhibiting the formation of biofilms is generally considered relevant to the anti-adhesion properties of the surface morphology [30,31,32]. Compared with the patterned surfaces, bacteria were reported more likely to adhere to smoother surfaces [32,33,34,35]. In the study reported by Guan et al., the bacterial coverage on the TiO2 NRDs samples was significantly lower than the bacterial coverage on the pure Ti group that formed a typical biofilm after 48 h ex vivo. The authors attributed this to the anti-adhesion effect of the topography on the implant. However, their in vivo experiment results using a Sprague Dawley rat model did not show the antibacterial activity of the TiO2 NRDs, and there was no difference in the infection rate between the TiO2 NRDs and pure Ti groups [22].
Antibacterial agents can be added to the surface of Ti materials to obtain/improve bactericidal properties [36,37,38]. Most of the studies in this regard were focused on the preparation of antibacterial metal nanoparticles (NPs) fixed on the surface of the Ti substrate by using a carrier or coated on the surface to achieve strong antibacterial ability, good biocompatibility, and stability [36,39,40]. The common metal particles selected for this purpose are Ag, Cu, zinc (Zn), magnesium (Mg), etc., which present antibacterial activities by generating ROS, destroying the structure of bacterial membranes, or regulating the signal transduction pathway of bacteria [41,42,43,44]. Among these, the advantages of a broad spectrum of antibacterial activity make Ag the most studied and widely used metal-based antibacterial agent on Ti substrates [45,46,47]. Ag NPs have been proven to have a good killing effect on both Gram-positive cocci (e.g., S. aureus) and Gram-negative bacilli (e.g., Escherichia coli) [48,49,50]. In the study reported by Guan et al., a novel surface strategy involving the formation of polydopamine (PDA) and Ag nanoparticle-loaded TiO2 NRDs coatings on Ti alloy was developed. In vitro antibacterial experiments showed that, compared to the pure Ti group, the Ag-TiO2@PDA NRDs coating group had adequate antibacterial effects at 7 and 14 days, according to the bacterial counting results. The efficacies were 88.6 ± 1.5% and 80.1 ± 1.1%, respectively, against methicillin-resistant S. aureus, and 91.3 ± 0.5% and 86.2 ± 2.6% against E. coli. Nevertheless, their TiO2 NRDs group also showed a 22.3 ± 3.9% antibacterial efficacy for methicillin-resistant S. aureus (MRSA) and 25.4 ± 12.3% for E. coli. During the in vivo experiments, the materials were implanted into the tibia of a MRSA infected Sprague Dawley rat model. After four weeks, the results of X-ray, micro-CT, and histopathological analysis showed that MRSA could be killed by Ag+, confirming that the Ag-TiO2@PDA NRDs coating also had good antibacterial activity in vivo. However, same as the pure Ti group, TiO2 NRDs on the surface did not show antibacterial activity in vivo [22].
Although the metallic agents can serve as excellent antibacterial elements on the Ti implant surface, the biggest challenge lies in enabling the stable release of such agents at a suitable concentration on the surface of the implants [51,52]. Metal ions released by the coatings are highly mobile and cytotoxic, and their entrance into living cells with high concentrations can kill healthy cells [53]. One solution to this problem is to create nanotube (NT) patterns on the surface of the substrate and load them into the structures for a controlled releasing and long-term antibacterial effect [47,54]. In the study performed by Yang et al., Mg-incorporated NT-modified Ti implants (NT-Mg) were designed and tested to measure the antimicrobial properties. The results demonstrated that NT-Mg implants maintained continuous and reliable release of Mg ion from the NTs, producing long-lasting antimicrobial activity both in vitro and in vivo. The nanotubular structure and alkaline microenvironment during degradation were the two main reasons responsible for the antimicrobial properties of NT-Mg. However, although the nanotubular structure itself exhibited slight anti-infection potential in vivo, the nanotubular structure alone could not combat such a severe implant-related bone infection [23].
In recent years, to avoid the cytotoxicity of the metal ion as the antibacterial agent, studies of non-metallic elements serving as the bactericidal agent of the surface of Ti implants have emerged [20,55,56,57]. It has been reported that fluorine (F)-doped nanopatterns on Ti material surfaces have excellent antibacterial ability against numerous bacteria in vitro and good cytocompatibility and osteoblastic activity [20,55,58,59]. In the study reported by Zhou et al., F-doped Sr1-HA (strontium containing hydroxyapatite) NRDs on microporous TiO2 implant led to the significantly improved antibacterial activities in a S. aureus infected New Zealand rabbit model at 8 weeks according to the bacteria counting results, and the effect was related to the incorporated F dose. However, the average CFU counting results exhibited the Sr1-HA NRDs on microporous TiO2 did not possess antibacterial activity against S. aureus in vivo at 8 weeks [20].
In addition to the inorganic antibacterial agents coating strategies, the antibiotic coating can also be added to the Ti implant surface to achieve antibacterial effects [60,61,62,63]. Under ideal conditions, the antibiotics loaded on the surface should release in a controlled path and speed, reach the effective drug concentration, and maintain a long sufficient sterilization time, reducing the risk of bacterial resistance problems caused by antibiotic abuse [64,65,66]. Considering that both aerobic and anaerobic organisms can cause bone infections and the high frequency of polymicrobial infections [67,68,69], broad-spectrum antibiotics such as rifampicin, gentamicin, vancomycin, etc., are recommended as the loaded agents [70,71]. Under ideal conditions, the antibiotics released by the prepared nanomaterials should reach the effective drug concentration and maintain a long sufficient sterilization time [64,72]. Zhang et al. compared the antibacterial efficacy of TiO2 NTs loaded with vancomycin (NT-V) with those of the NTs and commercially pure (Cp-Ti) groups in an S. aureus infected Sprague Dawley rat model. Compared with the other groups, NT-V showed an excellent antibacterial effect both in vitro and in vivo. Although the NTs reduced the surface bacterial adhesion in vitro, implant infection still developed in the in vivo experiments. The infection rate in the NT-V group was 0% at 30 days, while that of the NT and Cp-Ti groups reached 92% and 100%, respectively [21]. Although S. aureus is the primary pathogen responsible for bone infections [73,74,75], the evaluation of the in vivo anti-Staphylococcus activity alone further undermined the clinical translation value of the studies retrieved.
This systematic review has several limitations. Although a systematic literature search was performed and no similar reviews were identified, the registration of this systematic review was not completed before the data extraction was finished, and the limited number of studies and heterogeneity in reporting and experimental designs may influence outcomes, hinder result comparison, and preclude meta-analysis. Additionally, the findings of this review are supported by basic science studies (level 3). No studies of level 2 evidence or higher were identified. Finally, it is possible all relevant articles were not identified with our search criteria.

4. Materials and Methods

4.1. Systematic Literature Search

A comprehensive systematic literature search was conducted to answer the question, “Is the in vivo antibacterial efficacy of Ti implant enhanced by surface nanomorphology modification alone?” following the PRISMA (Preferred Reporting Items for Systematic Reviews and Meta-Analyses) guidelines [76]. The systematic search was conducted in August 2021 using three electronic databases: PubMed, Embase, and Cochrane. A general search was conducted using the following terms: (“nano”(All Fields) OR “nano-scale”(All Fields) OR “nanopattern”(All Fields) OR “nanomorphology”(All Fields) OR “nanostructure”(All Fields)) AND (“surface”(All Fields) OR “surfaces”(All Fields) OR “surfacing”(All Fields)) AND (“antimicrobial”(All Fields) OR “antibacterial”(All Fields)) AND (“titanium”(MeSH Terms) OR “titanium”(All Fields)) AND (“implantation”(All Fields) OR “implant”(All Fields) OR “implants”(All Fields)) AND “animals”(MeSH Terms).
The results were filtered to English and full-text articles. References from the articles were reviewed to confirm the completeness of the identified literature.

4.2. Exclusion and Inclusion Criteria

In vivo peer-reviewed studies evaluating the antibacterial efficacy of nanopatterns on Ti implant surfaces were included in this systematic review. Exclusion criteria for all studies were as follows: articles not written in English; review and expert opinion articles, conference proceedings, and presentations; ex vivo studies; studies that did not evaluate the antibacterial activity of nanopatterns or did not modify Ti surfaces with nanopatterns (Table S2). One author (Y.S.) performed the literature search, and two authors (Y.S., Y.Y.) independently reviewed the search results. Titles and abstracts were reviewed for all search results. Full-text articles were obtained to determine studies that met inclusion and exclusion criteria. If disagreement occurred, the senior author (J.W.) was consulted.

5. Conclusions

In contrast with the widely reported excellent in vitro antibacterial effectiveness, the in vivo antibacterial efficacy of the nanopatterns on Ti implants’ surfaces seems poor according to the preclinical studies we assessed. Moreover, given the small number of literature results, the variability in experimental designs, and the lack of reporting across studies, concluding the in vivo antibacterial effectiveness of the nanostructures on Ti substrates’ surfaces remains a big challenge. Surface modifications using metallic NPs or antibiotics coating are still practical approaches for achieving or improving the in vivo bactericidal activities. Despite our inability to ascertain the in vivo antibacterial efficacy of the surface nanopatterns, high-quality preclinical data are still needed to investigate the antibacterial effects of the nanopatterns on the implant surface and the mechanisms.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/antibiotics10121524/s1, Table S1: Checklist of ARRIVE criteria reported by the included studies, Table S2: Study inclusion and exclusion criteria.

Author Contributions

All the authors contributed to the conceptualization and design of this review. Data collection (literature search and study selection): Y.S. and Y.Y.; data analysis and interpretation: Y.S., Y.Y. and J.W.; writing—original draft preparation, Y.S., Y.Y., W.J., H.B., H.L. and J.W.; writing—review and editing, W.J. and J.W.; funding acquisition, Y.S. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Finance Department of Jilin Province, P.R. China (Grant Number 3D5197435429 and 3D5204944429), and the Science and Technology Department of Jilin Province, P.R. China (Grant Number 3D5203928429 and 3D5204471429).

Data Availability Statement

Not applicable.

Acknowledgments

The authors appreciate Jichao Liu and Shiyan Yang (Division of Orthopedics, Second Hospital of Jilin University, Jilin, China) for giving suggestions on the method design and data collection in this study.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Agsilver
ARRIVEAnimal Research Reporting of In Vivo Experiments
CFUColony-Forming Unit
Cp-Ticommercially pure titanium
Cucopper
E. coliEscherichia coli
Ffluorine
HAhydroxyapatite
Mgmagnesium
MRSAmethicillin-resistant Staphylococcus aureus
NIRnear-infrared
NRDnanorod
NTnanotube
PBSphosphate buffered saline
PDApolydopamine
PRISMAPreferred Reporting Items for Systematic Reviews and Meta-Analyses
ROSreactive oxygen species
Srstrontium
S. aureusStaphylococcus aureus
SYRCLESystematic Review Center for Laboratory Animal Experimentation
Tititanium
Znzinc

References

  1. Gittens, R.A.; Olivares-Navarrete, R.; Schwartz, Z.; Boyan, B.D. Implant osseointegration and the role of microroughness and nanostructures: Lessons for spine implants. Acta Biomater. 2014, 10, 3363–3371. [Google Scholar] [CrossRef] [Green Version]
  2. Souza, J.C.M.; Sordi, M.B.; Kanazawa, M.; Ravindran, S.; Henriques, B.; Silva, F.S.; Aparicio, C.; Cooper, L.F. Nano-scale modification of titanium implant surfaces to enhance osseointegration. Acta Biomater. 2019, 94, 112–131. [Google Scholar] [CrossRef] [PubMed]
  3. Roy-Chaudhury, P.; Munda, R. Infections associated with surgical implants. N. Engl. J. Med. 2004, 351, 193–195. [Google Scholar]
  4. 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] [PubMed] [Green Version]
  5. Liu, J.; Liu, J.; Attarilar, S.; Wang, C.; Tamaddon, M.; Yang, C.; Xie, K.; Yao, J.; Wang, L.; Liu, C.; et al. Nano-modified titanium implant materials: A way toward improved antibacterial properties. Front Bioeng. Biotechnol. 2020, 8, 576969. [Google Scholar] [CrossRef]
  6. Darouiche, R.O. Treatment of infections associated with surgical implants. N. Engl. J. Med. 2004, 350, 1422–1429. [Google Scholar] [CrossRef]
  7. Tanner, M.C.; Fischer, C.; Schmidmaier, G.; Haubruck, P. Evidence-based uncertainty: Do implant-related properties of titanium reduce the susceptibility to perioperative infections in clinical fracture management? A systematic review. Infection 2021, 49, 813–821. [Google Scholar] [CrossRef]
  8. 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]
  9. Andersson, D.I.; Hughes, D. Evolution of antibiotic resistance at non-lethal drug concentrations. Drug Resist. Updat. 2012, 15, 162–172. [Google Scholar] [CrossRef] [PubMed]
  10. Andersson, D.I.; Hughes, D. Microbiological effects of sublethal levels of antibiotics. Nat. Rev. Microbiol. 2014, 12, 465–478. [Google Scholar] [CrossRef] [PubMed]
  11. 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]
  12. Shen, X.T.; Zhang, Y.Z.; Xiao, F.; Zhu, J.; Zheng, X.D. Effects on cytotoxicity and antibacterial properties of the incorporations of silver nanoparticles into the surface coating of dental alloys. J. Zhejiang. Univ. Sci. B 2017, 18, 615–625. [Google Scholar] [CrossRef] [Green Version]
  13. Shivaram, A.; Bose, S.; Bandyopadhyay, A. Understanding long-term silver release from surface modified porous titanium implants. Acta Biomater. 2017, 58, 550–560. [Google Scholar] [CrossRef]
  14. 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] [PubMed]
  15. Hasan, J.; Crawford, R.J.; Ivanova, E.P. Antibacterial surfaces: The quest for a new generation of biomaterials. Trends Biotechnol. 2013, 31, 295–304. [Google Scholar] [CrossRef]
  16. Modaresifar, K.; Azizian, S.; Ganjian, M.; Fratila-Apachitei, L.E.; Zadpoor, A.A. Bactericidal effects of nanopatterns: A systematic review. Acta Biomater. 2019, 83, 29–36. [Google Scholar] [CrossRef] [Green Version]
  17. Ivanova, E.P.; Hasan, J.; Webb, H.K.; Truong, V.K.; Watson, G.S.; Watson, J.A.; Baulin, V.A.; Pogodin, S.; Wang, J.Y.; Tobin, M.J.; et al. Natural bactericidal surfaces: Mechanical rupture of Pseudomonas aeruginosa cells by cicada wings. Small 2012, 8, 2489–2494. [Google Scholar] [CrossRef] [PubMed]
  18. Bandara, C.D.; Singh, S.; Afara, I.O.; Wolff, A.; Tesfamichael, T.; Ostrikov, K.; Oloyede, A. Bactericidal Effects of Natural Nanotopography of Dragonfly Wing on Escherichia coli. ACS Appl. Mater. Interfaces 2017, 9, 6746–6760. [Google Scholar] [CrossRef] [Green Version]
  19. Linklater, D.P.; Juodkazis, S.; Rubanov, S.; Ivanova, E.P. Comment on “Bactericidal Effects of Natural Nanotopography of Dragonfly Wing on Escherichia coli”. ACS Appl. Mater. Interfaces 2017, 9, 29387–29393. [Google Scholar] [CrossRef] [Green Version]
  20. Zhou, J.; Li, B.; Zhao, L.; Zhang, L.; Han, Y. F-doped micropore/nanorod hierarchically patterned coatings for improving antibacterial and osteogenic activities of bone implants in bacteria-infected cases. ACS Biomater. Sci. Eng. 2017, 3, 1437–1450. [Google Scholar] [CrossRef] [PubMed]
  21. Zhang, H.Z.; Sun, Y.; Tian, A.; Xue, X.X.; Wang, L.; Alquhali, A.; Bai, X.Z. Improved antibacterial activity and biocompatibility on vancomycin-loaded TiO2 nanotubes: In vivo and in vitro studies. Int. J. Nanomed. 2013, 8, 4379–4389. [Google Scholar] [CrossRef] [Green Version]
  22. Guan, M.; Chen, Y.; Wei, Y.; Song, H.; Gao, C.; Cheng, H.; Li, Y.; Huo, K.; Fu, J.; Xiong, W. Long-lasting bactericidal activity through selective physical puncture and controlled ions release of polydopamine and silver nanoparticles-loaded TiO2 nanorods in vitro and in vivo. Int. J. Nanomed. 2019, 14, 2903–2914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Yang, Y.; Liu, L.; Luo, H.; Zhang, D.; Lei, S.; Zhou, K. Dual-purpose magnesium-incorporated titanium nanotubes for combating bacterial infection and ameliorating osteolysis to realize better osseointegration. ACS Biomater. Sci. Eng. 2019, 5, 5368–5383. [Google Scholar] [CrossRef]
  24. Zhang, X.; Zhang, G.; Chai, M.; Yao, X.; Chen, W.; Chu, P.K. Synergistic antibacterial activity of physical-chemical multi-mechanism by TiO2 nanorod arrays for safe biofilm eradication on implant. Bioactive Mater. 2021, 6, 12–25. [Google Scholar] [CrossRef]
  25. Hooijmans, C.R.; Rovers, M.M.; de Vries, R.B.M.; Leenaars, M.; Ritskes-Hoitinga, M.; Langendam, M.W. SYRCLE’s risk of bias tool for animal studies. BMC Med. Res. Methodol. 2014, 14, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Percie du Sert, N.; Hurst, V.; Ahluwalia, A.; Alam, S.; Avey, M.T.; Baker, M.; Browne, W.J.; Clark, A.; Cuthill, I.C.; Dirnagl, U.; et al. The ARRIVE guidelines 2.0: Updated guidelines for reporting animal research. Br. J. Pharmacol. 2020, 177, 3617–3624. [Google Scholar] [CrossRef] [PubMed]
  27. López-Valverde, N.; Macedo-de-Sousa, B.; López-Valverde, A.; Ramírez, J.M. Effectiveness of Antibacterial Surfaces in Osseointegration of Titanium Dental Implants: A Systematic Review. Antibiotics 2021, 10, 360. [Google Scholar] [CrossRef] [PubMed]
  28. Diu, T.; Faruqui, N.; Sjöström, T.; Lamarre, B.; Jenkinson, H.F.; Su, B.; Ryadnov, M.G. Cicada-inspired cell-instructive nanopatterned arrays. Sci. Rep. 2014, 4, 7122. [Google Scholar] [CrossRef] [Green Version]
  29. Kelleher, S.M.; Habimana, O.; Lawler, J.; O’reilly, B.; Daniels, S.; Casey, E.; Cowley, A. Cicada wing surface topography: An investigation into the bactericidal properties of nanostructural features. ACS Appl. Mater. Interfaces 2016, 8, 14966–14974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Bürgers, R.; Gerlach, T.; Hahnel, S.; Schwarz, F.; Handel, G.; Gosau, M. In vivo and in vitro biofilm formation on two different titanium implant surfaces. Clin. Oral. Implants Res. 2010, 21, 156–164. [Google Scholar] [CrossRef]
  31. Crawford, R.J.; Webb, H.K.; Truong, V.K.; Hasan, J.; Ivanova, E.P. Surface topographical factors influencing bacterial attachment. Adv. Colloid Interface Sci. 2012, 179–182, 142–149. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, B.; Luo, Y.; Pearlstein, A.J.; Aplin, J.; Liu, Y.; Bauchan, G.R.; Payne, G.F.; Wang, Q.; Nou, X.; Millner, P.D. Fabrication of biomimetically patterned surfaces and their application to probing plant-bacteria interactions. ACS Appl. Mater. Interfaces 2014, 6, 12467–12478. [Google Scholar] [CrossRef] [PubMed]
  33. Vasudevan, R.; Kennedy, A.J.; Merritt, M.; Crocker, F.H.; Baney, R.H. Microscale patterned surfaces reduce bacterial fouling-microscopic and theoretical analysis. Coll. Surf. B Biointerfaces 2014, 117, 225–232. [Google Scholar] [CrossRef]
  34. Ge, X.; Leng, Y.; Lu, X.; Ren, F.; Wang, K.; Ding, Y.; Yang, M. Bacterial responses to periodic micropillar array. J. Biomed. Mater. Res. A 2015, 103, 384–396. [Google Scholar] [CrossRef]
  35. Valle, J.; Burgui, S.; Langheinrich, D.; Gil, C.; Solano, C.; Toledo-Arana, A.; Helbig, R.; Lasagni, A.; Lasa, I. Evaluation of surface microtopography engineered by direct laser interference for bacterial anti-biofouling. Macromol. Biosci. 2015, 15, 1060–1069. [Google Scholar] [CrossRef] [Green Version]
  36. Liu, W.; Su, P.; Gonzales, A., 3rd; Chen, S.; Wang, N.; Wang, J.; Li, H.; Zhang, Z.; Webster, T.J. Optimizing stem cell functions and antibacterial properties of TiO2 nanotubes incorporated with ZnO nanoparticles: Experiments and modeling. Int. J. NanoMed. 2015, 10, 1997–2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Gunputh, U.F.; Le, H.; Handy, R.D.; Tredwin, C. Anodised TiO2 nanotubes as a scaffold for antibacterial silver nanoparticles on titanium implants. Mater. Sci. Eng. C Mater. Biol. Appl. 2018, 91, 638–644. [Google Scholar] [CrossRef] [PubMed]
  38. Cheng, Y.F.; Zhang, J.Y.; Wang, Y.B.; Li, C.M.; Lu, Z.S.; Hu, X.F.; Xu, L.Q. Deposition of catechol-functionalized chitosan and silver nanoparticles on biomedical titanium surfaces for antibacterial application. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 98, 649–656. [Google Scholar] [CrossRef]
  39. Ahmed, K.B.A.; Raman, T.; Veerappan, A. Future prospects of antibacterial metal nanoparticles as enzyme inhibitor. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 68, 939–947. [Google Scholar] [CrossRef]
  40. Kheiri, S.; Liu, X.; Thompson, M. Nanoparticles at biointerfaces: Antibacterial activity and nanotoxicology. Coll. Surf. B Biointerfaces 2019, 184, 110550. [Google Scholar] [CrossRef]
  41. Park, H.J.; Kim, J.Y.; Kim, J.; Lee, J.H.; Hahn, J.S.; Gu, M.B.; Yoon, J. Silver-ion-mediated reactive oxygen species generation affecting bactericidal activity. Water Res. 2009, 43, 1027–1032. [Google Scholar] [CrossRef] [PubMed]
  42. Vimbela, G.V.; Ngo, S.M.; Fraze, C.; Yang, L.; Stout, D.A. Antibacterial properties and toxicity from metallic nanomaterials. Int. J. Nanomed. 2017, 12, 3941–3965. [Google Scholar] [CrossRef] [Green Version]
  43. Li, J.; Tan, L.; Liu, X.; Cui, Z.; Yang, X.; Yeung, K.W.K.; Chu, P.K.; Wu, S. Balancing bacteria-osteoblast competition through selective physical puncture and biofunctionalization of ZnO/polydopamine/arginine-glycine-aspartic acid-cysteine nanorods. ACS Nano 2017, 11, 11250–11263. [Google Scholar] [CrossRef] [PubMed]
  44. Anitha, S.; Muthukumaran, S. Structural, optical and antibacterial investigation of La, Cu dual doped ZnO nanoparticles prepared by co-precipitation method. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 108, 110387. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, L.; Zhang, L.; Yang, Y.; Zhang, W.; Lv, H.; Yang, F.; Lin, C.; Tang, P. Inhibitory effect of super-hydrophobicity on silver release and antibacterial properties of super-hydrophobic Ag/TiO2 nanotubes. J. Biomed. Mater. Res. B Appl. Biomater. 2016, 104, 1004–1012. [Google Scholar] [CrossRef]
  46. Yang, Z.; Gu, H.; Sha, G.; Lu, W.; Yu, W.; Zhang, W.; Fu, Y.; Wang, K.; Wang, L. TC4/Ag metal matrix nanocomposites modified by friction stir processing: Surface characterization, antibacterial property, and cytotoxicity in vitro. ACS Appl. Mater. Interfaces 2018, 10, 41155–41166. [Google Scholar] [CrossRef]
  47. Gunputh, U.F.; Le, H.; Lawton, K.; Besinis, A.; Tredwin, C.; Handy, R.D. Antibacterial properties of silver nanoparticles grown in situ and anchored to titanium dioxide nanotubes on titanium implant against Staphylococcus aureus. Nanotoxicology 2020, 14, 97–110. [Google Scholar] [CrossRef]
  48. Deshmukh, S.P.; Mullani, S.B.; Koli, V.B.; Patil, S.M.; Kasabe, P.J.; Dandge, P.B.; Pawar, S.A.; Delekar, S.D. Ag nanoparticles connected to the surface of TiO2 electrostatically for antibacterial photoinactivation studies. Photochem. Photobiol. 2018, 94, 1249–1262. [Google Scholar] [CrossRef]
  49. Bala Subramaniyan, S.; Megarajan, S.; Vijayakumar, S.; Mariappan, M.; Anbazhagan, V. Evaluation of the toxicities of silver and silver sulfide nanoparticles against Gram-positive and Gram-negative bacteria. IET Nanobiotechnol. 2019, 13, 326–331. [Google Scholar] [CrossRef]
  50. Mohamed, D.S.; Abd El-Baky, R.M.; Sandle, T.; Mandour, S.A.; Ahmed, E.F. Antimicrobial activity of silver-treated bacteria against other multi-drug resistant pathogens in their environment. Antibiotics 2020, 9, 181. [Google Scholar] [CrossRef] [Green Version]
  51. Yang, Z.; Ma, C.; Wang, W.; Zhang, M.; Hao, X.; Chen, S. Fabrication of Cu2O-Ag nanocomposites with enhanced durability and bactericidal activity. J. Colloid. Interface Sci. 2019, 557, 156–167. [Google Scholar] [CrossRef]
  52. Lai, Y.; Dong, L.; Zhou, H.; Yan, B.; Chen, Y.; Cai, Y.; Liu, J. Coexposed nanoparticulate Ag alleviates the acute toxicity induced by ionic Ag+ in vivo. Sci. Total Environ. 2020, 723, 138050. [Google Scholar] [CrossRef]
  53. 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]
  54. Gao, C.; Cheng, H.; Xu, N.; Li, Y.; Chen, Y.; Wei, Y.; Gao, B.; Fu, J.; Huo, K.; Xiong, W. Poly(dopamine) and Ag nanoparticle-loaded TiO2 nanotubes with optimized antibacterial and ROS-scavenging bioactivities. Nanomedicine 2019, 14, 803–818. [Google Scholar] [CrossRef] [PubMed]
  55. Zhou, J.; Li, B.; Han, Y. F-doped TiO2 microporous coating on titanium with enhanced antibacterial and osteogenic activities. Sci. Rep. 2018, 8, 17858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Tran, P.A.; O’Brien-Simpson, N.; Palmer, J.A.; Bock, N.; Reynolds, E.C.; Webster, T.J.; Deva, A.; Morrison, W.A.; O’Connor, A.J. Selenium nanoparticles as anti-infective implant coatings for trauma orthopedics against methicillin-resistant Staphylococcus aureus and epidermidis: In vitro and in vivo assessment. Int. J. Nanomed. 2019, 14, 4613–4624. [Google Scholar] [CrossRef] [Green Version]
  57. Yonezawa, K.; Kawaguchi, M.; Kaneuji, A.; Ichiseki, T.; Iinuma, Y.; Kawamura, K.; Shintani, K.; Oda, S.; Taki, M.; Kawahara, N. Evaluation of antibacterial and cytotoxic properties of a fluorinated diamond-like carbon coating for the development of antibacterial medical implants. Antibiotics 2020, 9, 495. [Google Scholar] [CrossRef]
  58. Wang, Y.; Zhang, S.; Zeng, X.; Ma, L.L.; Weng, W.; Yan, W.; Qian, M. Osteoblastic cell response on fluoridated hydroxyapatite coatings. Acta Biomater. 2007, 3, 191–197. [Google Scholar] [CrossRef]
  59. Zhou, J.; Zhao, L. Multifunction Sr, Co and F co-doped microporous coating on titanium of antibacterial, angiogenic and osteogenic activities. Sci. Rep. 2016, 6, 29069. [Google Scholar] [CrossRef]
  60. Li, D.; Lv, P.; Fan, L.; Huang, Y.; Yang, F.; Mei, X.; Wu, D. The immobilization of antibiotic-loaded polymeric coatings on osteoarticular Ti implants for the prevention of bone infections. Biomater. Sci. 2017, 5, 2337–2346. [Google Scholar] [CrossRef] [PubMed]
  61. Mohan Raj, R.; Priya, P.; Raj, V. Gentamicin-loaded ceramic-biopolymer dual layer coatings on the Ti with improved bioactive and corrosion resistance properties for orthopedic applications. J. Mech. Behav. Biomed. Mater. 2018, 82, 299–309. [Google Scholar] [CrossRef] [PubMed]
  62. Baghdan, E.; Raschpichler, M.; Lutfi, W.; Pinnapireddy, S.R.; Pourasghar, M.; Schäfer, J.; Schneider, M.; Bakowsky, U. Nano spray dried antibacterial coatings for dental implants. Eur. J. Pharm. Biopharm. 2019, 139, 59–67. [Google Scholar] [CrossRef]
  63. 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. [Google Scholar] [CrossRef] [PubMed]
  64. Salwiczek, M.; Qu, Y.; Gardiner, J.; Strugnell, R.A.; Lithgow, T.; McLean, K.M.; Thissen, H. Emerging rules for effective antimicrobial coatings. Trends Biotechnol. 2014, 32, 82–90. [Google Scholar] [CrossRef]
  65. Acosta, S.; Ibañez-Fonseca, A.; Aparicio, C.; Rodríguez-Cabello, J.C. Antibiofilm coatings based on protein-engineered polymers and antimicrobial peptides for preventing implant-associated infections. Biomater. Sci. 2020, 8, 2866–2877. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, S.H.; Tang, T.W.; Wu, E.; Wang, D.W.; Liao, Y.D. Anionic surfactant-facilitated coating of antimicrobial peptide and antibiotic reduces biomaterial-associated infection. ACS Biomater. Sci. Eng. 2020, 6, 4561–4572. [Google Scholar] [CrossRef]
  67. Lipsky, B.A.; Berendt, A.R.; Cornia, P.B.; Pile, J.C.; Peters, E.J.; Armstrong, D.G.; Deery, H.G.; Embil, J.M.; Joseph, W.S.; Karchmer, A.W.; et al. Infectious Diseases Society of America clinical practice guideline for the diagnosis and treatment of diabetic foot infections. Clin. Infect. Dis. 2012, 54, e132–e173. [Google Scholar] [CrossRef] [Green Version]
  68. Tan, T.L.; Kheir, M.M.; Tan, D.D.; Parvizi, J. Polymicrobial periprosthetic joint infections: Outcome of treatment and identification of risk factors. J. Bone Joint Surg. Am. 2016, 98, 2082–2088. [Google Scholar] [CrossRef]
  69. Puca, V.; Marulli, R.Z.; Grande, R.; Vitale, I.; Niro, A.; Molinaro, G.; Prezioso, S.; Muraro, R.; Di Giovanni, P. Microbial species isolated from infected wounds and antimicrobial resistance analysis: Data emerging from a three-years retrospective study. Antibiotics 2021, 10, 1162. [Google Scholar] [CrossRef]
  70. Simchi, A.; Tamjid, E.; Pishbin, F.; Boccaccini, A.R. Recent progress in inorganic and composite coatings with bactericidal capability for orthopaedic applications. Nanomedicine 2011, 7, 22–39. [Google Scholar] [CrossRef]
  71. Hickok, N.J.; Shapiro, I.M. Immobilized antibiotics to prevent orthopaedic implant infections. Adv. Drug. Deliv. Rev. 2012, 64, 1165–1176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Nguyen-Tri, P.; Tran, H.N.; Plamondon, C.O.; Tuduri, L.; Vo, D.-V.N.; Nanda, S.; Mishra, A.; Chao, H.-P.; Bajpai, A.K. Recent progress in the preparation, properties and applications of superhydrophobic nano-based coatings and surfaces: A review. Prog. Org. Coat. 2019, 132, 235–256. [Google Scholar] [CrossRef]
  73. Muthukrishnan, G.; Masters, E.A.; Daiss, J.L.; Schwarz, E.M. Mechanisms of immune evasion and bone tissue colonization that make Staphylococcus aureus the primary pathogen in Osteomyelitis. Curr. Osteoporos Rep. 2019, 17, 395–404. [Google Scholar] [CrossRef]
  74. Kavanagh, N.; Ryan, E.J.; Widaa, A.; Sexton, G.; Fennell, J.; O’Rourke, S.; Cahill, K.C.; Kearney, C.J.; O’Brien, F.J.; Kerrigan, S.W. Staphylococcal Osteomyelitis: Disease progression, treatment challenges, and future directions. Clin. Microbiol. Rev. 2018, 31, e00084-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Anagnostakos, K.; Sahan, I. Are cement spacers and beads loaded with the correct antibiotic(s) at the site of periprosthetic hip and knee joint infections? Antibiotics 2021, 10, 143. [Google Scholar] [CrossRef]
  76. Moher, D.; Liberati, A.; Tetzlaff, J.; Altman, D.G. Preferred reporting items for systematic reviews and meta-analyses: The PRISMA statement. PLoS Med. 2009, 6, e1000097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. PRISMA (Preferred Reporting Items for Systematic Reviews and Meta-Analyses) diagram including study algorithm.
Figure 1. PRISMA (Preferred Reporting Items for Systematic Reviews and Meta-Analyses) diagram including study algorithm.
Antibiotics 10 01524 g001
Figure 2. Risk of bias assessment results based on SYRCLE’s (Systematic Review Centre for Laboratory Animal Experimentation) risk of bias tool. White indicates low risk; black high risk; and gray unclear risk [20,21,22,23,24].
Figure 2. Risk of bias assessment results based on SYRCLE’s (Systematic Review Centre for Laboratory Animal Experimentation) risk of bias tool. White indicates low risk; black high risk; and gray unclear risk [20,21,22,23,24].
Antibiotics 10 01524 g002
Table 1. Study characteristics.
Table 1. Study characteristics.
StudiesAnimal Model (n)Location of Implant PlacementBacteria and Infection Set-UpFollow-UpIn Vivo Antibacterial Efficacy MeasuresIn Vivo Antibacterial Efficacy Conclusions (Nanopatterns)
Zhang et al., 2013 [21]Sprague Dawley rat (36)Femoral intercondylar fossaS. aureus. bacterial suspension (107 CFU/100μL, 100 μL) was introduced into the femoral canal through the hole in the femoral intercondylar fossa.30 daysClinical assessment
Bacterial Culture
Although better than the pure Ti, TiO2 NTs showed a poor antibacterial effect in vivo.
Zhou et al., 2017 [20]New Zealand rabbit (24)Left femurPBS-diluted suspension of S. aureus (105 CFU/mL, 20 μL) was injected into the medullary cavity of the femur.8 weeksBacterial Culture
CFU Counting
The Sr1-HA NRDs on microporous TiO2 showed no antibacterial activity in vivo.
Guan et al., 2019 [22]Sprague Dawley rat (20)Tibia plateau of the right knee30 μL bacteria suspension (MRSA, 1.5 × 106 CFU/mL) was injected into the exposed tibia hole.4 weeksX-ray
Micro-CT
Histopathological analysis
Same as the pure Ti group, TiO2 NRDs did not show antibacterial activity in vivo.
Yang et al., 2019 [23]Sprague Dawley rat (45)Femoral medullary cavity at the middle of intercondylar fossa50 μL of PBS containing MRSA at a 1 × 106 CFU/mL concentration was injected into the medullary cavity.5 weeksX-ray
Micro-CT
Histopathological analysis
The NT structure
itself demonstrated limited antimicrobial activities in vivo.
Zhang et al., 2021 [24]Kunming mice (24)Subcutaneous tissue on the back and tibiaAntibacterial assay: samples were soaked in 50 μL of S. aureus (1 × 107 CFU/mL) for 1 h and then implanted.
Antibiofilm assay: samples were cultivated in 2 mL of the S. aureus solution (107 CFU/ mL) to form biofilms.
2–14 daysBacterial Culture
Histopathological analysis
TiO2 nanostructures under the irradiation of 808 nm NIR light had an excellent anti-biofilm effect in vivo.
CFU: Colony-Forming Unit; HA: hydroxyapatite; S. aureus: Staphylococcus aureus; MRSA: methicillin-resistant Staphylococcus aureus; NRDs: nanorods; NTs: nanotubes; NIR: near-infrared; Sr: strontium; Ti: titanium.
Table 2. Implant characteristics and in vivo antibacterial activities (Outcomes).
Table 2. Implant characteristics and in vivo antibacterial activities (Outcomes).
StudiesImplants Number (n)Implant Dimensions D(Ø) × L (mm)Ti Implant ShapeSurface NanopatternsNanopattern Dimensions D (Ø) × L (nm)In Vivo Antibacterial Activities (Nanopatterns)
Zhang et al., 2013 [21]361 (Ø) × L 20CylinderTiO2 NTs80 (Ø) × L 800The infection rate was lower in the NT group compared to the Cp-Ti group (92% vs. 100%).
Zhou et al., 2017 [20]1202 (Ø) × L 10CylinderSr1-HA NRDs on microporous TiO2NRDs Ø and interrod spacing: 70 ± 6
Pore Ø: 1000-3000
The average CFU counting results exhibited the Sr1-HA NRDs on microporous TiO2 did not possess antibacterial activity against S. aureus in vivo.
Guan et al., 2019 [22]201 (Ø) × L 10CylinderTiO2 NRDs50–100 (Ø) × L 1000–2000TiO2 NRDs showed no difference in the infection rate compared to the pure Ti group.
Yang et al., 2019 [23]452 (Ø) × L 15CylinderTiO2 NTs80 (Ø)The NT structure itself exhibited slight anti-infection potential in vivo, but the NTs structure alone could not combat such a severe implant-related bone infection.
Zhang et al., 2021 [24]+10 (length) × (wide) 5FoilTiO2 NRDsNanoleaf;
NRDs: 40–50 (Ø) × L 1000
The TiO2 NRDs arrays under irradiation with 808 NIR light produced excellent antibacterial activity in vivo and could eradicate the attached biofilms on the implant surface.
CFU: Colony-Forming Unit; Cp-Ti: commercially pure titanium; HA: hydroxyapatite; NIR: near-infrared; PBS: phosphate buffered saline; Sr: strontium; S. aureus: Staphylococcus aureus; NRDs: nanorods; NTs: nanotubes; Ti: titanium.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sun, Y.; Yang, Y.; Jiang, W.; Bai, H.; Liu, H.; Wang, J. In Vivo Antibacterial Efficacy of Nanopatterns on Titanium Implant Surface: A Systematic Review of the Literature. Antibiotics 2021, 10, 1524. https://doi.org/10.3390/antibiotics10121524

AMA Style

Sun Y, Yang Y, Jiang W, Bai H, Liu H, Wang J. In Vivo Antibacterial Efficacy of Nanopatterns on Titanium Implant Surface: A Systematic Review of the Literature. Antibiotics. 2021; 10(12):1524. https://doi.org/10.3390/antibiotics10121524

Chicago/Turabian Style

Sun, Yang, Yang Yang, Weibo Jiang, Haotian Bai, He Liu, and Jincheng Wang. 2021. "In Vivo Antibacterial Efficacy of Nanopatterns on Titanium Implant Surface: A Systematic Review of the Literature" Antibiotics 10, no. 12: 1524. https://doi.org/10.3390/antibiotics10121524

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