Next Article in Journal
Efficient and Stable Air-Processed Ternary Organic Solar Cells Incorporating Gallium-Porphyrin as an Electron Cascade Material
Previous Article in Journal
The Tunable Electronic and Optical Properties of Two-Dimensional Bismuth Oxyhalides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Progress in Nanostructured Mechano-Bactericidal Polymeric Surfaces for Biomedical Applications

by
S. P. S. N. Buddhika Sampath Kumara
1,2,3,
S. W. M. Amal Ishantha Senevirathne
1,3,
Asha Mathew
1,4,
Laura Bray
1,3,
Mohammad Mirkhalaf
1,3,5,* and
Prasad K. D. V. Yarlagadda
1,2,3,4,*
1
School of Mechanical, Medical and Process Engineering, Faculty of Engineering, Queensland University of Technology (QUT), Brisbane, QLD 4000, Australia
2
Australian Research Council Training Centre for Multiscale 3D Imaging, Modelling, and Manufacturing, Queensland University of Technology (QUT), Brisbane, QLD 4000, Australia
3
Centre for Biomedical Technologies, Queensland University of Technology (QUT), Brisbane, QLD 4000, Australia
4
School of Engineering, University of Southern Queensland, Springfield, QLD 4300, Australia
5
Centre for Materials Science, Queensland University of Technology (QUT), Brisbane, QLD 4000, Australia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(20), 2799; https://doi.org/10.3390/nano13202799
Submission received: 30 September 2023 / Revised: 16 October 2023 / Accepted: 17 October 2023 / Published: 20 October 2023

Abstract

:
Bacterial infections and antibiotic resistance remain significant contributors to morbidity and mortality worldwide. Despite recent advances in biomedical research, a substantial number of medical devices and implants continue to be plagued by bacterial colonisation, resulting in severe consequences, including fatalities. The development of nanostructured surfaces with mechano-bactericidal properties has emerged as a promising solution to this problem. These surfaces employ a mechanical rupturing mechanism to lyse bacterial cells, effectively halting subsequent biofilm formation on various materials and, ultimately, thwarting bacterial infections. This review delves into the prevailing research progress within the realm of nanostructured mechano-bactericidal polymeric surfaces. It also investigates the diverse fabrication methods for developing nanostructured polymeric surfaces with mechano-bactericidal properties. We then discuss the significant challenges associated with each approach and identify research gaps that warrant exploration in future studies, emphasizing the potential for polymeric implants to leverage their distinct physical, chemical, and mechanical properties over traditional materials like metals.

1. Introduction

Bacterial cells usually attach to surfaces, colonise, and form a matrix of polysaccharide material called biofilms. Bacteria can adhere to many different surfaces, including human tissues, metals, and polymeric surfaces [1]. The bacteria that form biofilms are highly resistant to disinfectants, antiseptics, and other antimicrobial agents that are normally effective in their planktonic state [2,3,4]. Moreover, overuse of antibiotics has led to an era where bacteria are becoming increasingly resistant to antibacterial agents, making it more difficult to treat infections [4,5]. The discovery of the mechano-bactericidal response to nano-topography found on insect wings, such as cicadas and dragonflies, has inspired the development of nanostructured topography on synthetic materials, which can reduce or inhibit the growth of bacteria on their surfaces. Most of the studies on artificial antibacterial micro/nanostructured surfaces were based on materials like ceramics, metals, and alloys, because of their superior mechanical and thermal properties that lead to potential applications in medical and industrial sectors. Metals are among the most used engineering materials and are employed extensively as biomaterials [6,7]. However, they can cause inflammation, bone loss, and imaging artifacts in computed tomography (CT) scans and magnetic resonance imaging (MRI) due to their magnetic properties, low flexibility, and low biocompatibility compared to biocompatible polymers [8,9]. High cost and density characteristics also make them less effective in specific medical implants such as cardiovascular implants, absorbable implants, paediatric implants, and cochlear implants [9,10,11,12,13,14,15]. Biocompatible polymers are attractive alternatives that can be used in such applications due to ease of processing, low weight, high biocompatibility, and biodegradability [16,17,18,19]. The methods for developing nanostructured bactericidal surfaces have been reviewed in the recent literature [20,21,22,23,24]. However, a comprehensive and critical overview of the techniques developed for polymers is yet missing. This review provides such an overview showing the progress in the field, identifying research gaps, and discussing opportunities and possible future directions.

2. Polymers Used in Biomedical Applications

Many biocompatible polymers are used in medical implants with some demonstrating high biocompatibility [25]. The most commonly used biocompatible polymers are polyglycolic acid (PGA) [26,27], poly(lactic-co-glycolic acid) (PLGA) [26,28,29,30], polycaprolactone (PCL) [26,31,32], polyurethane (PU) [33], polyvinyl alcohol (PVA) [34,35,36], silicone [33,37,38], polylactic acid (PLA) [26,39,40,41,42,43,44], polypropylene (PP) [45,46], and polymethyl methacrylate (PMMA) [47,48,49]. Each polymer has distinct advantages and limitations in various biomedical applications, as mentioned in Table 1. For instance, PMMA is a frequently used polymer for dental implants due to its non-degradability and high strength while PLA is used in bone tissue engineering applications due to its biodegradability and osteoblast stimulation [39,50]. Moreover, polyetheretherketone (PEEK) and polyetherketoneketone (PEKK) are emerging biomedical materials with high biocompatibility, thermal stability, and non-degradability with excellent mechanical properties which are suitable for various biomedical applications [51,52].
In such applications, one of the most concerning issues is implant failures due to bacterial infections as described in the Introduction section. Bacteria colonise various implant surfaces, such as dental implants, catheters, and orthopaedic implants [53]. Furthermore, antibiotic resistance directly contributes to implant failures due to the formation of biofilms on these surfaces [54,55]. In response to this concern, numerous studies have been undertaken to develop physical and chemical methods aimed at reducing bacterial colonisation [56,57]. To implement these strategies successfully, the choice of implant material plays a pivotal role in developing surfaces that are bactericidal and/or antifouling. Among various materials, polymers offer advantageous characteristics for biomedical implant applications. These attributes are highly considered in the biomedical field, particularly for ensuring a successful implant surgery in in vivo conditions (Figure 1).
Table 1. Polymers used in biomedical applications. Table abbreviations: Polydimethylsiloxane—PDMS; Polyethylene terephthalate—PET; Polytetrafluoroethylene—PTFE.
Table 1. Polymers used in biomedical applications. Table abbreviations: Polydimethylsiloxane—PDMS; Polyethylene terephthalate—PET; Polytetrafluoroethylene—PTFE.
PolymerCharacteristics, Physical, and Mechanical PropertiesBiomedical
Applications
AdvantagesLimitationsRef.
PGABiodegradable, biocompatible, tensile stress: 890 MPa, density: 1.5 g/cm3, melting point: 225–230 °C, glass transition temperature: 35–40 °CTissue engineering applications in bone, tendon, cartilage, tooth, and spinal regeneration; nerve grafts; absorbable suturesStimulates cartilage regeneration; 3D-printability; high tensile strengthHigh friction coefficient and “binds and snags” when wet, high brittleness, high degradation causes inflammatory response[26,27,58,59,60]
PLGABiodegradable, bioadsorbable, biocompatible, tensile stress: 3.4 MPa, density: 1.2 g/cm3, melting point: dependent on the percent composition (PLA, PGA), glass transition temperature: 40–60 °CTherapeutic tools; drug delivery; tissue engineeringStimulates osteoblasts; 3D-printabilityRelease of acidic byproducts leads to inflammation, degrade due to hydrolysis, poor strength[26,28,29,30,61]
PCLBiodegradable, bioadsorbable, biocompatible, tensile stress: 12.8 MPa, density: 1.15 g/cm3, melting point: 60 °C, glass transition temperature: −60 °CDental splints; drug delivery; tissue engineeringStimulates osteoblasts; 3D printable; slow degradation rate; low cost in 3D printing due to low melting point; high biocompatibilityPoor mechanical properties; low cell adhesion[26,31,62]
PUCan be biodegradable or non-biodegradable based on chemical composition, non-bioabsorbable, biocompatible, tensile strength: 34.5–56 MPa, density: 1.23 g/cm3, melting point: 163 °C, glass transition temperature: −35 °CDrug delivery; catheters, pacemaker leads insulation, vascular prostheses, heart valves, cardiac assist devices (cardiovascular applications)High durability; high toughness; good biostability; low costEnvironmental stress cracking; material degradation in vivo; metal ion oxidation[16,63,64,65]
PPNon-biodegradable, non-bioabsorbable, biocompatible, tensile stress: 28 MPa, density: 0.9 g/cm3, melting point: 170 °C, glass transition temperature: −25 °CSutures; scaffolds (ligament or tendon repair); meshes for hernia and pelvic organ repair; heart valve structure,
oxygenator and plasmapheresis membranes,
finger joint prosthesis
High melting point; less toxic; low costLimited biocompatibility; poor strength[16,33,66,67,68]
PVABiodegradable, biocompatible, tensile stress: 40–90 MPa, density: 1.26 g/cm3, melting point: 228 °C, glass transition temperature: 85 °CWound dressings, drug delivery, targeted-tissue transportation systems; soft biomaterial implants.High chemical and thermal stability; non-toxicWeak hydrogel endurance in high temperature; relatively weak polymer; limited biocompatibility; degrades due to hydrolysis[34,69,70,71,72]
Silicone or PDMSNon-absorbable, non-biodegradable, biocompatible, hydrophobic, tensile stress: 2–10 MPa, density: 0.97 g/cm3, melting point: 228 °C
glass transition temperature: ~120–123 °C
Oxygenator membrane; tubing; shunts; prostheses; heart peacemaker leads; heart valve structures; burn dressingChemically inert; low toxicity; thermal stability; high biocompatibilityProne to damage; non-durable; contamination of monomers; low mechanical strength[16,47,63]
PLABiodegradable, bioabsorbable, biocompatible, tensile stress: 21–60 MPa, density: 1.21–1.25 g/cm3, melting point: 150–160 °C, glass transition temperature: 60–65 °CBone tissue engineering; drug delivery; plates, screws, pins, and wires in bone fixation; bio-absorbable implants; sutures in dermatology; drug-eluting stentsHigh biocompatibility; stimulates osteoblasts; less brittle; one of the highly used 3D-printable materials; degradation products are also non-toxic to humans and the environment.Low mechanical strength[26,33,39,40,66,67,73]
PMMANon-degradable, biocompatible, tensile stress: 48–76 MPa, density: 1.2 g/cm3, melting point: 130–180 °C, glass transition temperature: 80 °CDental implants; bone cement; lenses; drug deliveryOne of the hardest thermoplastics with high scratch resistance; high mechanical strengthLess biocompatibility; high curing temperature; does not support osteointegration; causes necrosis effect[16,74,75]
PEEKNon-degradable, biocompatible, tensile stress: 84 MPa, density: 1.4 g/cm3, melting point: 343 °C, glass transition temperature: 143 °CDental implants; knee implants; spine implants; cranioplasticity; hip replacement; anterior plate fixation; heart valves; face reconstructionsHigh biocompatibility; 3D-printable; light weight; compatible with hydroxyapatite (natural bone tissue materials) hence substitute to metallic implants; stable at high temperatures; mechanical stabilityLow thermoformability; bioinert (does not promote tissue integration); complex and costly manufacturing process[76,77,78]
PEKKNon-degradable, biocompatible, tensile stress: 115 MPa, density: 1.3 g/cm3, melting point: 363–386 °C, glass transition temperature: 162 °CDental implants; crown and bridge in dentistry; endodontic post; removable denture framework; restorative and prosthetic applicationsHigh biocompatibility; 3D-printable; light weight; high mechanical strength; excellent chemical resistanceBioinert (does not promote tissue integration); more complex and costly manufacturing process than PEEK[52,79]
PETNon-degradable, high biocompatibility, tensile stress: 75–100 MPa, density:1.38 g/cm3, melting point: 255–265 °C, glass transition temperature: 85 °CSutures; heart valves; surgical meshes; scaffolds; urinary and bloodstream catheters; commercial vascular prosthesis3D-printable; cost effective; excellent chemical resistanceBioinert (does not promote tissue integration)[80,81,82]
PTFENon-degradable, biocompatible, tensile stress: 30.5 MPa, density: 2.175 g/cm3, melting point: 327 °C, glass transition temperature: 127 °CVascular graft prostheses; heart patches; stapes prosthesisHigh mechanical strength; chemically inertDifficult to 3D-print[16,83]
ChitosanBiodegradable, biocompatible, tensile stress: 32.2 MPa, density: 0.20–0.38 g/cm3, melting point:105 °C, glass transition temperature: 75 °CAntitumor drug delivery; protein and peptide drug delivery; gene delivery; antibiotic delivery; polyphenol delivery; wound healing applicationsAntimicrobial; anti-inflammatory; antifungal; nontoxicity; antitumor activity; antioxidant activityLow mechanical strength; significant variations of properties based on the source of material[84,85,86,87,88,89,90]

3. Mechanism of Bacteria Adhesion on Surfaces

Bacteria are present in different environments, such as animals, soil, plants, fresh water, and air [91]. Bacterial adhesion refers to the capability of bacteria to attach to a range of surfaces, such as human tissues, medical implants, polymers, metals, and glasses [92]. This process is vital for bacterial colonisation and is achieved through a complex process that involves few stages with multiple factors. As an overview, it is affected by the distinct characteristics of bacteria such as motility, cell wall structure and appendages (flagella, pili, and curli), exposure duration to the surface, amount of nutrients, coaggregation, cohesion, and bacterial density [92,93,94]. Bacteria colonisation occurs in two stages: (1) prior attachment stage, known as the primary stage or reversible adhesion, and (2) post-attachment stage, known as the secondary stage or irreversible adhesion [95]. During the reversible (volatile) adhesion, once the bacteria reach a certain proximity from the surface, their adhesion depends on the superposition of attractive and/or repulsive forces such as electrostatic, van der Waals, hydrophobic, and hydrodynamic forces [95], whereas the attraction of the bacteria to the surface is high during this stage and occurs within few minutes [96]. The majority of bacteria carry a predominantly negative surface charge [97,98], especially in the initial growth stage [98], and tend to selectively bond with surfaces that are positively charged [99]. However, if the environment (where the bacteria and surface are located) is in high ionic condition, then this electrostatic interaction will be reduced due to the charge screening effect (neutralizing) caused by oppositely charged ions in the environment [93,99]. In the second stage, adhesion becomes irreversible (permanent) without any need for physical or chemical intervention, firmly anchoring the organism to the surface within several hours [95,99]. Similar to the primary stage, van der Waal interactions are involved, which are between the outer cell wall and the surface. Moreover, polysaccharides and proteins play a crucial role in the transition from reversible to irreversible cell attachment, whereas the irreversible attachment is mostly dominated by the production of extracellular polymeric substances (EPSs) [92,99,100]. EPSs consist of polysaccharides, proteins, extracellular DNA, and lipids. EPSs, which are released by cells in biofilms adhered to surfaces, protect against mechanical damage and shear generated by flow [99]. Notably, within biofilms, EPSs exhibit a non-homogeneous distribution pattern among cells [101]. In the biofilm, an EPS offers various advantages to the cells. These benefits include adhesion, protection, and structural support. Specifically, the aggregative polysaccharides function as molecular glue, facilitating the adhesion of bacterial cells both to each other and to surfaces [102]. In addition to these factors, the surface characteristics of the substrate, including factors like surface charge density, wettability, roughness, stiffness, and surface architecture, are also regarded as significant factors that influence the initial adhesion of bacteria to surfaces [94]. Once these bacterial cells colonise a surface, they can create numerous problems such as infections. Many different methods are used to mitigate bacterial colonisation.

4. Antimicrobial Strategies

4.1. Chemical Bactericidal Strategies

Bactericidal chemicals play a major role in mitigating bacterial colonisation. Upon attachment, the planktonic bacterial cells begin clustering and start to form biofilms. Biofilms are highly resistant to antiseptics, antibiotics, and immune killing [103,104,105,106]. These biofilms create adverse effects on medical implants such as infections and continual inflammatory reactions, that can lead to implant failure [103,107,108,109]. Many researchers investigated chemical antibacterial strategies to avoid this issue. A common strategy is the widespread use of bactericidal and bacteria-repelling chemical agents, such as antibiotics and antiseptics [110]. Under the same category, antimicrobial surface coatings are more popular in many antibacterial applications [111,112,113,114,115,116,117,118]. Metal (Ag, Au, Cu, Zn) nanoparticles [113,119,120], metal oxide (ZnO, MgO, CuO, TiO2) nanoparticles [120,121], graphene family materials [122,123,124,125], fullerene [126,127], and carbon nanotubes [125] are frequently used as coatings [128] and additives for composites [129] in implants. For instance, Rezic et al. developed an antibacterial coating on PLA polymer using different types of commercially available nanoparticles such as Ag, Al2O3, Au, Pt, TiO2, Y2O3, ZnO, and ZrO2 [56]. This coating was bactericidal against Gram-positive Staphylococcus aureus (S. aureus) strains. Moreover, Patil and colleagues developed a silver-nanoparticle-embedded polystyrene polymer composite via thermal annealing and soft moulding process [130]. This composite showed an excellent bactericidal effect on both Gram-negative Escherichia coli (E. coli) and Gram-positive S. aureus bacteria.
The interaction of these nanoparticles with bacterial cells leads to several effects, including the inhibition of enzymes, deactivation of proteins, induction of oxidative stress, disruption of electrolyte balance, and alterations in gene expression levels resulting in lysing the bacteria [131]. These materials demonstrate excellent bactericidal efficacy. But the underlying antimicrobial mechanisms might induce a cytotoxicity effect for human cells [132,133,134,135]. For instance, Berardis et al. and Guan et al. studied toxicity assessments of ZnO nanoparticles on human colon carcinoma cells and human hepatocyte cell lines via toxicity assay methods. They found that these nanoparticles cause a reduction in human cell viability, presence of inflammatory biomarkers, DNA, and mitochondrial damage [136,137]. Moreover, Haase et al. studied the toxicity of silver nanoparticles using human leukemia cells and concluded that the silver nanoparticles cause elevated lactate dehydrogenase and a reduction in cell viability [138]. Moreover, the biggest threat to chemical-based antibacterial agents is the ability of bacteria to evolve into antimicrobial-resistant strains [139,140]. Given the widespread usage of antibacterial chemicals, they have the potential to be redundant due to antimicrobial resistance, less susceptibility of bacteria in biofilms, and potential cytotoxicity. These challenges necessitate alternative methods that are not prone to those threats.

4.2. Physical/Mechanical Bactericidal Strategies

4.2.1. Natural Bactericidal Nanostructured Surfaces

During the last decade, patterns with impressive bactericidal properties have been found on shark skin, gecko skin, lotus leaves, and the wings of cicadas, damselflies, and dragonflies [23,141,142,143]. These wings contain nanostructured surfaces that can prevent the growth of biofilms and kill a variety of bacteria by generating high bactericidal levels to different bacteria strains, as depicted in Table 2 [141,144]. For example, Ivanova et al. showed that cicada wing surfaces can eliminate Gram-negative Pseudomonas aeruginosa (P. aeruginosa) bacteria within 3 min of contact [145]. Interestingly, highly patterned or regularly arrayed nanoscale pillars on cicada wings are effective against Gram-negative bacteria, while the random nanofeatures or sigmoidal nanoarchitecture of dragonfly wings can kill both Gram-negative and Gram-positive cells [141,146,147,148,149]. Tripathy et al. reviewed that cicada wings are proficient at effectively lysing Gram-negative bacteria but not Gram-positive bacteria. The reason is due to the Gram-positive bacteria’s thick peptidoglycan cell wall, which is approximately 4 to 5 times thicker than that of Gram-negative bacteria [24]. The various nanopatterns comprise either patterned or random arrays of surface features (Figure 2b,h), and are easily distinguished by the height, tip diameter, base width, and spacing of their individual surface features [57,150], as shown in Table 2 and Figure 2.
The cicada wing species demonstrate ordered nanoarchitecture while dragonfly, damselfly, and gecko skin present random or hierarchical nanoarchitecture (Table 2 and Figure 3). All these surfaces are composed of nanopillars while gecko skin and lotus leaves are composed of spinules and nanotubes, respectively. The spinules are wire-like features with high aspect ratio having a height around 3000 nm [110]. There is a significant difference between the height of nanofeatures in gecko skin and other natural bactericidal nano-topographies (Figure 4). Moreover, cicada wing nano-topographies are bactericidal to Gram-negative bacteria species such as P. aeruginosa and P. fluorescens, while dragonfly wings, gecko skin, and damselfly wings are bactericidal to both Gram-negative (P. aeruginosa, E. coli, and P. gingivalis) and Gram-positive (S. aureus) bacteria. This bacterial cell rupture is caused by the prolonged suspension of the bacterial cell membrane on the nanostructured surface, which causes the membrane to stretch beyond its elastic limit [24,150,151,152,153]. It is also noticeable that the tip diameter is low when it comes to dragonfly wing, gecko skin, and damselfly wing architecture compared with cicada wings. Hence, it results in creating high stress on the bacterial cells to lyse the bacteria. As per the literature results included in Table 2, it is shown that there is an effect on bactericidal activity from the nanostructure pattern type for the different bacteria species. The absence of threats of antimicrobial resistance development and antibacterial tolerance due to biofilm formation makes bactericidal nano-topographies more appealing for both medical and industrial applications.
Figure 3. Micrographs of natural bactericidal nanostructured surface topographies. (a) Helium ion micrograph of a periodical cicada wing (Magicicada septendecim) showing hemispherical features. (b) Helium ion micrograph of an annual cicada wing (Tibicen tibicen) showing spherically capped conical features. (c) Helium ion micrographs of a common sanddragon dragonfly wing (Progomphus obscurus) showing spherically capped cylindrical features with a high aspect ratio. Adapted with permission from ref. [154]. (d) Scanning electron microscopy (SEM) micrograph (tilted 45°) showing the self-organised free-standing and clustered nanopillar arrays on a black damselfly wing (C. haemorrhoidalis). Adapted with permission from ref. [155]. (e) SEM micrograph of natural gecko skin (Strophurus williamsi). Adapted from ref. [110]. (f) SEM micrograph of the surface of a natural lotus leaf (scale: 10 μm). Adapted with permission from ref. [156]. Copyright© 2015, American Chemical Society. Dragonfly (c) and damselfly (d) wings comprise similar non-uniform (random) nanoarchitectures with different heights and orientations. However, the nanofeatures in gecko skin (e) comprise a high aspect ratio and high spacing nanofeatures compared to all the other natural nanostructures.
Figure 3. Micrographs of natural bactericidal nanostructured surface topographies. (a) Helium ion micrograph of a periodical cicada wing (Magicicada septendecim) showing hemispherical features. (b) Helium ion micrograph of an annual cicada wing (Tibicen tibicen) showing spherically capped conical features. (c) Helium ion micrographs of a common sanddragon dragonfly wing (Progomphus obscurus) showing spherically capped cylindrical features with a high aspect ratio. Adapted with permission from ref. [154]. (d) Scanning electron microscopy (SEM) micrograph (tilted 45°) showing the self-organised free-standing and clustered nanopillar arrays on a black damselfly wing (C. haemorrhoidalis). Adapted with permission from ref. [155]. (e) SEM micrograph of natural gecko skin (Strophurus williamsi). Adapted from ref. [110]. (f) SEM micrograph of the surface of a natural lotus leaf (scale: 10 μm). Adapted with permission from ref. [156]. Copyright© 2015, American Chemical Society. Dragonfly (c) and damselfly (d) wings comprise similar non-uniform (random) nanoarchitectures with different heights and orientations. However, the nanofeatures in gecko skin (e) comprise a high aspect ratio and high spacing nanofeatures compared to all the other natural nanostructures.
Nanomaterials 13 02799 g003
Table 2. Natural antibacterial nanostructured surfaces.
Table 2. Natural antibacterial nanostructured surfaces.
SchemeNanofeature TypeNanofeature
Arrangement
Dimensions of the NanofeatureWettabilityAntimicrobial
Activity
Ref.
Tip Diameter (nm)Base Diameter (nm)Height (nm)Spacing (nm)
Periodical cicada wing
(Magicicada ssp.)
Hemispherical shape nanofeaturesOrdered16716783.5252~80°Lethal to S. cerevisiae fungus[154]
Dog day annual cicada
(Tibicen ssp.)
Nanoneedles (spherically capped, conical, nanoscale pillars)Random/hierarchical57104183175132°Bactericidal to P. aeruginosa (Gram-negative)[154]
Cicada wing
(Psaltoda claripennis)
NanopillarsOrdered60100200170~159°Bactericidal to P. aeruginosa (Gram-negative)[145]
Cicada wing
(Megapomponia intermedia)
Tubular nanofeatures (nanopillars)Ordered156~312241165~136°Bactericidal to P. fluorescens (Gram-negative)[101]
Cicada wing
(Cryptotympana Aguila)
Tubular nanofeatures (nanopillars)Ordered159~318182187~113°Bactericidal to P. fluorescens (Gram-negative)[101]
Cicada wing
(Ayuthia spectabile)
Tubular nanofeatures (nanopillars)Ordered207~414182251~96°Bactericidal to P. fluorescens (Gram-negative)[101]
Sanddragon dragonfly wing (Progomphus obscurus) NanograssRandom/hierarchical53-241123119°Bactericidal to P. aeruginosa (Gram-negative), S. aureus and B. subtilis (Gram-positive)[154]
Dragonfly wing
(D. bipunctata)
NanograssRandom/hierarchical50–70-240-153°Bactericidal to P. aeruginosa (Gram-negative), S. aureus and B. subtilis (Gram-positive)[146]
Dragonfly wing
(Orthetrum villosovittatum)
NanopillarRandom/hierarchical37 and 57-189 and 311--Bactericidal to E. coli (Gram-negative)[148]
Gecko skin
(Strophurus williamsii)
SpinulesRandom/hierarchical50~4003000500>136°-[110]
Gecko skin
(Lucasium steindachneri)
SpinulesRandom/hierarchical--2000–4000~500150°Bactericidal to Streptococcus mutans (Gram-positive) and Porphyromonas gingivalis (Gram-negative)[102,131]
Lotus leaves (Nelumbo nucifera)NanotubulesRandom/hierarchical-100–200300–1100~150–250154°Bactericidal to E. coli (Gram-negative)[157]
Damselfly wing
(Calopteryx haemorrhoidalis)
NanopillarsRandom/hierarchical~48-~433116157°Bactericidal to P. aeruginosa (Gram-negative) and S. aureus (Gram-positive)[149]

4.2.2. Artificial Nanostructure Fabrication Methods in Polymers

To generate different morphologies of polymeric nanostructured surfaces, many studies were conducted using different chemical and mechanical methods. Nanoimprinting lithography (NIL) [57,132], electron beam lithography (EBL) [141,158], reactive ion etching (RIE) [159,160], colloidal lithography [161,162], laser-based lithography techniques, and anodic aluminium oxide template (AAOT) [141,163,164] are polymer nanostructured surface fabrication methods which are used in the literature (Table 3). The NIL, EBL, and AAOT processes demonstrate high nanofeature parameter controllability. However, limitations like high fabrication cost and multiple process steps make them non-facile and inefficient [141,158,163,165].

Reactive Ion Etching

Reactive ion etching (RIE) is perhaps the most frequently used method to develop non-polymer nanostructured surfaces, and produces high-aspect-ratio nanostructures by mimicking cicada and dragonfly nanostructures [159,166]. It is based on high energetic ion which is bombarded into the substrate coated with a photoresist pattern mask to remove material from the surface which is unexposed to photoresist, resulted in nanostructured surfaces [167,168]. Ivanova et al., Hasan et al., Roy et al., and Ganjian et al. have used the RIE process to fabricate high-aspect-ratio antibacterial nanostructured surfaces on non-polymer materials such as black silicon and black titanium [146,169,170]. However, limited literature related to polymer surfaces was found. Kayes et al. developed a nanostructured surface on PP polymer via maskless RIE (without using a photoresist pattern mask) method under different plasma gas conditions, and achieved a bactericidal activity on E.coli bacteria [171]. Moreover, Patil et al. was successful in obtaining more than 85% of bactericidal efficacy against P. aeruginosa bacteria on commercially pure nitrile (CPN) and nitrile gloves without using a mask in RIE [160]. Moreover, the etching time, type of gas, gas pressure, plasma gas flowrate, and plasma power are crucial parameters in this process to optimise the nanostructures to obtain the required bactericidal efficacy. Another study involving maskless RIE on the PET polymer fabricated bactericidal nanostructures against P. aeruginosa. Nevertheless, the same etched surface did not demonstrate bactericidal efficacy against S. aureus bacteria [172]. Interestingly, the generated surfaces are random arrays of nanofeatures as depicted in Figure 5d,e. This process is comparatively a facile method due to the absence of the photoresist mask. Yet, there is no evidence in the literature that maskless RIE is used to fabricate nanostructures on polymer surfaces which are effective in lysing both Gram-negative and Gram-positive strains.

Colloidal Lithography

Colloidal lithography is a similar process to RIE which uses nano- or microspheres instead of a photoresist mask. Distinct nanostructured surfaces are obtained by controlling the micro/nanosphere size, plasma gas pressure and flow rate, and etching time [161,162,173]. Hazel et al. fabricated bactericidal PET nanocones via a combination of colloidal lithography and RIE [161]. To fabricate the nanocones, initially the polystyrene (PS) microspheres were spin-coated as a 2D hexagonally close-packed array, as depicted in Figure 6 (step 1). Nanostructure size is dependent on the size of the microsphere, and 200 nm and 500 nm sized microspheres were used to produce uniform nanocones in this work. Oxygen plasma is bombarded on the PS coated substrate with relevant power and flow rate to etch the PS and substrate. Further etching will reduce the size of the PS microsphere and remove the substrate material while taking the PS microsphere as a mask (Figure 6 (step 2)). For each of the PS microsphere sizes, different etching times were performed to obtain distinct nanofeatures on the surface with different height, tip diameter, and spacings, as depicted in Figure 6b–i. The researchers were able to produce nanostructures with a 20 nm tip width and 400 nm height which were bactericidal to Gram-negative E. coli and Klebsiella pneumoniae (K. pneumoniae).
Mo et al. fabricated different types of nanofeatures (nanopillars, micropillars, microcones) on PEEK polymer using PS microspheres as the mask [162]. The fabricated nanofeatures were lethal to E. coli bacteria. A similar process has been used as the one depicted in Figure 6a. However, the researchers used two gases such as Ar and O2. Using Ar in colloidal lithography can produce nanotube-like features via anisotropic etching (etching in vertical direction) without decreasing the size of the microspheres. On the other hand, using O2 as the plasma gas can produce nanocone features while etching material isotropically (etching in both vertical and horizontal directions) by decreasing the size of the microspheres. Hence, this process can be effectively used to fabricate nanostructured surfaces on polymers via controlling the parameters like microsphere size, plasma power, gas flowrate, and etching time. Colloidal lithography can produce a uniform array of nanostructures with varying nanofeature sizes more effectively than the RIE procedure.

Nanoimprinting Lithography

NIL is a high-throughput fabrication approach for producing nanostructures. It entails transferring a pattern from a mould to a deposited polymer. A liquid polymer layer, known as a “resist”, is placed on the substrate surface and mechanically pressed with a silicon stamp, creating a mould pattern imprint on the polymer substrate. The mould is then removed to reveal the polymer impression, and a RIE procedure is used to remove the polymer resist (Figure 7), resulting in the final nanostructure [174,175]. The process can be divided into two types: thermal nanoimprinting lithography (TNIL) and ultraviolet nanoimprinting lithography (UVNIL) [175]. In the TNIL process, the pattern is made using both pressure and increased temperature (Figure 7a), which is often higher than the polymer’s glass transition temperature [176]. However, the mould removal process cause more damage to the nanostructures. The UVNIL process reduces the structural deformation during the demoulding process in the TNIL method. UVNIL uses UV radiation to crosslink the polymeric nanostructures (Figure 7b) [177,178]. However, some researchers have used a combination of both processes to get the maximum advantage of producing highly bactericidal nanostructured surfaces [179,180].
Oopath et al. created surfaces that resemble rose petals and have antibacterial properties using a combination of TNIL, UVNIL, and the hydrothermal technique [179]. Initially, they used NIL to imitate the microstructures found on rose petals on the surface of a PVDF-HFP film, as shown in Figure 7b. The hydrothermal approach was then used to grow ZnO nanostructures on the PVDF-HFP film, resembling the structure of rose petals. As shown in Table 3, the experimental findings showed that PVDF-HFP films mimicking rose petals had about 100% bactericidal efficiency against both strains of bacteria.
Dickson et al. also used NIL to construct nanopillars inspired by cicada wing structures on a PMMA substrate (Figure 7d–f). The PMMA film was applied onto the glass substrate using a spin coating method, followed by an annealing procedure on a hot plate at 100 °C prior to the imprinting stage. Different samples were acquired via distinct approaches to obtain various sizes of nanopillars (Table 3). PDMS nano hole moulds, a nickel stamp, and natural cicada wings were used in the NIL process to imprint the pattern on PMMA surfaces [57].
Shung et al. developed a nano-in-microstructured hierarchical superhydrophobic surface using a combination of UVNIL and RIE [181]. A PDMS mould with nanostructures was fabricated using a silicon mould. Then, the UV-curable polymer resist was imprinted using the prepared PDMS mould. Later, this nanostructured polymer layer was transferred into a thermal shrinkage film with an annealing process and RIE to create the final nanostructure (Figure 7c). The created nanostructured polymer film was superhydrophobic with a water contact angle of 150°. The researchers harnessed these surfaces to create superhydrophobic surfaces, primarily for self-cleaning purposes. While the bactericidal properties of these surfaces have not been experimentally evaluated, the nanofeature characteristics seem promising for potential mechano-bactericidal effects, with respect to the mechano-bactericidal nanofeature parameter ranges in Figure 4.
Even though NIL demonstrates high controllability in nanostructure dimensions, the multiple steps in the process lead to a high cost.

Laser-Based Lithography Techniques

In laser-based lithography techniques, a laser beam or set of beams is used to create nanopatterns on different material substrates via a photosensitive material (typically a photoresist). Laser lithography, laser interference lithography, and femtosecond laser lithography are key methods used in the literature for fabricating nanostructured surfaces [176,182,183,184]. The laser lithography process uses a laser beam to directly produce a pattern on the photosensitive material. Resists used in this process can have either negative or positive tones, and the region on the resist that is exposed can be crosslinked or chain-scissioned [185]. The patterned photoresist material acts as a mask on the target substrate, and plasma etching usually follows in order to etch the unexposed area of photoresist to obtain the final nanostructure on the substrate. Kim et al. produced a nanostructured bactericidal and superhydrophobic surface on PMMA polymer using the KrF laser lithography process with dry etching, thermal oxidation, high-density plasma chemical vapour deposition, and TNIL, as depicted in Figure 8a–i [176]. BARC and LX-429 photoresists were coated on a Si wafer, and the initial nano pattern was produced using KrF laser beam. Then, a dry etching process was used with a combination of Cl2 and HBr gases to etch the Si wafer by taking the produced photoresist nanopattern as a mask. Subsequently, a nanocone structure was obtained, as depicted in Figure 8e. Later, the high-density plasma chemical vapour deposition process used to develop nanopillars on Si, and the TNIL process replicated the pattern on PMMA with dimensions of 250 nm (base diameter), 490 nm (height), and 300 nm (spacing) (Figure 8i). The fabricated surface exhibited bactericidal properties against P. aeruginosa and E. coli bacteria. In the early stages of attachment, which occurred during the first 1–4 h of bacterial incubation, most adhered E. coli cells were lysed, primarily attributable to the mechano-bactericidal effect, when compared to a flat substrate. Conversely, there was no substantial bactericidal impact on P. aeruginosa during this initial attachment phase. However, over a period of 2–7 days, a significant portion of both bacteria lysed due to the inhibition of adhesion and bacterial growth because of the surface’s superhydrophobic nature.
Laser interference lithography (LIL) creates detailed patterns on a substrate using the interference pattern of coherent laser beams. Complex interference patterns are formed by adjusting the angles and wavelengths of the laser beams, resulting in high-resolution patterns on the substrate [184,186,187,188]. During the process of fabricating polymer nanostructured surfaces, the initial photoresist pattern can be made using the LIL method with high-resolution patterns due to the interference of multiple beams rather than using the conventional laser lithography processes [189].
Quilis et al. fabricated arrays of thermoresponsive poly(N-isopropylacrylamide) (pNIPAAm)-based hydrogel nanostructures with a gold nanoparticle array for applications in highly sensitive chemical and biological sensing [190]. The four-beam UV laser interference lithography (UVLIL) method was used to pattern the photoresist on the pNIPAAm hydrogel, and the hydrogel was etched using the dry etching process to obtain the final nanostructure, as depicted in Figure 8j. This resulted in nanostructures with 132 nm (diameter), 50 nm (height), and 463 nm (spacing) dimensions (Figure 8k). Even though the spacing is too high compared to diameter and height, these methods can be used in the development of mechano-bactericidal nanostructures on polymers. Interestingly, Valle et al. fabricated microstructures on polystyrene polymer surfaces using the LIL process [191]. These microstructures took various forms, including line and pillar-like patterns as well as lamella shapes. The researchers examined bacterial attachment, using S. aureus, to assess the antibacterial properties in both static and fluid flow conditions. Surprisingly, the line and pillar-like microstructures seemed to promote bacterial adhesion, while the lamella-shaped patterns reduced bacterial adhesion under both static and flow conditions. Furthermore, the researchers primarily demonstrated an anti-adhesion effect as a key contributor to the antibacterial activity of these microstructures. Additionally, they proposed that these microstructures provided a mechano-bactericidal effect inspired from the microfeatures found on shark skin.
To conclude, the laser-based lithography technique is an extended version of the NIL process. Given the high precision of the nanofeatures, the various steps employed in the process make it a non-facile method in commercial implant fabrication. However, LIL can also be used as a facile method without using NIL and other supportive processes like etching as per the study of Valle et al. Moreover, to make such surfaces mechano-bactericidal, nanofeature parameters should be carefully controlled as per Figure 4.
Figure 8. (ai) Laser-based lithography process assisted with TNIL. Adapted from ref. [176]. (j) Schematic view of the crosslinking of pNIPAAm-based hydrogel polymer via arrays of interference laser beams. (k) Change of nanofeature parameters with irradiation dose. (ln) AFM micrographs of pNIPAAm-based hydrogel polymer nanofeature parameters under different irradiation doses. Images (jn) adapted with permission from ref. [190].
Figure 8. (ai) Laser-based lithography process assisted with TNIL. Adapted from ref. [176]. (j) Schematic view of the crosslinking of pNIPAAm-based hydrogel polymer via arrays of interference laser beams. (k) Change of nanofeature parameters with irradiation dose. (ln) AFM micrographs of pNIPAAm-based hydrogel polymer nanofeature parameters under different irradiation doses. Images (jn) adapted with permission from ref. [190].
Nanomaterials 13 02799 g008

Anodic Aluminium Oxide Template

To create well-ordered nanofibers, AAOT is used as a template to recreate fine nanopores employing a molten polymer. Nanostructure fabrication via AAOT consists of several steps, as depicted in Figure 9d. First anodization, acid etching, and second anodization are the key steps of producing the template. The AAOT process is used to fabricate the template which is required to use in the NIL process to impart the nanostructure to a polymer material [192]. During the initial anodization process, electoral density distribution is formed over the thin aluminium sheet. Then, the acid etching is used to remove the initial aluminium oxide layer to form an organized barrier layer to produce nano pores in an ordered manner. Finally, a secondary anodizing process follows to allow the developed pores to penetrate into the original aluminium sheet. Temperature, voltage, and the pH level of the anodising electrolyte play a crucial role of determining the pore depth, spacing, and diameter [193,194].
Cui et al. modified the polycarbonate (PC) substrate surface using AAOT-assisted wet etching and hot embossing [195]. This is similar to the procedure depicted in Figure 9d. The ability to change the geometric parameters (height, diameter, and spacing of nanopillars) of the nanostructure more efficiently (Figure 9a–c) than other manufacturing methods, such as NIL, RIE, and colloidal lithography, is an essential component of this nanostructure fabrication process. As mentioned in Table 3, the researchers fabricated bactericidal nanostructures with a bactericidal efficacy of more than 95% against E. coli bacteria. Moreover, the interaction between bactericidal efficacy and geometric parameters were also studied in this work, and the researchers concluded that high efficacy is obtained at 200 nm of nanopillar height with smaller cap diameters. Also, as per the study, there was an optimum value (170 nm) for the spacing which influence to have more bactericidal efficacy.
Zhang and colleagues devised a hybrid nano-topography via a mechano-bactericidal mechanism using an AAOT assisted technique [196]. They employed a layer-by-layer assembly method to create this structure, involving the sequential application of tannic acid (TA) and iron ions (Fe3+) through a dip-coating process. Polymeric nanopillars were crafted by forming a polymeric resin from diglycidyl ether of bisphenol A resin (DGEBA), polyether amine resin (D230), and ethyl acetate using the AAOT process. The coating of TA and Fe3+ on these nanopillars exhibited a significant photothermal effect resulting in enhanced antibacterial properties in addition to its mechano-bactericidal effect. These hybrid nanopillars demonstrated exceptional efficacy in lysing P. aeruginosa and S. aureus bacteria, with a bactericidal rate exceeding 99%. It is worth noting that the photothermal coating can have inherent cytotoxicity, but the developed surface exhibited excellent compatibility with mammalian cells (more than 88% cell viability) in in vitro studies.
Similar to NIL and laser-based lithography methods, AAOT involves a series of process steps and is primarily suitable for creating nanostructures on flat surfaces due to the difficulties involved in mould removal of NIL on curved surfaces. As AAOT utilizes the NIL process, any drawbacks associated with NIL also apply to this method.

Electron Beam Lithography (EBL)

Lithography is a nanofabrication method which uses light or focused particle beams to remove material from a substrate and obtain nanostructures [197]. EBL and focused ion beam (FIB) are the two main categories of the particle beam lithography (PBL) process [198]. Particles can be identified as electrons and ions (Ga+, He+, Ne+). In EBL, high-energy electrons are employed, whereas in focused ion beam (FIB) milling, high-energy ions are utilized. EBL is a maskless lithography process which uses an electron gun to fabricate nanoscale patterns on a substrate [197]. The electron beam produced by the gun either images (in SEM) a surface or fabricates a previously deposited resist on a substrate. The resist that is exposed can be crosslinked or chain-scissioned [185]. Then, exposed and nonexposed resist regions are selectively removed via a solvent. While the covered regions are protected, the exposed areas can be further processed for etching or thin-film deposition [199]. PMMA polymer is a common positive e-beam resist whose solubility changes when exposed to an e-beam [185,200]. Moreover, polyacrylic acid (PAA), polyethylene glycol (PEG), and hydrogen silsesquioxane (HSQ) are the mostly used resist materials in the EBL process [201].
Kallas et al. produced nanopillars on polycarbonate polymers using a combination of the UV-NIL, dry-etching, and EBL process [201]. The developed nanostructures were bactericidal against E. coli bacteria. The EBL process is used to make the pattern for the UVNIL process. The pattern was made from hydrogen silsesquioxane (HSQ) resist. Compared to photolithography processes, EBL provides minimum nanofeature size as fine as ~2 nm while photolithography fabricates around ~50 nm [180]. However, the direct writing of patterns by scanning electron beam is a slow process with low throughput. As a result, EBL is utilised for photomask preparation for photolithography or the direct writing of low-volume and small-area patterns.

Hydrothermal Synthesis

The hydrothermal synthesis process develops a high-temperature and high-pressure environment in a closed Teflon-lined stainless steel reactor [202]. During this process, the sample (or implant) is submerged completely in alkaline solution to perform a reaction to achieve nanostructures. Because of its reliability, environmental friendliness, simplicity, low cost in comparison to other procedures, and flexibility for material morphology control, hydrothermal synthesis is the most versatile nanostructured surface fabrication method for metals, ceramics, and non-flat complex geometry surfaces [159,166,203]. As per the literature relevant to hydrothermal synthesis on titanium surfaces, it can resemble the nanofeatures of cicada wings or dragonfly wings with high bactericidal efficacy [159,166,204]. Hence, theoretically this method can be used to fabricate mechano-bactericidal nanostructures on 3D implants. For instance, Jaggessar and colleagues fabricated bactericidal nanostructures on 3D titanium surfaces via the hydrothermal synthesis process, and were successful in obtaining 91% more efficiency in anti-adhesion compared to 2D surfaces against S. aureus bacteria, showing the same or higher bactericidal efficacy than 2D surfaces [205].
On the other hand, due to the low melting point and low glass transition temperature of polymers, it is challenging to use the hydrothermal synthesis process to develop nanostructures on polymer surfaces. Hence, using the hydrothermal synthesis process for polymers has been overlooked in the bactericidal nanostructured surface field. However, Yoo et al. created ZnO nanowires on flexible plastics such as PET, PC, and polyimide (PI) using a low-temperature (approx. 90 °C) hydrothermal synthesis method, as depicted in Figure 10a [206]. After coating the flexible plastic substrate with Ag ink solution, the Ag layer was transformed into a nanoporous layer at 130 °C (Figure 10b,c). The hydrothermal reaction was processed inside the Z n 2 + ion solution of this Ag-coated plastic substrate. Finally, as shown in Figure 10d–f, ZnO crystallised on the Ag-seeded substrate at 90 °C after 9 h. This nano-topography is similar to the mechano-bactericidal nano-topography developed by Bhadra et al. [159] and Zhao et al. [203]. Even though there are a few studies that used the hydrothermal synthesis process to develop nanostructures via metal oxides on polymer substrates, no study has yet been reported in the literature for using hydrothermal synthesis to fabricate micro/nanostructured surfaces on pure polymers.
Table 3. Summary of polymer nanostructured surface fabrication methods. Table abbreviations: Poly(ethylene glycol) dimethacrylate—PEGDMA; Poly(vinylid ne fluoride-co-hexafluoropopylene)—PVDF-HFP; Height—H; Tip diameter—Dt; Base diameter—Db; Diameter—D; Spacing—S; Width—W; Pillar density—PD; Pitch—P; Aspect ratio—AR; Roughness—R.
Table 3. Summary of polymer nanostructured surface fabrication methods. Table abbreviations: Poly(ethylene glycol) dimethacrylate—PEGDMA; Poly(vinylid ne fluoride-co-hexafluoropopylene)—PVDF-HFP; Height—H; Tip diameter—Dt; Base diameter—Db; Diameter—D; Spacing—S; Width—W; Pillar density—PD; Pitch—P; Aspect ratio—AR; Roughness—R.
MethodSubstrate
Material
Supportive ProcessesFeature TypeNanofeature ParametersBactericidal Effect (Max. Efficiency %, Incubation Time in Hours)Ref.
NILPMMASilicon mould preparationNanopillarsH210 nm, 300 nmLethal to E. coli bacteria (~50%, 24 h)[57]
Dt70 nm, 190 nm, 215 nm
Db100 nm, 130 nm, 380 nm
S170 nm, 320 nm, 595 nm
OrmoStamp (glass)AAOT, deep RIENanopillarsH200, 300, 400 nmLethal to S. aureus bacteria (~100%, 0.5 h)[207]
D70, 80 nm
PD40 µm2, 70 µm2
PEGDMARIENanoneedleH300 nmLethal to E. coli bacteria (N/A)[208]
Dt50 nm
Db200 nm
P500 nm
PVDF-HFPHydrothermal methodArray of micropapillae with nanofold structuresD (papillae)20 µmLethal to Gram-positive S. agalactiae: Micropapillae (~97%, 12 h), Nanopillars (100%, 12 h); Gram-negative E. coli, Micropapillae (~100%, 12 h), Nanopillars (100%, 12 h)[179]
D (nanofold)300–400 nm
D300 nm
P600 nm
PCRIE, EBLNanopillarH25 nmLethal to E. coli (N/A)[201]
Db40 nm
S100, 200, 500 nm
Laser LithographyPMMAThermal oxidation, RIE,
chemical vapour deposition, TNIL
NanopillarsH450 nmLethal to P. aeruginosa (~70%, 8 h) and E. coli (~70%, 8 h)[176]
Db250 nm
AR3
LILPSNoneBoth pillar-like and lamella-shaped patternsH2 µmLethal and anti-adhesion under flow conditions to S. aureus[191]
W3 µm
S3 µm
Colloidal lithographyPETRIENanoconesH530–350 nmLethal to E. coli (~30%, 1 h) and K. pneumoniae (~30%, 1 h)[161]
Dt20–300 nm
Db55–380 nm
AR1–7
PEEKRIENanopillars, nanocones, micropillars, micro-conesH350 nmLethal to E. coli: Nanopillars (>95%, 12 h), Micropillars (~90%, 12 h), Micro-cones (>75%, 12 h)[162]
P200, 250 nm
D100, 50 nm
AAOTPolymer made from DGEBA and E230Wet etching, dip coatingNanopillarsH500–600 nm Lethal to P. aeruginosa (~99%, 3 h) and S. aureus (~99%, 3 h)[196]
Dt100 nm
Db350 nm
PCHot embossing and wet etchingNanopillarsH143–408 nmLethal to E. coli bacteria (>95%, 3 h)[195]
Dt26.7–33.4 nm
Db66.3–154.3 nm
S100–307.8 nm
RIEPPNoneNanofibrilsH0.5, 1 µmLethal to E. coli (~99.6%, 24 h)[171]
D30, 40 nm
CPNNoneNanopillarsH1–1.4 µmLethal to P. aeruginosa (~80%, 24 h)[160]
R0.92–1.42 µm
PETNoneNanopillarsH128.67 nmLethal to P. aeruginosa (~99%, 24 h)[172]
D21.7 nm
NanowiresH501.4 nm
D29.54 nm

5. Factors Affecting Bactericidal Activity of Mechano-Bactericidal Mechanism

The stress experienced by the bacterial cell membrane due to a nanopattern is determined by both the shape and arrangement of the nanofeatures and by how strongly the cell membrane is attached to the surface, with various factors mentioned in Section 3. By fine-tuning the various geometric aspects of the nanopillars, it is possible to increase the level of stretching that the membrane undergoes [24,150,153,209]. Seng et al. showed that S. aureus is the most common pathogenic bacterium for breast implant infections (49% of 47 cases), whereas P. aeruginosa was identified as the second most pathogenic bacterium in this study (19% of 47 cases) [210]. Hence, it is more important to consider both bacteria species when developing bactericidal nanostructured surfaces.
Cui et al. found that the bactericidal efficacy of nanostructures developed on PC surfaces increased with spacing and reached a maximum at 170 nm of nanopillar spacing, but then decreased. On the other hand, the bactericidal efficiency increased exponentially with the height of the nanopillar up to 300 nm, but then plateaued [195]. Different parameters of the nanostructured surface, including the shape of the feature [20,145,211], height [145,150,209], tip radius [145,209], spacing [145,209], aspect ratio [146,209], base width [209], and the arrangement (periodic or random), influence its bactericidal efficiency [207,212]. For example, Fisher et al. found that sharp diamond nanocones on a silicon wafer arranged in a nonuniform array (random) with a lower distribution density were much more bactericidal against P. aeruginosa than uniformly (periodic) arranged, high-density nanocones [213]. However, this cannot imply the effect of nanostructure density (or spacing) and nanoarchitecture type on bactericidal efficacy solely. Effects and relationships among the factors should be studied separately, one parameter at a time, to get a clear understanding. In addition to that, Velic et al. has summarised the effect of tip radius (or diameter) and spacing on the maximum von Mises stress σ v m a x , areal stress ( σ A ) , and contact pressure ( P m a x ) . These stresses were then correlated to the bactericidal efficacy from the available data in the literature (Figure 11) [209]. It was shown that the reduction in both tip radius and spacing increases bactericidal efficacy in most of the cases plotted in the graph (Figure 11). Hence, it is beneficial to use various nanostructure configurations with distinct parameters to achieve highly effective bactericidal nanostructured surfaces.
In a separate study, Velic et al. studied the effect of geometrical parameters of a nanostructured surface for lysing bacteria via a finite element analysis approach, and concluded that the reduction in nanopillar radius and spacing increased the maximal strains and the frequency of perturbation points accordingly [214]. The maximum strain is in the highest point of the pillar. This suggests that whenever there is stretching that induces pore formation and potential rupture of the bilayer, it will consistently begin within the inner leaflet at the highest point of the pillar.
Anti-bio adhesion is another mechanism which uses the wettability of the nature-inspired nanostructured surface to achieve the anti-adhesion of the bacteria [173,215]. This phenomenon is not a bactericidal activity, but it uses the non-adhesion mechanism resulting in a reduction in viability of the bacteria on the surface. There were many researchers who investigated the effect of surface wettability (hydrophobicity [154,215,216,217,218] or hydrophilicity [216,219,220]) on the bactericidal efficacy of nanostructured surfaces. Valiei et al. studied the mechano-bactericidal efficacy of etched silicon nanopillars against P. aeruginosa bacteria. Their results showed that the bactericidal efficacy decreases when the surfaces hydrophobicity increases. In other words, bactericidal activity was greatest on superhydrophilic surfaces and decreased as the surface became more hydrophobic. It significantly decreased when the contact angle of the substrate exceeded 90° [215]. These results suggest that super-hydrophilic nanopillared surfaces are more suitable for mechano-bactericidal activity. In contrast, superhydrophobic surfaces, although not bactericidal, may possess antibiofouling properties owing to their self-cleaning characteristics. Interestingly, according to Ivanova and colleagues, despite the cicada wing nanostructures demonstrating a superhydrophobic nature, the adhesion of bacteria was significant to kill the P. aeruginosa bacterium [145]. They reduced the hydrophobicity of the cicada surface by coating it with gold, which reduced the water contact angle from 158° to 105°. However, no significant difference in bactericidal action was observed, implying that the topography of the nanostructure is more important than the surface chemistry [145]. A large number of contributory factors makes it challenging to study their impact and the interactions between them. Therefore, more comprehensive studies are needed to explain the effects of those factors on the bactericidal effect of the surface.

6. Conclusions and Outlook

Antibacterial implants are highly desired to prevent failures due to bacterial infections. To achieve that, different approaches have been developed in the literature, briefly outlined in the Introduction. Among different pathways, “bactericidal nanostructured surfaces” have taken the spotlight in the field for years due to their excellent bactericidal efficacy on different pathogenic bacteria species, and possibility of implementation on different implant materials. Polymers are crucial materials for biomedical implants due to their inherent low density and ease of manufacturing, for example, additive manufacturing processes. TNIL, UVNIL, AAOT, EBL, and laser-based lithography processes have been efficiently used to achieve high-precision bactericidal nanofeatures on different polymeric surfaces. However, the multiple steps involved in each process make them non-feasible in the commercial scale of the relevant biomedical application. Moreover, almost all fabrication processes which were carried out in the literature were related to flat surfaces. However, most implants are not flat. Implants contain intricate shapes of internal and external features. Hence, there is no successful fabrication method to produce bactericidal nanostructures on the internal features of polymeric implants. Hydrothermal synthesis can be an effective solution to this requirement. However, the low melting point and glass transition temperature has made this process non-feasible. On the other hand, RIE and colloidal lithography are facile methods compared with lithography and templating. To optimise the topography of nanofeatures for the majority of the methods described in Table 3, assistance from the RIE process is required. However, several researchers have succeeded in fabricating nanostructures on various polymers using RIE, with or without a photoresist mask. Interestingly, these structures have shown great bactericidal efficacy against Gram-negative bacteria. However, there is no research that has shown bactericidal efficacy on Gram-positive bacteria using maskless RIE. This process has reduced most of the steps in other popular methods and could be useful in developing bactericidal nanostructures on different polymers. Due to the directional etching principle (nanofeatures will be created perpendicular to the surface) of RIE and colloidal lithography, again it is difficult to fabricate nanostructures on 3D surfaces. As a result, the density of the nanofeatures will not be homogenous due to the agglomeration of nanofeatures in low curvature areas and the surfaces which are parallel to plasma direction. In a broader context, the challenge of scaling up these processes to fabricate successful implants with strong bactericidal efficacy on a commercial scale remains unmet. Given their attributes of additive manufacturing and excellent biocompatibility, polymers are poised to play a pivotal role in addressing the aforementioned gaps in the biomedical field.

Author Contributions

Conceptualization, S.P.S.N.B.S.K., S.W.M.A.I.S., M.M. and P.K.D.V.Y.; writing—original draft preparation, S.P.S.N.B.S.K.; writing—review and editing, S.P.S.N.B.S.K., S.W.M.A.I.S., A.M., L.B., M.M. and P.K.D.V.Y.; visualization, S.P.S.N.B.S.K., S.W.M.A.I.S., A.M. and M.M.; supervision, S.W.M.A.I.S., A.M., L.B., M.M. and P.K.D.V.Y.; funding acquisition, P.K.D.V.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Australian Research Council (ARC) Industrial Transformation and Training Centre for Multiscale 3D Imaging (ARC ITTC M3D) Stryker Scholarship (Project IC 180100008) and Queensland University of Technology (QUT)”.

Data Availability Statement

No data were created in this review.

Acknowledgments

M.M. acknowledges the Australian Research Council (ARC) Early Career Award (DE210100975). A.M. acknowledges the support from VC postdoctoral fellowship for women in STEMM (UniSQ). The authors acknowledge continued support from the Queensland University of Technology (QUT).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Donlan, R.M. Biofilms: Microbial Life on Surfaces. Emerg. Infect. Dis. 2002, 8, 881–890. [Google Scholar] [CrossRef]
  2. Davies, D. Understanding Biofilm Resistance to Antibacterial Agents. Nat. Rev. Drug Discov. 2003, 2, 114–122. [Google Scholar] [CrossRef] [PubMed]
  3. Hernando-Amado, S.; Coque, T.M.; Baquero, F.; Martínez, J.L. Defining and Combating Antibiotic Resistance from One Health and Global Health Perspectives. Nat. Microbiol. 2019, 4, 1432–1442. [Google Scholar] [CrossRef] [PubMed]
  4. Li, B.; Webster, T.J. Bacteria Antibiotic Resistance: New Challenges and Opportunities for Implant-Associated Orthopedic Infections. J. Orthop. Res. 2017, 36, 22–32. [Google Scholar] [CrossRef] [PubMed]
  5. Varela, M.F.; Stephen, J.; Lekshmi, M.; Ojha, M.; Wenzel, N.; Sanford, L.M.; Hernandez, A.J.; Parvathi, A.; Kumar, S.H. Bacterial Resistance to Antimicrobial Agents. Antibiotics 2021, 10, 593. [Google Scholar] [CrossRef] [PubMed]
  6. Ashtiani, R.E.; Alam, M.; Tavakolizadeh, S.; Abbasi, K. The Role of Biomaterials and Biocompatible Materials in Implant-Supported Dental Prosthesis. Complement. Alternat. Med. 2021, 2021, 3349433. [Google Scholar]
  7. Santos, G.A. The Importance of Metallic Materials as Biomaterials. Adv. Tissue Eng. Regen. Med. Open Access 2017, 3, 200–302. [Google Scholar] [CrossRef]
  8. Klinke, T.; Daboul, A.; Maron, J.; Gredes, T.; Puls, R.; Jaghsi, A.; Biffar, R. Artifacts in Magnetic Resonance Imaging and Computed Tomography Caused by Dental Materials. PLoS ONE 2012, 7, e31766. [Google Scholar] [CrossRef]
  9. Prakasam, M.; Locs, J.; Salma-Ancane, K.; Loca, D.; Largeteau, A.; Berzina-Cimdina, L. Biodegradable Materials and Metallic Implants-A Review. J. Funct. Biomater. 2017, 8, 44. [Google Scholar] [CrossRef]
  10. Shekhawat, D.; Singh, A.; Bhardwaj, A.; Patnaik, A. A Short Review on Polymer, Metal and Ceramic Based Implant Materials. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1017, 012038. [Google Scholar] [CrossRef]
  11. Mahanty, A.; Shikha, D. Changes in the Morphology, Mechanical Strength and Biocompatibility of Polymer and Metal/Polymer Fabricated Hydroxyapatite for Orthopaedic Implants: A Review. J. Polym. Eng. 2022, 42, 298–322. [Google Scholar] [CrossRef]
  12. Peschke, E.; Ulloa, P.; Jansen, O.; Hoevener, J.-B. Metallic Implants in MRI—Hazards and Imaging Artifacts. Rofo 2021, 193, 1285–1293. [Google Scholar] [CrossRef] [PubMed]
  13. Hargreaves, B.A.; Worters, P.W.; Pauly, K.B.; Pauly, J.M.; Koch, K.M.; Gold, G.E. Metal-Induced Artifacts in MRI. AJR Am. J. Roentgenol. 2011, 197, 547–555. [Google Scholar] [CrossRef] [PubMed]
  14. Ramakrishna, S.; Mayer, J.; Wintermantel, E.; Leong, K.W. Biomedical Applications of Polymer-Composite Materials: A Review. Compos. Sci. Technol. 2001, 61, 1189–1224. [Google Scholar] [CrossRef]
  15. Jaganathan, S.K.; Supriyanto, E.; Murugesan, S.; Balaji, A.; Asokan, M.K. Biomaterials in Cardiovascular Research: Applications and Clinical Implications. Biomed Res. Int. 2014, 2014, 459465. [Google Scholar] [CrossRef] [PubMed]
  16. Teo, A.J.T.; Mishra, A.; Park, I.; Kim, Y.-J.; Park, W.-T.; Yoon, Y.-J. Polymeric Biomaterials for Medical Implants and Devices. ACS Biomater. Sci. Eng. 2016, 2, 454–472. [Google Scholar] [CrossRef] [PubMed]
  17. Bozkurt, Y.; Karayel, E. 3D Printing Technology; Methods, Biomedical Applications, Future Opportunities and Trends. J. Mater. Res. Technol. 2021, 14, 1430–1450. [Google Scholar] [CrossRef]
  18. Gonzalez, G.; Roppolo, I.; Pirri, C.F.; Chiappone, A. Current and Emerging Trends in Polymeric 3D Printed Microfluidic Devices. Addit. Manuf. 2022, 55, 102867. [Google Scholar] [CrossRef]
  19. Kalirajan, C.; Dukle, A.; Nathanael, A.J.; Oh, T.-H.; Manivasagam, G. A Critical Review on Polymeric Biomaterials for Biomedical Applications. Polymers 2021, 13, 3015. [Google Scholar] [CrossRef]
  20. Senevirathne, S.W.M.A.I.; Hasan, J.; Mathew, A.; Woodruff, M.; Yarlagadda, P.K.D.V. Bactericidal Efficiency of Micro- and Nanostructured Surfaces: A Critical Perspective. RSC Adv. 2021, 11, 1883–1900. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. Slepicka, P.; Kasalkova, N.S.; Siegel, J.; Kolska, Z.; Bacakova, L.; Svorcik, V. Nano-Structured and Functionalized Surfaces for Cytocompatibility Improvement and Bactericidal Action. Biotechnol. Adv. 2015, 33, 1120–1129. [Google Scholar] [CrossRef] [PubMed]
  23. Patil, D.; Overland, M.; Stoller, M.; Chatterjee, K. Bioinspired Nanostructured Bactericidal Surfaces. Curr. Opin. Chem. Eng. 2021, 34, 100741. [Google Scholar] [CrossRef]
  24. Tripathy, A.; Sen, P.; Su, B.; Briscoe, W.H. Natural and Bioinspired Nanostructured Bactericidal Surfaces. Adv. Colloid Interface Sci. 2017, 248, 85–104. [Google Scholar] [CrossRef] [PubMed]
  25. Al-Shalawi, F.D.; Mohamed Ariff, A.H.; Jung, D.-W.; Mohd Ariffin, M.K.A.; Seng Kim, C.L.; Brabazon, D.; Al-Osaimi, M.O. Biomaterials as Implants in the Orthopedic Field for Regenerative Medicine: Metal versus Synthetic Polymers. Polymers 2023, 15, 2601. [Google Scholar] [CrossRef] [PubMed]
  26. Asghari, F.; Samiei, M.; Adibkia, K.; Akbarzadeh, A.; Davaran, S. Biodegradable and Biocompatible Polymers for Tissue Engineering Application: A Review. Artif. Cells Nanomed. Biotechnol. 2017, 45, 185–192. [Google Scholar] [CrossRef] [PubMed]
  27. Mi, H.-Y.; Salick, M.R.; Jing, X.; Jacques, B.R.; Crone, W.C.; Peng, X.-F.; Turng, L.-S. Characterization of Thermoplastic Polyurethane/Polylactic Acid (TPU/PLA) Tissue Engineering Scaffolds Fabricated by Microcellular Injection Molding. Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 4767–4776. [Google Scholar] [CrossRef] [PubMed]
  28. Qian, J.; Xu, W.; Yong, X.; Jin, X.; Zhang, W. Fabrication and in Vitro Biocompatibility of Biomorphic PLGA/NHA Composite Scaffolds for Bone Tissue Engineering. Mater. Sci. Eng. C Mater. Biol. Appl. 2014, 36, 95–101. [Google Scholar] [CrossRef]
  29. Badaraev, A.D.; Sidelev, D.V.; Yurjev, Y.N.; Bukal, V.R.; Tverdokhlebov, S.I. Modes Development of PLGA Scaffolds Modification by Magnetron Co-Sputtering of Cu and Ti Targets. J. Phys. Conf. Ser. 2021, 1799, 012001. [Google Scholar] [CrossRef]
  30. Sun, F.; Sun, X.; Wang, H.; Li, C.; Zhao, Y.; Tian, J.; Lin, Y. Application of 3D-Printed, PLGA-Based Scaffolds in Bone Tissue Engineering. Int. J. Mol. Sci. 2022, 23, 5831. [Google Scholar] [CrossRef]
  31. Maheshwari, U.; Samuel, S.; Nagiah, V.K. Fabrication and Evaluation of (PVA/HAp/PCL) Bilayer Composites as Potential Scaffolds for Bone Tissue Regeneration Application. Ceram Int. 2014, 40, 8469–8477. [Google Scholar] [CrossRef]
  32. Hajiali, F.; Tajbakhsh, S.; Shojaei, A. Fabrication and Properties of Polycaprolactone Composites Containing Calcium Phosphate-Based Ceramics and Bioactive Glasses in Bone Tissue Engineering: A Review. Polym. Rev. 2018, 58, 164–207. [Google Scholar] [CrossRef]
  33. Greene, J.P. Microstructures of Polymers. In Automotive Plastics and Composites; Elsevier: Amsterdam, The Netherlands, 2021; pp. 27–37. [Google Scholar]
  34. Gaaz, T.S.; Sulong, A.B.; Akhtar, M.N.; Kadhum, A.A.H.; Mohamad, A.B.; Al-Amiery, A.A. Properties and Applications of Polyvinyl Alcohol, Halloysite Nanotubes and Their Nanocomposites. Molecules 2015, 20, 22833–22847. [Google Scholar] [CrossRef] [PubMed]
  35. Jiang, S.; Liu, S.; Feng, W. PVA Hydrogel Properties for Biomedical Application. J. Mech. Behav. Biomed. Mater. 2011, 4, 1228–1233. [Google Scholar] [CrossRef] [PubMed]
  36. Nkhwa, S.; Lauriaga, K.F.; Kemal, E.; Deb, S. Poly(Vinyl Alcohol): Physical Approaches to Designing Biomaterials for Biomedical Applications. Conf. Pap. Sci. 2014, 2014, 403472. [Google Scholar] [CrossRef]
  37. Zare, M.; Ghomi, E.R.; Venkatraman, P.D.; Ramakrishna, S. Silicone-based Biomaterials for Biomedical Applications: Antimicrobial Strategies and 3D Printing Technologies. J. Appl. Polym. Sci. 2021, 138, 50969. [Google Scholar] [CrossRef]
  38. Ding, K.; Wang, Y.; Liu, S.; Wang, S.; Mi, J. Preparation of Medical Hydrophilic and Antibacterial Silicone Rubber via Surface Modification. RSC Adv. 2021, 11, 39950–39957. [Google Scholar] [CrossRef] [PubMed]
  39. Carvalho, J.R.G.; Conde, G.; Antonioli, M.L.; Santana, C.H.; Littiere, T.O.; Dias, P.P.; Chinelatto, M.A.; Canola, P.A.; Zara, F.J.; Ferraz, G.C. Long-Term Evaluation of Poly (Lactic Acid) (PLA) Implants in a Horse: An Experimental Pilot Study. Molecules 2021, 26, 7224. [Google Scholar] [CrossRef]
  40. Singhvi, M.S.; Zinjarde, S.S.; Gokhale, D.V. Polylactic Acid: Synthesis and Biomedical Applications. J. Appl. Microbiol. 2019, 127, 1612–1626. [Google Scholar] [CrossRef]
  41. Casalini, T.; Rossi, F.; Castrovinci, A.; Perale, G. A Perspective on Polylactic Acid-Based Polymers Use for Nanoparticles Synthesis and Applications. Front. Bioeng. Biotechnol. 2019, 7, 259. [Google Scholar] [CrossRef]
  42. Chen, T.; Zhao, X.; Weng, Y. Self-Assembled Polylactic Acid (PLA): Synthesis, Properties and Biomedical Applications. Front. Chem. 2022, 10, 1107620. [Google Scholar] [CrossRef]
  43. Ebrahimi, F.; Ramezani Dana, H. Poly Lactic Acid (PLA) Polymers: From Properties to Biomedical Applications. Int. J. Polym. Mater. 2022, 71, 1117–1130. [Google Scholar] [CrossRef]
  44. Pawar, R.P.; Tekale, S.U.; Shisodia, S.U.; Totre, J.T.; Domb, J.A. Biomedical Applications of Poly (Lactic Acid). Rec. Pat. Regen. Med. 2014, 4, 40–51. [Google Scholar] [CrossRef]
  45. Ludwicka, K.; Kolodziejczyk, M.; Gendaszewska-Darmach, E.; Chrzanowski, M.; Jedrzejczak-Krzepkowska, M.; Rytczak, P.; Bielecki, S. Stable Composite of Bacterial Nanocellulose and Perforated Polypropylene Mesh for Biomedical Applications. J. Biomed. Mater. Res. B Appl. Biomater. 2019, 107, 978–987. [Google Scholar] [CrossRef] [PubMed]
  46. Acik, G.; Altinkok, C.; Acik, B. Biodegradable and Antibacterial Chlorinated Polypropylene/Chitosan Based Composite Films for Biomedical Applications. Polym. Bull. 2022, 79, 9997–10011. [Google Scholar] [CrossRef]
  47. Ali, U.; Karim, K.J.B.A.; Buang, N.A. A Review of the Properties and Applications of Poly (Methyl methacrylate) (PMMA). Polym. Rev. 2015, 55, 678–705. [Google Scholar] [CrossRef]
  48. Frazer, R.Q.; Byron, R.T.; Osborne, P.B.; West, K.P. PMMA: An Essential Material in Medicine and Dentistry. J. Long Term Eff. Med. Implants 2005, 15, 629–639. [Google Scholar] [CrossRef] [PubMed]
  49. Díez-Pascual, A.M. PMMA-Based Nanocomposites for Odontology Applications: A State-of-the-Art. Int. J. Mol. Sci. 2022, 23, 10288. [Google Scholar] [CrossRef] [PubMed]
  50. Leigh, J.A. Use of PMMA in Expansion Dental Implants. J. Biomed. Mater. Res. 1975, 9, 233–242. [Google Scholar] [CrossRef]
  51. Wang, H.; Lu, T.; Meng, F.; Zhu, H.; Liu, X. Enhanced Osteoblast Responses to Poly Ether Ether Ketone Surface Modified by Water Plasma Immersion Ion Implantation. Colloids Surf. B Biointerfaces 2014, 117, 89–97. [Google Scholar] [CrossRef]
  52. Alqurashi, H.; Khurshid, Z.; Syed, A.U.Y.; Rashid Habib, S.; Rokaya, D.; Zafar, M.S. Polyetherketoneketone (PEKK): An Emerging Biomaterial for Oral Implants and Dental Prostheses. J. Adv. Res. 2021, 28, 87–95. [Google Scholar] [CrossRef] [PubMed]
  53. Pietrocola, G.; Campoccia, D.; Motta, C.; Montanaro, L.; Arciola, C.R.; Speziale, P. Colonization and Infection of Indwelling Medical Devices by Staphylococcus Aureus with an Emphasis on Orthopedic Implants. Int. J. Mol. Sci. 2022, 23, 5958. [Google Scholar] [CrossRef] [PubMed]
  54. Ardila, C.M.; Vivares-Builes, A.M. Antibiotic Resistance in Patients with Peri-Implantitis: A Systematic Scoping Review. Int. J. Environ. Res. Public Health 2022, 19, 15609. [Google Scholar] [CrossRef] [PubMed]
  55. Momand, P.; Becktor, J.P.; Naimi-Akbar, A.; Tobin, G.; Götrick, B. Effect of Antibiotic Prophylaxis in Dental Implant Surgery: A Multicenter Placebo-controlled Double-blinded Randomized Clinical Trial. Clin. Implant Dent. Relat. Res. 2022, 24, 116–124. [Google Scholar] [CrossRef] [PubMed]
  56. Rezić, I.; Majdak, M.; Ljoljić Bilić, V.; Pokrovac, I.; Martinaga, L.; Somogyi Škoc, M.; Kosalec, I. Development of Antibacterial Protective Coatings Active against MSSA and MRSA on Biodegradable Polymers. Polymers 2021, 13, 659. [Google Scholar] [CrossRef] [PubMed]
  57. Dickson, M.N.; Liang, E.I.; Rodriguez, L.A.; Vollereaux, N.; Yee, A.F. Nanopatterned Polymer Surfaces with Bactericidal Properties. Biointerphases 2015, 10, 021010. [Google Scholar] [CrossRef] [PubMed]
  58. Trott, A.T. Instruments, Suture Materials, and Closure Choices. In Wounds and Lacerations; Elsevier: Amsterdam, The Netherlands, 2012; pp. 82–94. [Google Scholar]
  59. Palispis, W.A.; Gupta, R. Biologic Augmentation in Peripheral Nerve Repair. In Biologics in Orthopaedic Surgery; Elsevier: Amsterdam, The Netherlands, 2019; pp. 141–163. [Google Scholar]
  60. Manoukian, O.S.; Sardashti, N.; Stedman, T.; Gailiunas, K.; Ojha, A.; Penalosa, A.; Mancuso, C.; Hobert, M.; Kumbar, S.G. Biomaterials for Tissue Engineering and Regenerative Medicine. In Encyclopedia of Biomedical Engineering; Elsevier: Amsterdam, The Netherlands, 2019; pp. 462–482. [Google Scholar]
  61. Houchin, M.L.; Topp, E.M. Physical Properties of PLGA Films during Polymer Degradation. J. Appl. Polym. Sci. 2009, 114, 2848–2854. [Google Scholar] [CrossRef]
  62. Vignali, A.; Iannace, S.; Falcone, G.; Utzeri, R.; Stagnaro, P.; Bertini, F. Lightweight Poly(ε-Caprolactone) Composites with Surface Modified Hollow Glass Microspheres for Use in Rotational Molding: Thermal, Rheological and Mechanical Properties. Polymers 2019, 11, 624. [Google Scholar] [CrossRef]
  63. Zhou, Z.-M.; Wang, K.; Lin, K.-W.; Wang, Y.-H.; Li, J.-Z. Influence of Characteristics of Thermoplastic Polyurethane on Graphene-Thermoplastic Polyurethane Composite Film. Micromachines 2021, 12, 129. [Google Scholar] [CrossRef]
  64. Wang, W.; Wang, C. Polyurethane for Biomedical Applications: A Review of Recent Developments. In The Design and Manufacture of Medical Devices; Elsevier: Amsterdam, The Netherlands, 2012; pp. 115–151. [Google Scholar]
  65. Vermette, P.; Griesser, H.J.; Laroche, G.; Guidoin, R. Biomedical Applications of Polyurethanes. In Tissue Engineering Intelligence Unit 6 Landes Bioscience; Landes Bioscience: Georgetown, TX, USA, 2001. [Google Scholar]
  66. Hoque, M.B.; Solaiman; Alam, A.B.M.H.; Mahmud, H.; Nobi, A. Mechanical, Degradation and Water Uptake Properties of Fabric Reinforced Polypropylene Based Composites: Effect of Alkali on Composites. Fibers 2018, 6, 94. [Google Scholar] [CrossRef]
  67. Islam, N.M.; Othman, N.; Ahmad, Z.; Ismail, H. Effect of Pro-Degradant Additives Concentration on Aging Properties of Polypropylene Films. Polym. Plast. Technol. Eng. 2010, 49, 272–278. [Google Scholar] [CrossRef]
  68. Premkumar, J.; SonicaSree, K.; Sudhakar, T. Polymers in Biomedical Use. In Handbook of Polymer and Ceramic Nanotechnology; Springer International Publishing: Cham, Switzerland, 2021; pp. 1–28. [Google Scholar]
  69. Paradossi, G.; Cavalieri, F.; Chiessi, E.; Spagnoli, C.; Cowman, M.K.J. Poly(vinyl alcohol) as versatile biomaterial for potential biomedical applications. Mater. Sci. Mater. Med. 2003, 14, 687–691. [Google Scholar] [CrossRef] [PubMed]
  70. Peng, C.; Chen, G. Preparation and Assessment of Heat-Treated α-Chitin Nanowhiskers Reinforced Poly(Viny alcohol) Film for Packaging Application. Materials 2018, 11, 1883. [Google Scholar] [CrossRef] [PubMed]
  71. Hashim, H.; El-Mekawey, F.; El-Kashef, H.; Ghazy, R. Determination of Scattering Parameters of Polyvinyl Alcohol by Static Laser Scattering. Beni-Suef Univ. J. Basic Appl. Sci. 2014, 3, 203–208. [Google Scholar] [CrossRef]
  72. Tubbs, R.K. Melting Point and Heat of Fusion of Poly(Vinyl Alcohol). J. Polym. Sci. A 1965, 3, 4181–4189. [Google Scholar] [CrossRef]
  73. Gundapaneni, D.; Goswami, T. Thermal Isotherms in PMMA and Cell Necrosis during Total Hip Arthroplasty. J. Appl. Biomater. Funct. Mater. 2014, 12, 193–202. [Google Scholar] [CrossRef] [PubMed]
  74. Clet, J.A.G.; Liou, N.-S.; Weng, C.-H.; Lin, Y.-S. A Parametric Study for Tensile Properties of Silicone Rubber Specimen Using the Bowden-Type Silicone Printer. Materials 2022, 15, 1729. [Google Scholar] [CrossRef] [PubMed]
  75. Zhao, J.; Jiang, N.; Zhang, D.; He, B.; Chen, X. Study on Optimization of Damping Performance and Damping Temperature Range of Silicone Rubber by Polyborosiloxane Gel. Polymers 2020, 12, 1196. [Google Scholar] [CrossRef]
  76. Verma, S.; Sharma, N.; Kango, S.; Sharma, S. Developments of PEEK (Polyetheretherketone) as a Biomedical Material: A Focused Review. Eur. Polym. J. 2021, 147, 110295. [Google Scholar] [CrossRef]
  77. Senra, M.R.; de Marques, M.F.V.; Monteiro, S.N. Poly (Ether-Ether-Ketone) for Biomedical Applications: From Enhancing Bioactivity to Reinforced-Bioactive Composites—An Overview. Polymers 2023, 15, 373. [Google Scholar] [CrossRef]
  78. Mbogori, M.; Vaish, A.; Vaishya, R.; Haleem, A.; Javaid, M. Poly-Ether-Ether-Ketone (PEEK) in Orthopaedic Practice- A Current Concept Review. J. Orthop. Rep. 2022, 1, 3–7. [Google Scholar] [CrossRef]
  79. Sakihara, M.; Taira, Y.; Sawase, T. Effects of Sulfuric and Vinyl Sulfonic Acid Etchants on Bond Strength of Resin Composite to Polyetherketoneketone. Odontology 2019, 107, 158–164. [Google Scholar] [CrossRef] [PubMed]
  80. Çaykara, T.; Sande, M.G.; Azoia, N.; Rodrigues, L.R.; Silva, C.J. Exploring the Potential of Polyethylene Terephthalate in the Design of Antibacterial Surfaces. Med. Microbiol. Immunol. 2020, 209, 363–372. [Google Scholar] [CrossRef] [PubMed]
  81. Panowicz, R.; Konarzewski, M.; Durejko, T.; Szala, M.; Łazińska, M.; Czerwińska, M.; Prasuła, P. Properties of Polyethylene Terephthalate (PET) after Thermo-Oxidative Aging. Materials 2021, 14, 3833. [Google Scholar] [CrossRef] [PubMed]
  82. Dhaka, V.; Singh, S.; Anil, A.G.; Sunil Kumar Naik, T.S.; Garg, S.; Samuel, J.; Kumar, M.; Ramamurthy, P.C.; Singh, J. Occurrence, Toxicity and Remediation of Polyethylene Terephthalate Plastics. A Review. Environ. Chem. Lett. 2022, 20, 1777–1800. [Google Scholar] [CrossRef] [PubMed]
  83. Aumsuwan, N.; Heinhorst, S.; Urban, M.W. Antibacterial Surfaces on Expanded Polytetrafluoroethylene; Penicillin Attachment. Biomacromolecules 2007, 8, 713–718. [Google Scholar] [CrossRef] [PubMed]
  84. Kalaycıoğlu, Z.; Torlak, E.; Akın-Evingür, G.; Özen, İ.; Erim, F.B. Antimicrobial and Physical Properties of Chitosan Films Incorporated with Turmeric Extract. Int. J. Biol. Macromol. 2017, 101, 882–888. [Google Scholar] [CrossRef] [PubMed]
  85. Shahidi, F.; Arachchi, J.K.V.; Jeon, Y.-J. Food Applications of Chitin and Chitosans. Trends Food Sci. Technol. 1999, 10, 37–51. [Google Scholar] [CrossRef]
  86. Shih, P.-Y.; Liao, Y.-T.; Tseng, Y.-K.; Deng, F.-S.; Lin, C.-H. A Potential Antifungal Effect of Chitosan against Candida Albicans Is Mediated via the Inhibition of SAGA Complex Component Expression and the Subsequent Alteration of Cell Surface Integrity. Front. Microbiol. 2019, 10, 602. [Google Scholar] [CrossRef]
  87. Azuma, K.; Osaki, T.; Minami, S.; Okamoto, Y. Anticancer and Anti-Inflammatory Properties of Chitin and Chitosan Oligosaccharides. J. Funct. Biomater. 2015, 6, 33–49. [Google Scholar] [CrossRef]
  88. Cho, Y.I.; No, H.K.; Meyers, S.P. Physicochemical Characteristics and Functional Properties of Various Commercial Chitin and Chitosan Products. J. Agric. Food Chem. 1998, 46, 3839–3843. [Google Scholar] [CrossRef]
  89. Dong, Y.; Ruan, Y.; Wang, H.; Zhao, Y.; Bi, D. Studies on Glass Transition Temperature of Chitosan with Four Techniques. J. Appl. Polym. Sci. 2004, 93, 1553–1558. [Google Scholar] [CrossRef]
  90. Jiang, S.; Qiao, C.; Wang, X.; Li, Z.; Yang, G. Structure and Properties of Chitosan/Sodium Dodecyl Sulfate Composite Films. RSC Adv. 2022, 12, 3969–3978. [Google Scholar] [CrossRef] [PubMed]
  91. Cabral, J.P.S. Water Microbiology. Bacterial Pathogens and Water. Int. J. Environ. Res. Public Health 2010, 7, 3657–3703. [Google Scholar] [CrossRef] [PubMed]
  92. Kreve, S.; Reis, A.C.D. Bacterial Adhesion to Biomaterials: What Regulates This Attachment? A Review. Jpn. Dent. Sci. Rev. 2021, 57, 85–96. [Google Scholar] [CrossRef]
  93. Renner, L.D.; Weibel, D.B. Physicochemical Regulation of Biofilm Formation. MRS Bull. 2011, 36, 347–355. [Google Scholar] [CrossRef] [PubMed]
  94. Zheng, S.; Bawazir, M.; Dhall, A.; Kim, H.-E.; He, L.; Heo, J.; Hwang, G. Implication of Surface Properties, Bacterial Motility, and Hydrodynamic Conditions on Bacterial Surface Sensing and Their Initial Adhesion. Front. Bioeng. Biotechnol. 2021, 9, 643722. [Google Scholar] [CrossRef] [PubMed]
  95. Dunne, W.M., Jr. Bacterial Adhesion: Seen Any Good Biofilms Lately? Clin. Microbiol. Rev. 2002, 15, 155–166. [Google Scholar] [CrossRef]
  96. Boks, N.P.; Busscher, H.J.; van der Mei, H.C.; Norde, W. Bond-Strengthening in Staphylococcal Adhesion to Hydrophilic and Hydrophobic Surfaces Using Atomic Force Microscopy. Langmuir 2008, 24, 12990–12994. [Google Scholar] [CrossRef]
  97. Wilson, W.W.; Wade, M.M.; Holman, S.C.; Champlin, F.R. Status of Methods for Assessing Bacterial Cell Surface Charge Properties Based on Zeta Potential Measurements. J. Microbiol. Methods 2001, 43, 153–164. [Google Scholar] [CrossRef]
  98. Hayashi, H.; Seiki, H.; Tsuneda, S.; Hirata, A.; Sasaki, H. Influence of Growth Phase on Bacterial Cell Electrokinetic Characteristics Examined by Soft Particle Electrophoresis Theory. J. Colloid Interface Sci. 2003, 264, 565–568. [Google Scholar] [CrossRef] [PubMed]
  99. Tuson, H.H.; Weibel, D.B. Bacteria–Surface Interactions. Soft Matter 2013, 9, 4368. [Google Scholar] [CrossRef] [PubMed]
  100. Limoli, D.H.; Jones, C.J.; Wozniak, D.J. Bacterial Extracellular Polysaccharides in Biofilm Formation and Function. Microbiol. Spectr. 2015, 3, 223–247. [Google Scholar] [CrossRef] [PubMed]
  101. 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]
  102. Watson, G.S.; Green, D.W.; Schwarzkopf, L.; Li, X.; Cribb, B.W.; Myhra, S.; Watson, J.A. A Gecko Skin Micro/Nano Structure—A Low Adhesion, Superhydrophobic, Anti-Wetting, Self-Cleaning, Biocompatible, Antibacterial Surface. Acta Biomater. 2015, 21, 109–122. [Google Scholar] [CrossRef] [PubMed]
  103. Campoccia, D.; Montanaro, L.; Arciola, C.R. The Significance of Infection Related to Orthopedic Devices and Issues of Antibiotic Resistance. Biomaterials 2006, 27, 2331–2339. [Google Scholar] [CrossRef] [PubMed]
  104. Ventola, C.L. The Antibiotic Resistance Crisis: Part 1: Causes and Threats. Pharm. Ther. 2015, 40, 277–283. [Google Scholar]
  105. Montanaro, L.; Speziale, P.; Campoccia, D.; Ravaioli, S.; Cangini, I.; Pietrocola, G.; Giannini, S.; Arciola, C.R. Scenery of Staphylococcus Implant Infections in Orthopedics. Future Microbiol. 2011, 6, 1329–1349. [Google Scholar] [CrossRef]
  106. 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]
  107. Khatoon, Z.; McTiernan, C.D.; Suuronen, E.J.; Mah, T.-F.; Alarcon, E.I. Bacterial Biofilm Formation on Implantable Devices and Approaches to Its Treatment and Prevention. Heliyon 2018, 4, e01067. [Google Scholar] [CrossRef]
  108. Kandi, V.; Vadakedath, S. Implant-Associated Infections: A Review of the Safety of Cardiac Implants. Cureus 2020, 12, e12267. [Google Scholar] [CrossRef] [PubMed]
  109. Nicolle, L.E. Catheter Associated Urinary Tract Infections. Antimicrob. Resist. Infect. Control 2014, 3, 23. [Google Scholar] [CrossRef] [PubMed]
  110. Green, D.W.; Lee, K.K.-H.; Watson, J.A.; Kim, H.-Y.; Yoon, K.-S.; Kim, E.-J.; Lee, J.-M.; Watson, G.S.; Jung, H.-S. High Quality Bioreplication of Intricate Nanostructures from a Fragile Gecko Skin Surface with Bactericidal Properties. Sci. Rep. 2017, 7, 41023. [Google Scholar] [CrossRef] [PubMed]
  111. Ishihama, H.; Ishii, K.; Nagai, S.; Kakinuma, H.; Sasaki, A.; Yoshioka, K.; Kuramoto, T.; Shiono, Y.; Funao, H.; Isogai, N.; et al. An Antibacterial Coated Polymer Prevents Biofilm Formation and Implant-Associated Infection. Sci. Rep. 2021, 11, 3602. [Google Scholar] [CrossRef] [PubMed]
  112. Jahanmard, F.; Dijkmans, F.M.; Majed, A.; Vogely, H.C.; van der Wal, B.C.H.; Stapels, D.A.C.; Ahmadi, S.M.; Vermonden, T.; Amin Yavari, S. Toward Antibacterial Coatings for Personalized Implants. ACS Biomater. Sci. Eng. 2020, 6, 5486–5492. [Google Scholar] [CrossRef] [PubMed]
  113. Akshaya, S.; Rowlo, P.K.; Dukle, A.; Nathanael, A.J. Antibacterial Coatings for Titanium Implants: Recent Trends and Future Perspectives. Antibiotics 2022, 11, 1719. [Google Scholar] [CrossRef] [PubMed]
  114. Escobar, A.; Muzzio, N.; Moya, S.E. Antibacterial Layer-by-Layer Coatings for Medical Implants. Pharmaceutics 2020, 13, 16. [Google Scholar] [CrossRef] [PubMed]
  115. Chan, Y.; Wu, X.H.; Chieng, B.W.; Ibrahim, N.A.; Then, Y.Y. Superhydrophobic Nanocoatings as Intervention against Biofilm-Associated Bacterial Infections. Nanomaterials 2021, 11, 1046. [Google Scholar] [CrossRef]
  116. Steinerová, D.; Kalendová, A.; Machotová, J.; Knotek, P.; Humpolíček, P.; Vajdák, J.; Slang, S.; Krejčová, A.; Beneš, L.; Wolff-Fabris, F. Influence of Metal Oxide Nanoparticles as Antimicrobial Additives Embedded in Waterborne Coating Binders Based on Self-Crosslinking Acrylic Latex. Coatings 2022, 12, 1445. [Google Scholar] [CrossRef]
  117. Gallingani, T.; Resca, E.; Dominici, M.; Gavioli, G.; Laurita, R.; Liguori, A.; Mari, G.; Ortolani, L.; Pericolini, E.; Sala, A.; et al. A New Strategy to Prevent Biofilm and Clot Formation in Medical Devices: The Use of Atmospheric Non-Thermal Plasma Assisted Deposition of Silver-Based Nanostructured Coatings. PLoS ONE 2023, 18, e0282059. [Google Scholar] [CrossRef]
  118. Kara, F.; Aksoy, E.A.; Aksoy, S.; Hasirci, N. Coating of silver nanoparticles on polyurethane film surface by green chemistry approach and investigation of antibacterial activity against S. epidermidis. J. Bioact. Compat. Polym. 2022, 37, 205–219. [Google Scholar] [CrossRef]
  119. Zhang, J.; Gao, M.; Gao, W.; Yang, P.; Meng, F.; Liu, Q.; Chang, H. Functional Silver Nanoparticles as Broad-Spectrum Antimicrobial Agents. New J. Chem. 2022, 46, 16387–16393. [Google Scholar] [CrossRef]
  120. Gold, K.; Slay, B.; Knackstedt, M.; Gaharwar, A.K. Antimicrobial Activity of Metal and Metal-oxide Based Nanoparticles. Adv. Ther. 2018, 1, 1700033. [Google Scholar] [CrossRef]
  121. Azam, A.; Ahmed, A.S.; Oves, M.; Khan, M.S.; Habib, S.S.; Memic, A. Antimicrobial Activity of Metal Oxide Nanoparticles against Gram-Positive and Gram-Negative Bacteria: A Comparative Study. Int. J. Nano Med. 2012, 7, 6003–6009. [Google Scholar] [CrossRef] [PubMed]
  122. Mohammed, H.; Kumar, A.; Bekyarova, E.; Al-Hadeethi, Y.; Zhang, X.; Chen, M.; Ansari, M.S.; Cochis, A.; Rimondini, L. Antimicrobial Mechanisms and Effectiveness of Graphene and Graphene-Functionalized Biomaterials. A Scope Review. Front. Bioeng. Biotechnol. 2020, 8, 465. [Google Scholar] [CrossRef] [PubMed]
  123. Kumar, P.; Huo, P.; Zhang, R.; Liu, B. Antibacterial Properties of Graphene-Based Nanomaterials. Nanomaterials 2019, 9, 737. [Google Scholar] [CrossRef] [PubMed]
  124. Shang, L.; Qi, Y.; Lu, H.; Pei, H.; Li, Y.; Qu, L.; Wu, Z.; Zhang, W. Graphene and Graphene Oxide for Tissue Engineering and Regeneration. In Theranostic Bionanomaterials; Elsevier: Amsterdam, The Netherlands, 2019; pp. 165–185. [Google Scholar]
  125. Azizi-Lalabadi, M.; Hashemi, H.; Feng, J.; Jafari, S.M. Carbon Nanomaterials against Pathogens; the Antimicrobial Activity of Carbon Nanotubes, Graphene/Graphene Oxide, Fullerenes, and Their Nanocomposites. Adv. Colloid Interface Sci. 2020, 284, 102250. [Google Scholar] [CrossRef] [PubMed]
  126. Reynoso, E.; Durantini, A.M.; Solis, C.A.; Macor, L.P.; Otero, L.A.; Gervaldo, M.A.; Durantini, E.N.; Heredia, D.A. Photoactive Antimicrobial Coating Based on a PEDOT-Fullerene C60 Polymeric Dyad. RSC Adv. 2021, 11, 23519–23532. [Google Scholar] [CrossRef] [PubMed]
  127. Lyon, D.Y.; Adams, L.K.; Falkner, J.C.; Alvarezt, P.J.J. Antibacterial Activity of Fullerene Water Suspensions: Effects of Preparation Method and Particle Size. Environ. Sci. Technol. 2006, 40, 4360–4366. [Google Scholar] [CrossRef] [PubMed]
  128. Guo, L.; Yuan, W.; Lu, Z.; Li, C.M. Polymer/Nanosilver Composite Coatings for Antibacterial Applications. Colloids Surf. A Physicochem. Eng. Asp. 2013, 439, 69–83. [Google Scholar] [CrossRef]
  129. Ghazzy, A.; Naik, R.R.; Shakya, A.K. Metal–Polymer Nanocomposites: A Promising Approach to Antibacterial Materials. Polymers 2023, 15, 2167. [Google Scholar] [CrossRef] [PubMed]
  130. Patil, D.; Wasson, M.K.; Aravindan, S.; Perumal, V.; Rao, P.V. Fabrication of Silver Nanoparticles-Embedded Antibacterial Polymer Surface through Thermal Annealing and Soft Molding Technique. Mater. Res. Express 2019, 6, 045010. [Google Scholar] [CrossRef]
  131. Li, X.; Cheung, G.S.; Watson, G.S.; Watson, J.A.; Lin, S.; Schwarzkopf, L.; Green, D.W. The Nanotipped Hairs of Gecko Skin and Biotemplated Replicas Impair and/or Kill Pathogenic Bacteria with High Efficiency. Nanoscale 2016, 8, 18860–18869. [Google Scholar] [CrossRef] [PubMed]
  132. Linklater, D.P.; Saita, S.; Murata, T.; Yanagishita, T.; Dekiwadia, C.; Crawford, R.J.; Masuda, H.; Kusaka, H.; Ivanova, E.P. Nanopillar Polymer Films as Antibacterial Packaging Materials. ACS Appl. Nano Mater. 2022, 5, 2578–2591. [Google Scholar] [CrossRef]
  133. Sengul, A.B.; Asmatulu, E. Toxicity of Metal and Metal Oxide Nanoparticles: A Review. Environ. Chem. Lett. 2020, 18, 1659–1683. [Google Scholar] [CrossRef]
  134. Ng, C.T.; Yong, L.Q.; Hande, M.P.; Ong, C.N.; Yu, L.E.; Bay, B.H.; Baeg, G.H. Zinc Oxide Nanoparticles Exhibit Cytotoxicity and Genotoxicity through Oxidative Stress Responses in Human Lung Fibroblasts and Drosophila Melanogaster. Int. J. Nano Med. 2017, 12, 1621–1637. [Google Scholar] [CrossRef] [PubMed]
  135. Tan, K.X.; Barhoum, A.; Pan, S.; Danquah, M.K. Risks and Toxicity of Nanoparticles and Nanostructured Materials. In Emerging Applications of Nanoparticles and Architecture Nanostructures; Elsevier: Amsterdam, The Netherlands, 2018; pp. 121–139. [Google Scholar]
  136. De Berardis, B.; Civitelli, G.; Condello, M.; Lista, P.; Pozzi, R.; Arancia, G.; Meschini, S. Exposure to ZnO Nanoparticles Induces Oxidative Stress and Cytotoxicity in Human Colon Carcinoma Cells. Toxicol. Appl. Pharmacol. 2010, 246, 116–127. [Google Scholar] [CrossRef] [PubMed]
  137. Guan, R.; Kang, T.; Lu, F.; Zhang, Z.; Shen, H.; Liu, M. Cytotoxicity, Oxidative Stress, and Genotoxicity in Human Hepatocyte and Embryonic Kidney Cells Exposed to ZnO Nanoparticles. Nanoscale Res. Lett. 2012, 7, 602. [Google Scholar] [CrossRef]
  138. Haase, A.; Tentschert, J.; Jungnickel, H.; Graf, P.; Mantion, A.; Draude, F.; Plendl, J.; Goetz, M.E.; Galla, S.; Mašić, A.; et al. Toxicity of Silver Nanoparticles in Human Macrophages: Uptake, Intracellular Distribution and Cellular Responses. J. Phys. Conf. Ser. 2011, 304, 012030. [Google Scholar] [CrossRef]
  139. Mancuso, G.; Midiri, A.; Gerace, E.; Biondo, C. Bacterial Antibiotic Resistance: The Most Critical Pathogens. Pathogens 2021, 10, 1310. [Google Scholar] [CrossRef]
  140. Uddin, T.M.; Chakraborty, A.J.; Khusro, A.; Zidan, B.M.R.M.; Mitra, S.; Emran, T.B.; Dhama, K.; Ripon, M.K.H.; Gajdács, M.; Sahibzada, M.U.K.; et al. Antibiotic Resistance in Microbes: History, Mechanisms, Therapeutic Strategies and Future Prospects. J. Infect. Public Health 2021, 14, 1750–1766. [Google Scholar] [CrossRef]
  141. Jaggessar, A.; Shahali, H.; Mathew, A.; Yarlagadda, P.K.D.V. Bio-Mimicking Nano and Micro-Structured Surface Fabrication for Antibacterial Properties in Medical Implants. J. Nano Biotechnol. 2017, 15, 64. [Google Scholar] [CrossRef] [PubMed]
  142. Mainwaring, D.E.; Nguyen, S.H.; Webb, H.; Jakubov, T.; Tobin, M.; Lamb, R.N.; Wu, A.H.-F.; Marchant, R.; Crawford, R.J.; Ivanova, E.P. The Nature of Inherent Bactericidal Activity: Insights from the Nanotopology of Three Species of Dragonfly. Nanoscale 2016, 8, 6527–6534. [Google Scholar] [CrossRef] [PubMed]
  143. Nguyen, S.H.; Webb, H.K.; Mahon, P.J.; Crawford, R.J.; Ivanova, E.P. Natural Insect and Plant Micro-/Nanostructsured Surfaces: An Excellent Selection of Valuable Templates with Superhydrophobic and Self-Cleaning Properties. Molecules 2014, 19, 13614–13630. [Google Scholar] [CrossRef] [PubMed]
  144. Oopath, S.V.; Baji, A.; Abtahi, M.; Luu, T.Q.; Vasilev, K.; Truong, V.K. Nature-inspired Biomimetic Surfaces for Controlling Bacterial Attachment and Biofilm Development. Adv. Mater. Interfaces 2023, 10, 2201425. [Google Scholar] [CrossRef]
  145. 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]
  146. Ivanova, E.P.; Hasan, J.; Webb, H.K.; Gervinskas, G.; Juodkazis, S.; Truong, V.K.; Wu, A.H.F.; Lamb, R.N.; Baulin, V.A.; Watson, G.S.; et al. Bactericidal Activity of Black Silicon. Nat. Commun. 2013, 4, 2838. [Google Scholar] [CrossRef]
  147. Vassallo, E.; Pedroni, M.; Silvetti, T.; Morandi, S.; Toffolatti, S.; Angella, G.; Brasca, M. Bactericidal Performance of Nanostructured Surfaces by Fluorocarbon Plasma. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 80, 117–121. [Google Scholar] [CrossRef]
  148. 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]
  149. Truong, V.K.; Geeganagamage, N.M.; Baulin, V.A.; Vongsvivut, J.; Tobin, M.J.; Luque, P.; Crawford, R.J.; Ivanova, E.P. The Susceptibility of Staphylococcus Aureus CIP 65.8 and Pseudomonas Aeruginosa ATCC 9721 Cells to the Bactericidal Action of Nanostructured Calopteryx Haemorrhoidalis Damselfly Wing Surfaces. Appl. Microbiol. Biotechnol. 2017, 101, 4683–4690. [Google Scholar] [CrossRef]
  150. Linklater, D.P.; Baulin, V.A.; Juodkazis, S.; Crawford, R.J.; Stoodley, P.; Ivanova, E.P. Mechano-Bactericidal Actions of Nanostructured Surfaces. Nat. Rev. Microbiol. 2021, 19, 8–22. [Google Scholar] [CrossRef] [PubMed]
  151. Li, X. Bactericidal Mechanism of Nanopatterned Surfaces. Phys. Chem. Chem. Phys. 2016, 18, 1311–1316. [Google Scholar] [CrossRef] [PubMed]
  152. Jenkins, J.; Mantell, J.; Neal, C.; Gholinia, A.; Verkade, P.; Nobbs, A.H.; Su, B. Antibacterial Effects of Nanopillar Surfaces Are Mediated by Cell Impedance, Penetration and Induction of Oxidative Stress. Nat. Commun. 2020, 11, 1626. [Google Scholar] [CrossRef]
  153. Ivanova, E.P.; Linklater, D.P.; Werner, M.; Baulin, V.A.; Xu, X.; Vrancken, N.; Rubanov, S.; Hanssen, E.; Wandiyanto, J.; Truong, V.K.; et al. The Multi-Faceted Mechano-Bactericidal Mechanism of Nanostructured Surfaces. Proc. Natl. Acad. Sci. USA 2020, 117, 12598–12605. [Google Scholar] [CrossRef] [PubMed]
  154. Nowlin, K.; Boseman, A.; Covell, A.; LaJeunesse, D. Adhesion-Dependent Rupturing of Saccharomyces cerevisiae on Biological Antimicrobial Nanostructured Surfaces. J. R. Soc. Interface 2015, 12, 20140999. [Google Scholar] [CrossRef] [PubMed]
  155. Ivanova, E.P.; Linklater, D.P.; Aburto-Medina, A.; Le, P.; Baulin, V.A.; Khuong Duy Nguyen, H.; Curtain, R.; Hanssen, E.; Gervinskas, G.; Hock Ng, S.; et al. Antifungal versus Antibacterial Defence of Insect Wings. J. Colloid Interface Sci. 2021, 603, 886–897. [Google Scholar] [CrossRef] [PubMed]
  156. Yamamoto, M.; Nishikawa, N.; Mayama, H.; Nonomura, Y.; Yokojima, S.; Nakamura, S.; Uchida, K. Theoretical Explanation of the Lotus Effect: Superhydrophobic Property Changes by Removal of Nanostructures from the Surface of a Lotus Leaf. Langmuir 2015, 31, 7355–7363. [Google Scholar] [CrossRef]
  157. Jiang, R.; Hao, L.; Song, L.; Tian, L.; Fan, Y.; Zhao, J.; Liu, C.; Ming, W.; Ren, L. Lotus-Leaf-Inspired Hierarchical Structured Surface with Non-Fouling and Mechanical Bactericidal Performances. Chem. Eng. J. 2020, 398, 125609. [Google Scholar] [CrossRef]
  158. Chen, Y. Nanofabrication by Electron Beam Lithography and Its Applications: A Review. Microelectron. Eng. 2015, 135, 57–72. [Google Scholar] [CrossRef]
  159. Bhadra, C.M.; Truong, V.K.; Pham, V.T.H.; Kobaisi, M.; Seniutinas, G.; Wang, J.Y.; Juodkazis, S.; Crawford, R.J.; Ivanova, E.P. Antibacterial Titanium Nanopatterned Arrays Inspired by Dragonfly Wings. Sci. Rep. 2015, 5, 16817. [Google Scholar] [CrossRef]
  160. Patil, D.; Golia, V.; Overland, M.; Stoller, M.; Chatterjee, K. Mechanobactericidal nanotopography on nitrile surfaces toward antimicrobial protective gear. ACS Macro Lett. 2022, 12, 227–233. [Google Scholar] [CrossRef]
  161. Hazell, G.; Fisher, L.E.; Murray, W.A.; Nobbs, A.H.; Su, B. Bioinspired Bactericidal Surfaces with Polymer Nanocone Arrays. J. Colloid Interface Sci. 2018, 528, 389–399. [Google Scholar] [CrossRef] [PubMed]
  162. Mo, S.; Mehrjou, B.; Tang, K.; Wang, H.; Huo, K.; Qasim, A.M.; Wang, G.; Chu, P.K. Dimensional-Dependent Antibacterial Behavior on Bioactive Micro/Nano Polyetheretherketone (PEEK) Arrays. Chem. Eng. J. 2020, 392, 123736. [Google Scholar] [CrossRef]
  163. Kolodziej, C.M.; Maynard, H.D. Electron-Beam Lithography for Patterning Biomolecules at the Micron and Nanometer Scale. Chem. Mater. 2012, 24, 774–780. [Google Scholar] [CrossRef]
  164. Xu, R.; Zeng, Z.; Lei, Y. Well-Defined Nanostructuring with Designable Anodic Aluminum Oxide Template. Nat. Commun. 2022, 13, 2435. [Google Scholar] [CrossRef] [PubMed]
  165. Liu, W.-D.; Yang, B. Patterned Surfaces for Biological Applications: A New Platform Using Two Dimensional Structures as Biomaterials. Chin. Chem. Lett. 2017, 28, 675–690. [Google Scholar] [CrossRef]
  166. Zhu, K.; Hu, G. Supercritical Hydrothermal Synthesis of Titanium Dioxide Nanostructures with Controlled Phase and Morphology. J. Supercrit. Fluids 2014, 94, 165–173. [Google Scholar] [CrossRef]
  167. Zhang, C.; Yang, C.; Ding, D. Deep Reactive Ion Etching of PMMA. Appl. Surf. Sci. 2004, 227, 139–143. [Google Scholar] [CrossRef]
  168. Chekurov, N.; Grigoras, K.; Sainiemi, L.; Peltonen, A.; Tittonen, I.; Franssila, S. Dry Fabrication of Microdevices by the Combination of Focused Ion Beam and Cryogenic Deep Reactive Ion Etching. J. Micromech. Microeng. 2010, 20, 085009. [Google Scholar] [CrossRef]
  169. Roy, A.; Chatterjee, K. Bactericidal Anisotropic Nanostructures on Titanium Fabricated by Maskless Dry Etching. ACS Appl. Nano Mater. 2022, 5, 4447–4461. [Google Scholar] [CrossRef]
  170. Ganjian, M.; Modaresifar, K.; Zhang, H.; Hagedoorn, P.-L.; Fratila-Apachitei, L.E.; Zadpoor, A.A. Reactive Ion Etching for Fabrication of Biofunctional Titanium Nanostructures. Sci. Rep. 2019, 9, 18815. [Google Scholar] [CrossRef]
  171. Kayes, M.I. Micro/Nano Structured Materials for Enhanced Device Performance and Antibacterial Applications. Ph.D. Thesis, University of Pittsburgh, Pittsburgh, PA, USA, 2019. [Google Scholar]
  172. Roy, A.; Patil, D.; Yarlagadda, P.K.D.V.; Chatterjee, K. Cooperative Stiffening of Flexible High Aspect Ratio Nanostructures Impart Mechanobactericidal Activity to Soft Substrates. J. Colloid Interface Sci. 2023, 652, 2127–2138. [Google Scholar] [CrossRef]
  173. Liu, W.; Liu, X.; Fangteng, J.; Wang, S.; Fang, L.; Shen, H.; Xiang, S.; Sun, H.; Yang, B. Bioinspired Polyethylene Terephthalate Nanocone Arrays with Underwater Superoleophobicity and Anti-Bioadhesion Properties. Nanoscale 2014, 6, 13845–13853. [Google Scholar] [CrossRef]
  174. Guo, L.J. Nanoimprint Lithography: Methods and Material Requirements. Adv. Mater. 2007, 19, 495–513. [Google Scholar] [CrossRef]
  175. Stewart, M.D.; Willson, C.G. Imprint Materials for Nanoscale Devices. MRS Bull. 2005, 30, 947–951. [Google Scholar] [CrossRef]
  176. Kim, S.; Jung, U.T.; Kim, S.-K.; Lee, J.-H.; Choi, H.S.; Kim, C.-S.; Jeong, M.Y. Nanostructured Multifunctional Surface with Antireflective and Antimicrobial Characteristics. ACS Appl. Mater. Interfaces 2015, 7, 326–331. [Google Scholar] [CrossRef]
  177. Glinsner, T.; Kreindl, G.; Kast, M. Nanoimprint Lithography. Opt. Photonik 2010, 5, 42–45. [Google Scholar] [CrossRef]
  178. Kooy, N.; Mohamed, K.; Pin, L.T.; Guan, O.S. A Review of Roll-to-Roll Nanoimprint Lithography. Nanoscale Res. Lett. 2014, 9, 320. [Google Scholar] [CrossRef]
  179. Oopath, S.V.; Martins, J.; Kakarla, A.B.; Kong, I.; Petrovski, S.; Baji, A. Rose Petal Mimetic Surfaces with Antibacterial Properties Produced Using Nanoimprint Lithography. ACS Appl. Bio Mater. 2023, 6, 2690–2697. [Google Scholar] [CrossRef]
  180. Fruncillo, S.; Su, X.; Liu, H.; Wong, L.S. Lithographic Processes for the Scalable Fabrication of Micro- and Nanostructures for Biochips and Biosensors. ACS Sens. 2021, 6, 2002–2024. [Google Scholar] [CrossRef]
  181. Sung, Y.H.; Kim, Y.D.; Choi, H.-J.; Shin, R.; Kang, S.; Lee, H. Fabrication of Superhydrophobic Surfaces with Nano-in-Micro Structures Using UV-Nanoimprint Lithography and Thermal Shrinkage Films. Appl. Surf. Sci. 2015, 349, 169–173. [Google Scholar] [CrossRef]
  182. Lü, L.; Fuh, J.Y.H.; Wong, Y.S. Fundamentals of Laser-Lithography Processes. In Laser-Induced Materials and Processes for Rapid Prototyping; Springer: Boston, MA, USA, 2001; pp. 9–38. [Google Scholar]
  183. Rebollar, E.; Castillejo, M.; Ezquerra, T.A. Laser Induced Periodic Surface Structures on Polymer Films: From Fundamentals to Applications. Eur. Polym. J. 2015, 73, 162–174. [Google Scholar] [CrossRef]
  184. Tawfick, S.; De Volder, M.; Copic, D.; Park, S.J.; Oliver, C.R.; Polsen, E.S.; Roberts, M.J.; Hart, A.J. Engineering of Micro- and Nanostructured Surfaces with Anisotropic Geometries and Properties. Adv. Mater. 2012, 24, 1628–1674. [Google Scholar] [CrossRef] [PubMed]
  185. Shahali, H.; Hasan, J.; Cheng, H.-H.; Ramarishna, S.; Yarlagadda, P.K. A Systematic Approach towards Biomimicry of Nanopatterned Cicada Wings on Titanium Using Electron Beam Lithography. Nanotechnology 2021, 32, 065301. [Google Scholar] [CrossRef] [PubMed]
  186. Shimizu, Y. Laser Interference Lithography for Fabrication of Planar Scale Gratings for Optical Metrology. Nano Manuf. Metrol. 2021, 4, 3–27. [Google Scholar] [CrossRef]
  187. Du, K.; Jiang, Y.; Huang, P.-S.; Ding, J.; Gao, T.; Choi, C.-H. Self-Formation of Polymer Nanostructures in Plasma Etching: Mechanisms and Applications. J. Micromech. Microeng. 2018, 28, 014006. [Google Scholar] [CrossRef]
  188. Liu, R.; Cao, L.; Liu, D.; Wang, L.; Saeed, S.; Wang, Z. Laser Interference Lithography—A Method for the Fabrication of Controlled Periodic Structures. Nanomaterials 2023, 13, 1818. [Google Scholar] [CrossRef]
  189. Franssila, S.; Tuomikoski, S. MEMS Lithography. In Handbook of Silicon Based MEMS Materials and Technologies; Elsevier: Amsterdam, The Netherlands, 2015; pp. 427–443. [Google Scholar]
  190. Quilis, N.G.; Hageneder, S.; Fossati, S.; Auer, S.K.; Venugopalan, P.; Bozdogan, A.; Petri, C.; Moreno-Cencerrado, A.; Toca-Herrera, J.L.; Jonas, U.; et al. UV-Laser Interference Lithography for Local Functionalization of Plasmonic Nanostructures with Responsive Hydrogel. J. Phys. Chem. C Nanomater. Interfaces 2020, 124, 3297–3305. [Google Scholar] [CrossRef]
  191. 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]
  192. Cheng, H.-C.; Huang, C.-F.; Lee, J.-L.; Lin, Y.; Shen, Y.-K. Morphological Characteristics of Nanostructured Surface Using Anodic Aluminum Template. Adv. Polym. Technol. 2016, 35, 152–161. [Google Scholar] [CrossRef]
  193. Sulka, G.D.; Parkoła, K.G. Temperature Influence on Well-Ordered Nanopore Structures Grown by Anodization of Aluminium in Sulphuric Acid. Electrochim. Acta 2007, 52, 1880–1888. [Google Scholar] [CrossRef]
  194. Zhou, F.; Baron-Wiecheć, A.; Garcia-Vergara, S.J.; Curioni, M.; Habazaki, H.; Skeldon, P.; Thompson, G.E. Effects of Current Density and Electrolyte Temperature on the Volume Expansion Factor of Anodic Alumina Formed in Oxalic Acid. Electrochim. Acta 2012, 59, 186–195. [Google Scholar] [CrossRef]
  195. Cui, Q.; Liu, T.; Li, X.; Song, K.; Ge, D. Nanopillared Polycarbonate Surfaces Having Variable Feature Parameters as Bactericidal Coatings. ACS Appl. Nano Mater. 2020, 3, 4599–4609. [Google Scholar] [CrossRef]
  196. Zhang, X.; Zhang, J.; Han, X.; Wang, S.; Hao, L.; Zhang, C.; Fan, Y.; Zhao, J.; Jiang, R.; Ren, L. A Photothermal Therapy Enhanced Mechano-Bactericidal Hybrid Nanostructured Surface. J. Colloid Interface Sci. 2023, 645, 380–390. [Google Scholar] [CrossRef]
  197. Chaliha, C.; Kalita, E. Nanosystems for Antimicrobial Interventions: Advanced Synthesis and Implementation Strategies. In Antimicrobial Nanosystems; Elsevier: Amsterdam, The Netherlands, 2023; pp. 3–22. [Google Scholar]
  198. Shahali, H.; Hasan, J.; Wang, H.; Tesfamichael, T.; Yan, C.; Yarlagadda, P.K.D.V. Evaluation of Particle Beam Lithography for Fabrication of Metallic Nano-Structures. Procedia Manuf. 2019, 30, 261–267. [Google Scholar] [CrossRef]
  199. Pala, N. Karabiyik, Electron Beam Lithography (EBL). In Encyclopedia of Nanotechnology; Springer: Dordrecht, The Netherlands, 2012; pp. 718–740. [Google Scholar]
  200. Prakash, S.; Yeom, J. Advanced Fabrication Methods and Techniques. In Nanofluidics and Microfluidics; Elsevier: Amsterdam, The Netherlands, 2014; pp. 87–170. [Google Scholar]
  201. Kallas, P.; Haugen, H.J.; Gadegaard, N.; Stormonth-Darling, J.; Hulander, M.; Andersson, M.; Valen, H. Adhesion of Escherichia Coli to Nanostructured Surfaces and the Role of Type 1 Fimbriae. Nanomaterials 2020, 10, 2247. [Google Scholar] [CrossRef]
  202. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials: Synthesis, Properties, Modifications, and Applications. ChemInform 2007, 38, 2891–2959. [Google Scholar] [CrossRef]
  203. Zhao, L.; Liu, T.; Li, X.; Cui, Q.; Wu, Q.; Wang, X.; Song, K.; Ge, D. Low-Temperature Hydrothermal Synthesis of Novel 3D Hybrid Nanostructures on Titanium Surface with Mechano-Bactericidal Performance. ACS Biomater. Sci. Eng. 2021, 7, 2268–2278. [Google Scholar] [CrossRef]
  204. 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]
  205. Jaggessar, A.; Senevirathne, S.W.M.A.I.; Velic, A.; Yarlagadda, P.K.D.V. Antibacterial Activity of 3D versus 2D TiO2 Nanostructured Surfaces to Investigate Curvature and Orientation Effects. Curr. Opin. Biomed. Eng. 2022, 23, 100404. [Google Scholar] [CrossRef]
  206. Yoo, K.; Lee, W.; Kang, K.; Kim, I.; Kang, D.; Oh, D.K.; Kim, M.C.; Choi, H.; Kim, K.; Kim, M.; et al. Low-Temperature Large-Area Fabrication of ZnO Nanowires on Flexible Plastic Substrates by Solution-Processible Metal-Seeded Hydrothermal Growth. Nano Converg. 2020, 7, 24. [Google Scholar] [CrossRef] [PubMed]
  207. Wu, S.; Zuber, F.; Maniura-Weber, K.; Brugger, J.; Ren, Q. Nanostructured Surface Topographies Have an Effect on Bactericidal Activity. J. Nanobiotechnology 2018, 16, 20. [Google Scholar] [CrossRef]
  208. Park, H.-H.; Sun, K.; Seong, M.; Kang, M.; Park, S.; Hong, S.; Jung, H.; Jang, J.; Kim, J.; Jeong, H.E. Lipid-Hydrogel-Nanostructure Hybrids as Robust Biofilm-Resistant Polymeric Materials. ACS Macro Lett. 2019, 8, 64–69. [Google Scholar] [CrossRef] [PubMed]
  209. Velic, A.; Jaggessar, A.; Tesfamichael, T.; Li, Z.; Yarlagadda, P.K.D.V. Effects of Nanopillar Size and Spacing on Mechanical Perturbation and Bactericidal Killing Efficiency. Nanomaterials 2021, 11, 2472. [Google Scholar] [CrossRef] [PubMed]
  210. Seng, P.; Bayle, S.; Alliez, A.; Romain, F.; Casanova, D.; Stein, A. The Microbial Epidemiology of Breast Implant Infections in a Regional Referral Centre for Plastic and Reconstructive Surgery in the South of France. Int. J. Infect. Dis. 2015, 35, 62–66. [Google Scholar] [CrossRef]
  211. Xie, Y.; Qu, X.; Li, J.; Li, D.; Wei, W.; Hui, D.; Zhang, Q.; Meng, F.; Yin, H.; Xu, X.; et al. Ultrafast Physical Bacterial Inactivation and Photocatalytic Self-Cleaning of ZnO Nanoarrays for Rapid and Sustainable Bactericidal Applications. Sci. Total Environ. 2020, 738, 139714. [Google Scholar] [CrossRef] [PubMed]
  212. Ishak, M.I.; Liu, X.; Jenkins, J.; Nobbs, A.H.; Su, B. Protruding Nanostructured Surfaces for Antimicrobial and Osteogenic Titanium Implants. Coatings 2020, 10, 756. [Google Scholar] [CrossRef]
  213. Fisher, L.E.; Yang, Y.; Yuen, M.-F.; Zhang, W.; Nobbs, A.H.; Su, B. Bactericidal Activity of Biomimetic Diamond Nanocone Surfaces. Biointerphases 2016, 11, 011014. [Google Scholar] [CrossRef]
  214. Velic, A.; Hasan, J.; Li, Z.; Yarlagadda, P.K.D.V. Mechanics of Bacterial Interaction and Death on Nanopatterned Surfaces. Biophys. J. 2021, 120, 217–231. [Google Scholar] [CrossRef]
  215. Valiei, A.; Lin, N.; McKay, G.; Nguyen, D.; Moraes, C.; Hill, R.J.; Tufenkji, N. Surface Wettability Is a Key Feature in the Mechano-Bactericidal Activity of Nanopillars. ACS Appl. Mater. Interfaces 2022, 14, 27564–27574. [Google Scholar] [CrossRef]
  216. Michalska, M.; Gambacorta, F.; Divan, R.; Aranson, I.S.; Sokolov, A.; Noirot, P.; Laible, P.D. Tuning Antimicrobial Properties of Biomimetic Nanopatterned Surfaces. Nanoscale 2018, 10, 6639–6650. [Google Scholar] [CrossRef]
  217. Nistor, C.L.; Mihaescu, C.I.; Bala, D.; Gifu, I.C.; Ninciuleanu, C.M.; Burlacu, S.G.; Petcu, C.; Vladu, M.-G.; Ghebaur, A.; Stroea, L.; et al. Novel Hydrophobic Nanostructured Antibacterial Coatings for Metallic Surface Protection. Coatings 2022, 12, 253. [Google Scholar] [CrossRef]
  218. Mandal, P.; Shishodia, A.; Ali, N.; Ghosh, S.; Arora, H.S.; Grewal, H.S.; Ghosh, S.K. Effect of Topography and Chemical Treatment on the Hydrophobicity and Antibacterial Activities of Micropatterned Aluminium Surfaces. Surf. Topogr. 2020, 8, 025017. [Google Scholar] [CrossRef]
  219. Jindai, K.; Nakade, K.; Masuda, K.; Sagawa, T.; Kojima, H.; Shimizu, T.; Shingubara, S.; Ito, T. Adhesion and Bactericidal Properties of Nanostructured Surfaces Dependent on Bacterial Motility. RSC Adv. 2020, 10, 5673–5680. [Google Scholar] [CrossRef]
  220. Hasan, J.; Jain, S.; Chatterjee, K. Nanoscale Topography on Black Titanium Imparts Multi-Biofunctional Properties for Orthopedic Applications. Sci. Rep. 2017, 7, 41118. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Driving factors to choose a polymer as implant material. Compared to other materials polymers provide some favourable characteristics for biomedical implant applications, which is essential in the biomedical field for successful implant in in vivo applications.
Figure 1. Driving factors to choose a polymer as implant material. Compared to other materials polymers provide some favourable characteristics for biomedical implant applications, which is essential in the biomedical field for successful implant in in vivo applications.
Nanomaterials 13 02799 g001
Figure 2. (a) Major nanofeature parameters; (bj) different shapes of nanofeatures, type of nanoarchitectures (uniform (bf,i,j) and random (g,h)), and aspect ratio.
Figure 2. (a) Major nanofeature parameters; (bj) different shapes of nanofeatures, type of nanoarchitectures (uniform (bf,i,j) and random (g,h)), and aspect ratio.
Nanomaterials 13 02799 g002
Figure 4. Deviation of nanofeature parameters such as (a) tip diameter, (b) base diameter, (c) height, (d) spacing, and (e) surface wettability among the different natural bactericidal nano-topographies listed in Table 2. The mean values of each parameter are plotted with standard deviation error bars. Graph abbreviations: cicada wing—CW; dragonfly wing—DNFW; gecko skin—GS; damselfly wing—DSFW; lotus leaf—LL.
Figure 4. Deviation of nanofeature parameters such as (a) tip diameter, (b) base diameter, (c) height, (d) spacing, and (e) surface wettability among the different natural bactericidal nano-topographies listed in Table 2. The mean values of each parameter are plotted with standard deviation error bars. Graph abbreviations: cicada wing—CW; dragonfly wing—DNFW; gecko skin—GS; damselfly wing—DSFW; lotus leaf—LL.
Nanomaterials 13 02799 g004
Figure 5. (a,b) SEM micrographs (scale: 1 µm) of P. aeruginosa bacteria on unetched polymer surfaces; (d,e) SEM micrographs (scale: 1 µm) of lysed P. aeruginosa bacteria on etched polymer surfaces using RIE; (c,f) live/dead confocal micrographs (60×) of unetched and etched polymer surfaces with P. aeruginosa bacteria; (g) % of dead P. aeruginosa cells on each unetched and etched polymer surface. ** indicates significance for p < 0.01. Adapted with permission from ref. [160]. Copyright© 2023, American Chemical Society.
Figure 5. (a,b) SEM micrographs (scale: 1 µm) of P. aeruginosa bacteria on unetched polymer surfaces; (d,e) SEM micrographs (scale: 1 µm) of lysed P. aeruginosa bacteria on etched polymer surfaces using RIE; (c,f) live/dead confocal micrographs (60×) of unetched and etched polymer surfaces with P. aeruginosa bacteria; (g) % of dead P. aeruginosa cells on each unetched and etched polymer surface. ** indicates significance for p < 0.01. Adapted with permission from ref. [160]. Copyright© 2023, American Chemical Society.
Nanomaterials 13 02799 g005
Figure 6. (a) Colloidal lithography process steps. Step 1: Spin-coated PS microspheres on the polymer substrate. Step 2: Halfway etched microspheres and polymer substrate. Step 3: Completely etched microspheres and developed nanocone structures on the substrate. (bi). SEM micrographs of PET nanocone structures fabricated using 500 nm (be) and 200 nm (fi) PS microspheres at different etching times. Images (bi) adapted with permission from ref. [161]. Compared to the maskless RIE process, colloidal lithography is capable of producing different nanoarchitectures via mimicking cicada (bg), dragonfly (h,i), and damselfly (h,i) wings.
Figure 6. (a) Colloidal lithography process steps. Step 1: Spin-coated PS microspheres on the polymer substrate. Step 2: Halfway etched microspheres and polymer substrate. Step 3: Completely etched microspheres and developed nanocone structures on the substrate. (bi). SEM micrographs of PET nanocone structures fabricated using 500 nm (be) and 200 nm (fi) PS microspheres at different etching times. Images (bi) adapted with permission from ref. [161]. Compared to the maskless RIE process, colloidal lithography is capable of producing different nanoarchitectures via mimicking cicada (bg), dragonfly (h,i), and damselfly (h,i) wings.
Nanomaterials 13 02799 g006
Figure 7. (a) TNIL and UVNIL process steps. Adapted with permission from ref. [175]. (b) Fabrication steps of rose petal mimetic structures on PVDF-HFP films. Adapted with permission from ref. [179]. Copyright© 2023, American Chemical Society. (c) SEM micrographs of nanopatterned thermal shrinkage films via UVNIL process. Adapted with permission from ref. [181]. (df) SEM micrographs (30° tilted, scale bar: 1 µm) of nanostructure made on PMMA surface with different spacings ((a): 600 nm, (b): 300 nm, (c): 200 nm). Adapted with permission from ref. [57].
Figure 7. (a) TNIL and UVNIL process steps. Adapted with permission from ref. [175]. (b) Fabrication steps of rose petal mimetic structures on PVDF-HFP films. Adapted with permission from ref. [179]. Copyright© 2023, American Chemical Society. (c) SEM micrographs of nanopatterned thermal shrinkage films via UVNIL process. Adapted with permission from ref. [181]. (df) SEM micrographs (30° tilted, scale bar: 1 µm) of nanostructure made on PMMA surface with different spacings ((a): 600 nm, (b): 300 nm, (c): 200 nm). Adapted with permission from ref. [57].
Nanomaterials 13 02799 g007
Figure 9. (ac) SEM micrographs (scale: 200 nm) of fabricated nanostructures via AAOT on PC surface: (a) 100 nm, (b) 170 nm, and (c) 300 nm of interpillar spacing. 30° tilted SEM images of nanopillars (a1a4,b1b4,c1c4) with different heights (H). Adapted with permission from ref. [195]. Copyright© 2020, American Chemical Society. (d) Nanostructure fabrication using AAOT. Adapted with permission from ref. [192].
Figure 9. (ac) SEM micrographs (scale: 200 nm) of fabricated nanostructures via AAOT on PC surface: (a) 100 nm, (b) 170 nm, and (c) 300 nm of interpillar spacing. 30° tilted SEM images of nanopillars (a1a4,b1b4,c1c4) with different heights (H). Adapted with permission from ref. [195]. Copyright© 2020, American Chemical Society. (d) Nanostructure fabrication using AAOT. Adapted with permission from ref. [192].
Nanomaterials 13 02799 g009
Figure 10. (a) Schematic drawing of ZnO nanowire formation on plastic substrate. (b,c) SEM micrographs of formation of nanoporous morphology of Ag ink coating on plastic substrate. (df) SEM micrographs of grown ZnO nanowires on glass (d), PI (e), and PC (f) substrates. Adapted from ref. [206].
Figure 10. (a) Schematic drawing of ZnO nanowire formation on plastic substrate. (b,c) SEM micrographs of formation of nanoporous morphology of Ag ink coating on plastic substrate. (df) SEM micrographs of grown ZnO nanowires on glass (d), PI (e), and PC (f) substrates. Adapted from ref. [206].
Nanomaterials 13 02799 g010
Figure 11. Comparison of literature results on bactericidal nanostructured surfaces demonstrating the correlation between tip radius and spacing with bactericidal efficacy. Adapted from ref. [209]. In most of the analysed studies, when the pillar experiences high von Mises stress and contact pressure, it leads to high bactericidal efficacy [21,57,101,150,154,159,161,198].
Figure 11. Comparison of literature results on bactericidal nanostructured surfaces demonstrating the correlation between tip radius and spacing with bactericidal efficacy. Adapted from ref. [209]. In most of the analysed studies, when the pillar experiences high von Mises stress and contact pressure, it leads to high bactericidal efficacy [21,57,101,150,154,159,161,198].
Nanomaterials 13 02799 g011
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kumara, S.P.S.N.B.S.; Senevirathne, S.W.M.A.I.; Mathew, A.; Bray, L.; Mirkhalaf, M.; Yarlagadda, P.K.D.V. Progress in Nanostructured Mechano-Bactericidal Polymeric Surfaces for Biomedical Applications. Nanomaterials 2023, 13, 2799. https://doi.org/10.3390/nano13202799

AMA Style

Kumara SPSNBS, Senevirathne SWMAI, Mathew A, Bray L, Mirkhalaf M, Yarlagadda PKDV. Progress in Nanostructured Mechano-Bactericidal Polymeric Surfaces for Biomedical Applications. Nanomaterials. 2023; 13(20):2799. https://doi.org/10.3390/nano13202799

Chicago/Turabian Style

Kumara, S. P. S. N. Buddhika Sampath, S. W. M. Amal Ishantha Senevirathne, Asha Mathew, Laura Bray, Mohammad Mirkhalaf, and Prasad K. D. V. Yarlagadda. 2023. "Progress in Nanostructured Mechano-Bactericidal Polymeric Surfaces for Biomedical Applications" Nanomaterials 13, no. 20: 2799. https://doi.org/10.3390/nano13202799

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