Next Article in Journal
Mutations in embB406 Are Associated with Low-Level Ethambutol Resistance in Canadian Mycobacterium tuberculosis Isolates
Previous Article in Journal
Antiparasitic Activity of Isolated Fractions from Parthenium incanum Kunth against the Hemoflagellate Protozoan Trypanosoma cruzi
Previous Article in Special Issue
Chestnut Honey Is Effective against Mixed Biofilms at Different Stages of Maturity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Medical Device-Associated Infections Caused by Biofilm-Forming Microbial Pathogens and Controlling Strategies

by
Akanksha Mishra
1,
Ashish Aggarwal
1,* and
Fazlurrahman Khan
2,3,4,5,*
1
School of Bioengineering and Biosciences, Lovely Professional University, Phagwara 144001, Punjab, India
2
Institute of Fisheries Science, Pukyong National University, Busan 48513, Republic of Korea
3
International Graduate Program of Fisheries Science, Pukyong National University, Busan 48513, Republic of Korea
4
Marine Integrated Biomedical Technology Center, The National Key Research Institutes in Universities, Pukyong National University, Busan 48513, Republic of Korea
5
Research Center for Marine Integrated Bionics Technology, Pukyong National University, Busan 48513, Republic of Korea
*
Authors to whom correspondence should be addressed.
Antibiotics 2024, 13(7), 623; https://doi.org/10.3390/antibiotics13070623
Submission received: 30 May 2024 / Revised: 28 June 2024 / Accepted: 2 July 2024 / Published: 4 July 2024

Abstract

:
Hospital-acquired infections, also known as nosocomial infections, include bloodstream infections, surgical site infections, skin and soft tissue infections, respiratory tract infections, and urinary tract infections. According to reports, Gram-positive and Gram-negative pathogenic bacteria account for up to 70% of nosocomial infections in intensive care unit (ICU) patients. Biofilm production is a main virulence mechanism and a distinguishing feature of bacterial pathogens. Most bacterial pathogens develop biofilms at the solid-liquid and air-liquid interfaces. An essential requirement for biofilm production is the presence of a conditioning film. A conditioning film provides the first surface on which bacteria can adhere and fosters the growth of biofilms by creating a favorable environment. The conditioning film improves microbial adherence by delivering chemical signals or generating microenvironments. Microorganisms use this coating as a nutrient source. The film gathers both inorganic and organic substances from its surroundings, or these substances are generated by microbes in the film. These nutrients boost the initial growth of the adhering bacteria and facilitate biofilm formation by acting as a food source. Coatings with combined antibacterial efficacy and antifouling properties provide further benefits by preventing dead cells and debris from adhering to the surfaces. In the present review, we address numerous pathogenic microbes that form biofilms on the surfaces of biomedical devices. In addition, we explore several efficient smart antiadhesive coatings on the surfaces of biomedical device-relevant materials that manage nosocomial infections caused by biofilm-forming microbial pathogens.

1. Introduction

There is a high incidence of nosocomial infections caused by contaminated medical equipment, such as urinary catheters, intravascular catheters, and orthopedic implants, which contain pathogenic bacteria [1]. These pathogens include methicillin-resistant Staphylococcus aureus, Escherichia coli, Acinetobacter baumannii, and Pseudomonas aeruginosa [2]. Healthcare-associated infections (HAIs) are a global occurrence that leads to significant increases in mortality and morbidity. These present serious issues in both underdeveloped and highly developed European countries [3]. Numerous studies have shown that the types of bacteria that cause HAIs vary depending on the type of medical implant the patient has [4]. Accurate determination of whether the patient contracted the bacteria before hospital admission or during the hospital stay is crucial. Only infections that manifest in patients 48 h after admission are classified as HAIs. Symptoms, including fever, chills, fatigue, coughing, dyspnea, stomach pain, and loose stools, indicate that the patient is infected. Inflammation and sepsis are typical symptoms [5]. Catheterization-related nosocomial infections are often linked to antibiotic-resistant microorganisms, such as Staphylococcus, Enterococcus, and different enterobacterial species, as well as fungi, such as Candida spp. [6]. Approximately 40,000 hospitalized patients die each year globally from HAIs, which account for over 25% of the ailments attained during the process of health care in developing countries and up to 15% in wealthy countries [7]. Infections with different microorganisms (including P. aeruginosa, Klebsiella pneumoniae, A. baumannii, and E. coli) can cause bloodstream infections, which are a primary cause of death, prolonged hospital and Intensive Care Unit (ICU) stays, and increased healthcare costs [8]. The majority of indwelling central venous catheters are colonized by microorganisms embedded in a biofilm matrix, as demonstrated by scanning electron microscopy [9].
Research has demonstrated that bacterial adhesion and biofilm formation can occur in various medical devices, including dental chair water lines, indwelling stents, urinary catheters, intrauterine devices, and contact lenses [10]. There are multiple stages in the establishment and growth of biofilms; single bacterial attachment (both reversible and irreversible), bacterial aggregation, microcolony formation, maturation, and dispersion/detachment are the five main stages of bacterial attachment [11] (Figure 1). Quorum sensing (QS), a key mechanism in bacterial communication, involves the production, release, detection, and response to extracellular signaling molecules [12,13]. QS signaling enables bacteria to collectively modify their behavior, including the production of virulence factors and biofilms, in response to changes in cell density and community composition [12,14]. Since the beginning of time, bacteria have been on Earth in two states. Sessile bacteria are said to be 500–5000 times more resistant to antibiotics than their planktonic counterparts [15]. The complex, multi-step, and usually cyclic process of biofilm formation involves several bacterial species. Bacterial biofilms secrete extracellular polymeric material, which is a mixture of polysaccharides, proteins (mostly composed of D-amino acids), fatty acids, and different nucleic acids [16]. According to the National Institutes of Health, biofilms can be responsible for up to 80% of human microbial infections. These infections include meningitis, kidney infections, endocarditis, cystic fibrosis, periodontitis, rhinosinusitis, osteomyelitis, non-healing chronic wounds, and infections related to prosthetic and implantable devices [17]. The microbiological factors that control the development of biofilms have been identified by in vitro analyses of biofilm infection [18].
Figure 1. Stages of biofilm formation and mechanism showing antibiotic resistance by pathogenic bacterial strain. Reprinted from the [19], Copyright © 2023 by the authors and Licensee MDPI, Basel, Switzerland.
Figure 1. Stages of biofilm formation and mechanism showing antibiotic resistance by pathogenic bacterial strain. Reprinted from the [19], Copyright © 2023 by the authors and Licensee MDPI, Basel, Switzerland.
Antibiotics 13 00623 g001
The development of biofilms on the surfaces of medical devices is a major contributing factor to the infections associated with these devices [20,21]. These biofilms provide a protective barrier for bacteria, rendering them resistant to antimicrobial treatments and increasing the risk of infection [19]. At least three intricate components are involved in microbial colonization: the device, microbes, and the host environment (such as tissues and immune cells). The surface properties of medical implants, such as their chemical composition and morphology, play a crucial role in biofilm formation and bacterial adhesion [22]. Microbial colonization is frequently difficult to identify. In certain situations, it may go unnoticed for years; however, in other situations, it may be urgent enough to endanger life [23]. Slimy “biofilm” coatings created by invading bacteria have been found on a number of gadgets that were discarded following problems due to microbial colonization [24]. Recalcitrance, a biofilm lifestyle trait that causes medication failure and infection recurrence, refers to the ability of pathogenic biofilms to thrive in the presence of high doses of antibiotics [25]. The word “recalcitrance” refers to a subgroup of biofilm-forming bacteria that can survive in the presence of high doses of antibiotics (Figure 1). The distinct adaptive antimicrobial resistance mechanism known as biofilm-forming bacteria is directly responsible for a number of therapeutic problems that arise in clinical settings and ultimately result in the death of patients [26]. It encompasses the idea of non-susceptibility to (antimicrobial) control of refractory biofilms [25]. Although bacteria have multiple antibiotic resistance mechanisms, such as target modification, efflux pump expression, inactivation of antibiotic substances, and target bypass [27] (Figure 2), the formation of biofilms exhibits an adaptive resistance mechanism against antibiotics, as well as bypassing the host defense systems [28]. The biofilm matrix also prevents antibiotics from reaching the cells, contributing to the overall resistance [29]. This structure also facilitates the flow of antimicrobial-resistant genes between and within the species, protects against drug penetration, and increases their persistence [30].
The formation of a biofilm when bacteria contaminate an indwelling medical device depends on several factors [31]. Before permanently connecting, the microbes must cling for a sufficient time to expose the tool surfaces. The type and number of cells influence the rate of cell attachment in the liquid to which the device is exposed, the liquid flow rate through the device, and the physicochemical characteristics of the surface [11]. It is possible for liquid components to alter the surface properties as well as the pace of adhesion. The flow rate, nutritional composition of the medium, concentration of the antimicrobial medicine, and ambient temperature affect the pace of growth of these cells once they irreversibly attach to and generate proteins found outside the cells to create a biofilm [32]. Biofilms formed on three different types of indwelling medical devices were used to illustrate these variables: central venous catheters, urine (Foley) catheters, and mechanical heart valves [33]. Different stages of biofilm adherence to the surface of medical devices have been reported previously [34]. To strengthen the knowledge of possible strategies to control biomedical device-associated infections, the present review aims to discuss (1) the clinical significance of medical device-associated infections caused by biofilm-forming bacterial pathogens and (2) possible treatment strategies for controlling biomedical devices associated with biofilms using different materials.

2. Role of Biofilms in HAIs

The most prevalent biofilm-based illness caused by medical equipment is catheter-associated urinary tract infection (CAUTI), which affects over 150 million individuals worldwide annually [35]. Flexible multichannel endoscopes are a unique type of reusable medical device. If reprocessing guidelines are not followed properly, biofilm growth may occur. Biofilm production thrives in damp, nutrient-rich conditions inside the lumen of an endoscope used on a patient [36]. A protein-containing coating is formed around an implant or other device when it enters the body, facilitating bacterial colonization. After bacteria attach to themselves, biofilms begin to form. When bacteria reach maturity, they begin to spread, and some enter the bloodstream, potentially leading to serious infections [37]. Following a thorough examination of the data regarding bacterial adherence and device surface change, the following five main principles are identified: (1) Different bacteria could attach to the same device material in various ways; (2) the same bacteria may attach to various device materials in different ways; (3) bacteria may attach differently to the same device material under different environmental conditions, such as the type of flow (stationary vs. dynamic), temperature, and the hydrophobic versus hydrophilic medium in which the device is placed; (4) the prevention of bacterial colonization of the device in vitro cannot ensure anti-infective effectiveness in vivo; and (5) depending on the application, different surface-modifying approaches may have different therapeutic benefit [38].

3. Pathogenesis of Biofilm-forming Microbes

Bacterial biofilms are prevalent in the human body and can have significant impacts on health and disease [39]. Biofilms have been shown to grow on the surface of medical equipment, and the distribution of both single and clustered cells implies a substantial risk of microbial dissemination within the host and an elevated risk of infection. Hospitals, assisted living facilities, and even patients’ homes can have bacteria that cause HAIs [40]. Perry and Tan [39] summarized the formation of bacterial biofilms in the human body at different locations, such as the upper respiratory tract, middle ear, soft tissue wounds, urinary tract, male and female reproductive tract, bone, oral cavity, cardiovascular system, stomach, and colon.
The human oral cavity initially creates an aerobic environment where oxygen is first consumed by facultative anaerobic bacteria (such as Actinomyces and Streptococcus spp.) or aerobic bacteria (such as Neisseria spp.), creating an environment that is suitable for the survival of obligate anaerobic bacteria. Anaerobic bacteria predominantly populate the human oral cavity during biofilm formation [41]. When wound healing is disrupted, exogenous infections usually occur during surgery or early postoperatively. Patients with large hematomas typically present with this condition. Exogenous infections rarely develop late in the healing process during arthrocentesis or after device-induced or spontaneous skin rupture [42]. Spontaneous skin rupture occurs more frequently after osteosynthesis than after joint replacement [43].
Biofilms have been found in the circulatory system of atherosclerotic arteries and in heart infections (endocarditis). When bacteria, most frequently S. aureus, Streptococcus species, and Enterococcus species, adhere to the heart valves or the inner lining of the heart chambers, they can cause infectious endocarditis (Table 1). This condition usually affects patients with congenital valve defects or damaged heart tissue [44]. Recent data imply that planktonic bacteria have long been associated with acute respiratory illnesses, where aggregate-type biofilms in sputum are commonly thought to be the predominant form, as well as chronic lung infections, as is widely acknowledged [45].
The incidence of infections linked to medical device biofilms has become a significant clinical issue because biofilms are resilient and resistant to antimicrobial therapy, and many researchers are now focusing on the mechanisms by which they form and thrive using standard antibacterial techniques [46]. Locally acquired host defense deficiency is the primary cause of implanted devices that increase sensitivity to infection, and the rapid development of a biofilm that is resistant to both host defense and antimicrobial treatments is the main cause of persistence [47]. Patients in the ICU have a heightened vulnerability to device-associated nosocomial infections owing to their compromised immune systems and frequent contact with invasive medical equipment. The use of these devices increases the incidence of ventilator-associated pneumonia (VAP), central line-associated bloodstream infections (CLABSIs), and CAUTIs in patients admitted to the ICU [48].
Table 1. Various medical devices and common pathogens form biofilm.
Table 1. Various medical devices and common pathogens form biofilm.
Medical DeviceCommon PathogenReferences
Orthopedic devicesStaphylococci, Gram-negative bacilli[49]
Endotracheal tubesPseudomonas aeruginosa, Staphylococcus aureus[50]
Contact lensesPseudomonas aeruginosa, Staphylococci[51]
Intravascular cathetersStaphylococci, Enterococci, Gram-negative bacilli[52]
Valves, pacemakerStaphylococci, Streptococci[53]
Respiratory equipment, Indwelling cathetersAcinetobacter baumanii[54]
BronchoscopyKlebsiella pneumoniae[55]
UTIs devices,
Intravascular medical device
Enterobacter spp.[56]
Respiratory deviceAspergillus fumigatus[57]
Cardiac medical deviceCryptococcus neoformans[58]
Urinary cathetersEnterococcus faecalis[59]
Catheters, prosthetic devicesAspergillus fumigatus[60]
Central venous catheterStaphylococcus epidermidis[31]
Cerebrospinal shuntsStaphylococcus aureus, Propionibacterium[61]
Dental implantsPrevotella intermedia, Actinobacillus[62]

4. Establishment of the Biofilm on Biomedical Device Surfaces

The challenges in employing available antibiotics to treat biofilm-associated illnesses, especially implant-associated infections, are exacerbated by bacterial biofilm tolerance and resistance. In 36 countries in Latin America, Asia, Africa, and Europe, 422 ICUs were examined by the International Nosocomial Infection Control Consortium. The results indicated 7029, 6595, and 12,145 cases of CLABSIs, CAUTIs, and VAPs, respectively, between January 2004 and December 2009. Implant-associated infections have gained attention because of the rapid development of implantable biomedical devices [63]. Significant increases in economic losses, morbidity, and mortality are linked to HAIs, many of which are unavoidable. The most frequent hospital-acquired infections are surgical site-and device-associated HAIs, including VAP, CLABSIs, and CAUTIs [64].
Furthermore, infections are common and frequently associated with surfaces and devices that have biofilms established on them [65]. Physiological gradients, matrix diffusion restrictions, and innate and evolved resistance mechanisms make it challenging to control biofilms formed on medical devices using antimicrobials. Together, these factors promote antimicrobial resistance [66]. Cell adhesion, which is required for biofilm development and other biological processes, is an essential component of interactions between cells and surfaces. Surface wettability, rather than polymer type or surface topography, is a key factor that influences cell attachment and proliferation [67]. Biofilm formation initiates the disease process through various mechanisms, including the detachment of individual bacterial cells or clusters of cell aggregates, the production of endotoxins, increased evasion from host immune system surveillance, and the creation of a protective barrier that fosters the emergence of immune-resistant organisms (Figure 3). When pathogens, such as P. aeruginosa, attach themselves in real time to biomedical equipment in the clinic, biofilm formation can be monitored. This allows timely antibiotic treatment or device removal when the first signs of bacterial attachment are observed [68]. By doing this, the maturation of biofilms can be stopped, and chronic infections and related difficulties can be lessened [69]. Early action can stop biofilms from growing until they reach the dispersion stage of their life cycle, leading to systemic infections. This strategy would reduce healthcare costs for each patient while improving their health. Currently, symptoms are used to detect device-related infections. However, to examine the developed biofilm, the device must be disassembled for microscopic and microbiological examinations [70].
This laborious procedure postpones appropriate corrective measures to prevent the worsening of the infection [71]. When a biofilm forms on a medical device, removing the bacteria can be difficult and expensive because extensive hospital stays, surgeries, and long-term antibiotic treatments are frequently required [72]. It is widely acknowledged that biofilm formation is one of the key virulence factors in infections linked to medical devices [21].
Figure 3. Stages of biofilm formation on the surface of the medical devices. Reprinted from the [73], Copyright © 2021 by the authors and Licensee MDPI, Basel, Switzerland.
Figure 3. Stages of biofilm formation on the surface of the medical devices. Reprinted from the [73], Copyright © 2021 by the authors and Licensee MDPI, Basel, Switzerland.
Antibiotics 13 00623 g003

5. Recent Advancements in Antiadhesive or Antifouling Coating on the Surface of Biomedical Devices

The development of a promising antibacterial coating on the surface of biomaterials and biomedical devices is facilitated by the integration of bactericidal and antifouling coatings, as both functions work synergistically to provide benefits [74]. The most popular antimicrobial strategy involves the addition of bactericidal materials or reagents to a surface and carrying out antimicrobial activities through release or contact killing, which are referred to as releasable and non-releasable bactericidal compounds, respectively [75]. To effectively prevent microbial biofilms and biological biofouling infections caused by biomaterials and biomedical devices, they must demonstrate not only superior growth-inhibitory efficacy against pathogenic bacterial species but also the capacity to prevent the adhesion of live or dead microbiological species and nonspecific platelets, proteins, and other biological macromolecules [76]. By reducing the adhesion force between a solid surface and bacteria, antiadhesive/antifouling surfaces enable bacteria to be readily removed before a biofilm develops [77]. Barnacle cement has been successfully used as a surface anchor to attach antifouling and antimicrobial polymer brushes to stainless steel [78]. This approach has been shown to be stable and effective, reducing protein adsorption and bacterial adhesion [79]. Other studies have also explored the use of biomimetic anchors, such as polydopamine layers, for the attachment of functional polymer brushes to stainless steel, resulting in enhanced antifouling and anticorrosion properties [80]. A one-step anchoring method using tannic acid-scaffolded bifunctional coatings has also been developed, further improving resistance to protein adsorption and bacterial adhesion [81]. Recent studies have made significant strides in the development of antibacterial coatings to combat orthopedic implant-associated infections (Figure 4). Wang et al. [82] and Akay and Yaghmur [66] highlighted the potential of antibacterial hydrogel coatings and modified implant surfaces to prevent biofilm formation. These coatings are designed to inhibit bacterial attachment and colonization, thereby reducing the risk of infection. Wang et al. [83] discussed the use of biodegradable alloy materials with inherent antibacterial properties as orthopedic implant materials, thereby providing a promising alternative to traditional implants. However, the need for further research and clinical testing of these coatings was emphasized by Wang et al. [83], underscoring the importance of continued advancements in this field.
Several studies have demonstrated the application of antibacterial coatings consisting of nanoparticles and lipids to minimize implant-related infections [84,85,86]. Nanoparticles and lipid coatings play crucial roles in enhancing the performance of biomedical devices. Both Luchini et al. [87] and Simović et al. [86] highlight the potential of lipid-coated inorganic nanoparticles in improving the stability, performance, and biocompatibility of lipid-based colloids, as well as in drug delivery systems. Jiménez-Jiménez et al. [88] further emphasized the versatility of this technology, particularly in the use of cell membranes to coat nanoparticles, which can improve their performance in various applications. Mashaghi et al. [89] underscore the significance of lipid nanotechnology in these advancements, particularly in the fields of targeted drug delivery and bioimaging. These studies collectively demonstrate the potential of nanoparticles and lipid coatings to enhance the functionality and effectiveness of biomedical devices (Table 2). Numerous types of metal nanoparticles have proven to be effective antibacterial agents [90,91,92,93]. The most frequently used antibacterial nano-agents are oxide-based nanoparticles of silver, gold, copper, titanium, nickel, magnesium, and zinc [94].
Figure 4. Representation of biofilm formation control and its removal by various methods. Reprinted from the [95], Copyright © 2023 by the authors and Licensee, Elsevier Ltd. (Amsterdam, The Netherlands).
Figure 4. Representation of biofilm formation control and its removal by various methods. Reprinted from the [95], Copyright © 2023 by the authors and Licensee, Elsevier Ltd. (Amsterdam, The Netherlands).
Antibiotics 13 00623 g004
Table 2. Different types of antiadhesive or antibiofilm coating agents are employed on the biomedical device surface.
Table 2. Different types of antiadhesive or antibiofilm coating agents are employed on the biomedical device surface.
Name of Biomedical DevicesAntibacterial MaterialsAntiadhesive or Antibiofilm Coating AgentsTypes of PathogensEfficacy (Killing of Attached Cells in %)References
Polydimethylsiloxane (PDMS), stainless steel (SS) surfaceAntibacterial surfaces that are smart and photothermally activatedTannic acid(TA) and Fe3+ ionEscherichia coli>99[96]
Blood-contacting medical devicesmetal–phenolic and catecholamineCuII–GA/CySA coatings
copper ions (CuII)-gallic acid (GA)/ cystamine (CySA)
Escherichia coli, Staphylococcus aureus∼99[97]
Surfaces made of a variety of materials, including silicone, glass, poly(methyl methacrylate) (PMMA) plates, metal surfaces, polypropylene fiber, and filter paperElectrostatic attraction, physically rupturing cell wallsCoating with positively-charged Zeolitic imidazolate frameworks(ZIF) nano-dagger arraysStaphylococcus aureusNA[98]
Silicon (Si), PDMS, and SSGold nanoparticle layer(GNPL)Regenerable smart antibacterial surfacesEscherichia coli>99[99]
Implantable deviceHydroxyethyl methacrylate(HEMA) and quercetin(Qe)Dual-functional anti-biofilms surfacePseudomonas aeruginosa, Staphylococcus aureusNA[100]
Metal materials surfacesPoly (2-hydroxyethyl methacrylate) hydrophilic polymer with Quaternary ammonium salt(QAS)Intelligent composite material surface with titanium contentStaphylococcus aureus, Escherichia coli99.86 and 97.08[101]
Biomedical cathetersSulfamethoxazole (SMZ) and
trimethoprim (TMP)
Polyethylene glycol(PEG)Staphylococcus aureus,
Escherichia coli
80 (E. coli and S. aureus cells)[102]
Urinary catheterPoly(styrene sulfonate) (PSS),
quaternary ammonium, H2O2 enzyme
New zwitterionic copolymers (PTMAEMA-co-PSPE) with different proportions of sulfobetaineStaphylococcus aureus60[103]
Foley cathetersα-aminoisobutyric acidBiocompatible amino acidsEscherichia coli, Bacillus subtilisNA[104]
Catheters, stents, and
dialysis
equipment
Silver nanoparticles2-Methacryloyloxyethyl phosphorylcholineEscherichia coli, Escherichia coli K-12>99[105]
Dental implantQuaternized polyethyleneiminepoly(glycidyl methacrylate)Staphylococcus aureus95.6[106]
Implant medical devicesGold nanoparticlesbuilt a mixed-metal-organic network nanoclusterEscherichia coliNA[107]
Soft contact lensPolyelectrolytesDiclofenac sodium salt, doxifloxacin hydrochloride, and chlorhexidine diacetate monohydratePseudomonas aeruginosa, Staphylococcus aureusNA[108]
Dental implantsZinc (Zn) nanoparticles
Electrohydrodynamic deposition
Nanoparticles of hydroxyapatite (nHA) and zinc oxide (nZnO)Streptococcus spp.NA[109]
Sinus StentsCiprofloxacin nanoparticlePoly-l-lactic acid (PLLA), ciprofloxacinPseudomonas aeruginosaNA[110]
Bone implant devicesCombined vancomycin and MelittinChitosan, bioactive
glass, and melittin
Staphylococcus aureusNA[111]
Orthopedic implantsLiposome-encapsulated photosensitizers (PS), IR780, and perfluorohexane (PFH),Lecithin, cholesterol
and PEGylated DSPE
Escherichia coli, Staphylococcus aureus99.62 and 99.63[112]
Subcutaneous implantsNitric oxide (NO)NO-releasing xerogel coatings of silicone rubberStaphylococcus aureus82[113]
Biomedical ImplantsBlack phosphorusBlack phosphorus nanosheets with N,N′-
dimethyl propylene urea (DMPU)
Bacillus subtilis99.69[114]
Implantable medical devicesTitanium, titanium binding peptides (TiBP)Chimeric peptide
(TiBP(S)1–3 and
E14LKK/H14LKK
motifs.)
Streptococcus mutans, Staphylococcus
epidermidis, Escherichia coli
NA[115]
Biomedical implantsNanoparticles of (silver and
zinc pyrithione, ZnP)
Silver-based organomodified layered silicate additives nanoclaysStaphylococcus aureusNA[116]
NA, Not available.

6. Conclusions and Future Perspectives

Antimicrobial coatings are widely used; however, the standards for these coatings are particularly strict in biomedical applications, which is the subject of this study. The most common methods involve the prevention of bacterial adherence and killing microbes through surface-associated mechanisms or coatings that emit antibacterial chemicals. To combat the growing resistance to conventional antibiotics, metal and metal oxide nanoparticles, as well as 2D nanomaterials, have provided innovative substitutes for antibiotic treatments of hospital-acquired illnesses linked to biofilms. To prevent medical implants linked to infections, antimicrobial-releasing coatings have undergone the most research. The search for innovative anti-infective biomaterials might raise reasonable hope for averting infection problems, which are connected to both long-term medical implant use and surgical treatment. Future developments may involve customized coatings depending on the requirements of specific patients or the microbiological environment, given variations in patient reactions to biomaterials. Customizing coatings to target particular bacteria known to cause issues in a given patient may improve the effectiveness of antibacterial strategies. Future research perspectives are recommended to detect biofilms on biomedical device surfaces and to prevent biofilm formation on the surface.
  • Early biofilm detection will undoubtedly aid patient treatment and reduce costs. However, this is only possible if detection procedures and techniques are continuously improved, which might be accomplished with the use of artificial intelligence tools [117].
  • Mixed-species biofilms exhibit notable differences in growth rate, gene expression, living habits, and structural appearance compared with those of single species. These differences are primarily manifested in enhanced biofilm metabolic capacity, resilience to environmental stress, and community-level signaling. Further studies on mixed-species biofilms are required [118].
  • Mixed-species biofilms predominate in nature and are common in human hosts, such as the lungs and oral cavities of individuals with cystic fibrosis. Therefore, further studies are required to define the interactions within multispecies biofilms and the consequences of these interactions on biofilm community growth, makeup, and longevity [119].
  • An antibiofilm or antiadhesive coating could be developed to prevent biofilm formation that works upon three lines of defense with antiadhesive, bactericidal, and anti-quorum sensing properties, adapting to the bacterial biofilm formation mechanism. Further research is required on such antibiofilm coatings [120].

Author Contributions

A.M.: methodology, investigation, data curation, writing, and editing; A.A.: supervision, review, and editing; F.K.: conceptualization, funding, supervision, writing, and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Basic Science Research Program through the National Research Foundation (NRF) of Korea grant funded by the Ministry of Education (RS-2023-00241461). This research was also supported by the Pukyong National University Research Fund in 2023 (202315350001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors, Akanksha Mishra and Ashish Aggarwal, are grateful to the School of Bioengineering and Biosciences, Lovely Professional University, for financial assistance. This work is part of the thesis of Akanksha Mishra, which was approved by the School of Bioengineering and Biosciences, Lovely Professional University, in India.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Solis-Velazquez, O.A.; Gutiérrez-Lomelí, M.; Guerreo-Medina, P.J.; Rosas-García, M.D.L.; Iñiguez-Moreno, M.; Avila-Novoa, M.G. Nosocomial pathogen biofilms on biomaterials: Different growth medium conditions and components of biofilms produced in vitro. J. Microbiol. Immunol. Infect. 2021, 54, 1038–1047. [Google Scholar] [CrossRef] [PubMed]
  2. Devanga Ragupathi, N.K.; Veeraraghavan, B.; Karunakaran, E.; Monk, P.N. Editorial: Biofilm-mediated nosocomial infections and its association with antimicrobial resistance: Detection, prevention, and management. Front. Med. 2022, 9, 987011. [Google Scholar] [CrossRef] [PubMed]
  3. Walter, J.; Haller, S.; Quinten, C.; Kärki, T.; Zacher, B.; Eckmanns, T.; Abu Sin, M.; Plachouras, D.; Kinross, P.; Suetens, C. Healthcare-associated pneumonia in acute care hospitals in European Union/European Economic Area countries: An analysis of data from a point prevalence survey, 2011 to 2012. Eurosurveillance 2018, 23, 1700843. [Google Scholar] [CrossRef] [PubMed]
  4. Dadi, N.C.T.; Radochová, B.; Vargová, J.; Bujdáková, H. Impact of Healthcare-Associated Infections Connected to Medical Devices-An Update. Microorganisms 2021, 9, 2332. [Google Scholar] [CrossRef] [PubMed]
  5. Boev, C.; Kiss, E. Hospital-Acquired Infections. Crit. Care Nurs. Clin. N. Am. 2017, 29, 51–65. [Google Scholar] [CrossRef] [PubMed]
  6. Liao, W.-C.; Chung, W.-S.; Lo, Y.-C.; Shih, W.-H.; Chou, C.-H.; Chen, C.-Y.; Tu, C.-Y.; Ho, M.-W. Changing epidemiology and prognosis of nosocomial bloodstream infection: A single-center retrospective study in Taiwan. J. Microbiol. Immunol. Infect. 2022, 55, 1293–1300. [Google Scholar] [CrossRef]
  7. Lemiech-Mirowska, E.; Kiersnowska, Z.; Michałkiewicz, M.; Depta, A.; Marczak, M. Nosocomial infections as one of the most important problems of healthcare system. Ann. Agric. Environ. Med. 2021, 28, 361–366. [Google Scholar] [CrossRef] [PubMed]
  8. Di Franco, S.; Alfieri, A.; Pace, M.C.; Sansone, P.; Pota, V.; Fittipaldi, C.; Fiore, M.; Passavanti, M.B. Blood Stream Infections from MDR Bacteria. Life 2021, 11, 575. [Google Scholar] [CrossRef]
  9. Elliott, T.S.; Moss, H.A.; Tebbs, S.E.; Wilson, I.C.; Bonser, R.S.; Graham, T.R.; Burke, L.P.; Faroqui, M.H. Novel approach to investigate a source of microbial contamination of central venous catheters. Eur. J. Clin. Microbiol. Infect. Dis. 1997, 16, 210–213. [Google Scholar] [CrossRef]
  10. Pandey, V.K.; Srivastava, K.R.; Ajmal, G.; Thakur, V.K.; Gupta, V.K.; Upadhyay, S.N.; Mishra, P.K. Differential Susceptibility of Catheter Biomaterials to Biofilm-Associated Infections and Their Remedy by Drug-Encapsulated Eudragit RL100 Nanoparticles. Int. J. Mol. Sci. 2019, 20, 5110. [Google Scholar] [CrossRef]
  11. Sharma, S.; Mohler, J.; Mahajan, S.D.; Schwartz, S.A.; Bruggemann, L.; Aalinkeel, R. Microbial Biofilm: A Review on Formation, Infection, Antibiotic Resistance, Control Measures, and Innovative Treatment. Microorganisms 2023, 11, 1614. [Google Scholar] [CrossRef]
  12. Papenfort, K.; Bassler, B.L. Quorum sensing signal-response systems in Gram-negative bacteria. Nat. Rev. Microbiol. 2016, 14, 576–588. [Google Scholar] [CrossRef]
  13. Khan, F.; Javaid, A.; Kim, Y.M. Functional Diversity of Quorum Sensing Receptors in Pathogenic Bacteria: Interspecies, Intraspecies and Interkingdom Level. Curr. Drug Targets 2019, 20, 655–667. [Google Scholar] [CrossRef]
  14. Zaitseva, Y.V.; Koksharova, O.A.; Lipasova, V.A.; Plyuta, V.A.; Demidyuk, I.V.; Chernin, L.S.; Khmel, I.A. SprI/SprR Quorum Sensing System of Serratia proteamaculans 94. Biomed. Res. Int. 2019, 2019, 3865780. [Google Scholar] [CrossRef]
  15. Oppenheimer-Shaanan, Y.; Steinberg, N.; Kolodkin-Gal, I. Small molecules are natural triggers for the disassembly of biofilms. Trends Microbiol. 2013, 21, 594–601. [Google Scholar] [CrossRef]
  16. Gupta, P.; Sarkar, S.; Das, B.; Bhattacharjee, S.; Tribedi, P. Biofilm, pathogenesis and prevention—A journey to break the wall: A review. Arch. Microbiol. 2016, 198, 1–15. [Google Scholar] [CrossRef]
  17. Bjarnsholt, T. The role of bacterial biofilms in chronic infections. APMIS 2013, 121, 1–58. [Google Scholar] [CrossRef]
  18. Joo, H.-S.; Otto, M. Molecular Basis of In Vivo Biofilm Formation by Bacterial Pathogens. Chem. Biol. 2012, 19, 1503–1513. [Google Scholar] [CrossRef]
  19. Li, P.; Yin, R.; Cheng, J.; Lin, J. Bacterial Biofilm Formation on Biomaterials and Approaches to Its Treatment and Prevention. Int. J. Mol. Sci. 2023, 24, 11680. [Google Scholar] [CrossRef]
  20. Zheng, Y.; He, L.; Asiamah, T.K.; Otto, M. Colonization of medical devices by staphylococci. Environ. Microbiol. 2018, 20, 3141–3153. [Google Scholar] [CrossRef]
  21. Mack, D.; Rohde, H.; Harris, L.G.; Davies, A.P.; Horstkotte, M.A.; Knobloch, J.K. Biofilm formation in medical device-related infection. Int. J. Artif. Organs 2006, 29, 343–359. [Google Scholar] [CrossRef] [PubMed]
  22. Veerachamy, S.; Yarlagadda, T.; Manivasagam, G.; Yarlagadda, P.K. Bacterial adherence and biofilm formation on medical implants: A review. Proc. Inst. Mech. Eng. Part H J. Eng. Med. 2014, 228, 1083–1099. [Google Scholar] [CrossRef] [PubMed]
  23. Vertes, A.; Hitchins, V.; Phillips, K.S. Analytical Challenges of Microbial Biofilms on Medical Devices. Anal. Chem. 2012, 84, 3858–3866. [Google Scholar] [CrossRef]
  24. Wu, H.; Moser, C.; Wang, H.-Z.; Høiby, N.; Song, Z.-J. Strategies for combating bacterial biofilm infections. Int. J. Oral Sci. 2015, 7, 1–7. [Google Scholar] [CrossRef] [PubMed]
  25. Lebeaux, D.; Ghigo, J.-M.; Beloin, C. Biofilm-Related Infections: Bridging the Gap between Clinical Management and Fundamental Aspects of Recalcitrance toward Antibiotics. Microbiol. Mol. Biol. Rev. 2014, 78, 510–543. [Google Scholar] [CrossRef] [PubMed]
  26. Khan, J.; Tarar, S.M.; Gul, I.; Nawaz, U.; Arshad, M. Challenges of antibiotic resistance biofilms and potential combating strategies: A review. 3 Biotech 2021, 11, 169. [Google Scholar] [CrossRef] [PubMed]
  27. Uluseker, C.; Kaster, K.M.; Thorsen, K.; Basiry, D.; Shobana, S.; Jain, M.; Kumar, G.; Kommedal, R.; Pala-Ozkok, I. A Review on Occurrence and Spread of Antibiotic Resistance in Wastewaters and in Wastewater Treatment Plants: Mechanisms and Perspectives. Front. Microbiol. 2021, 12, 717809. [Google Scholar] [CrossRef]
  28. Anderson, G.G.; O’Toole, G.A. Innate and induced resistance mechanisms of bacterial biofilms. Curr. Top. Microbiol. Immunol. 2008, 322, 85–105. [Google Scholar] [CrossRef] [PubMed]
  29. Mah, T.F. Biofilm-specific antibiotic resistance. Future Microbiol. 2012, 7, 1061–1072. [Google Scholar] [CrossRef]
  30. Coenye, T.; Bové, M.; Bjarnsholt, T. Biofilm antimicrobial susceptibility through an experimental evolutionary lens. Npj Biofilms Microbiomes 2022, 8, 82. [Google Scholar] [CrossRef]
  31. 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] [PubMed]
  32. Zhao, A.; Sun, J.; Liu, Y. Understanding bacterial biofilms: From definition to treatment strategies. Front. Cell. Infect. Microbiol. 2023, 13, 1137947. [Google Scholar] [CrossRef] [PubMed]
  33. Donlan, R. Biofilms and Device-Associated Infections. Emerg. Infect. Dis. 2001, 7, 277–281. [Google Scholar] [CrossRef] [PubMed]
  34. Carniello, V.; Peterson, B.W.; van der Mei, H.C.; Busscher, H.J. Physico-chemistry from initial bacterial adhesion to surface-programmed biofilm growth. Adv. Colloid Interface Sci. 2018, 261, 1–14. [Google Scholar] [CrossRef] [PubMed]
  35. Werneburg, G.T. Catheter-Associated Urinary Tract Infections: Current Challenges and Future Prospects. Res. Rep. Urol. 2022, 14, 109–133. [Google Scholar] [CrossRef]
  36. Roberts, C.G. The role of biofilms in reprocessing medical devices. Am. J. Infect. Control 2013, 41, S77–S80. [Google Scholar] [CrossRef]
  37. Di Domenico, E.G.; Oliva, A.; Guembe, M. The Current Knowledge on the Pathogenesis of Tissue and Medical Device-Related Biofilm Infections. Microorganisms 2022, 10, 1259. [Google Scholar] [CrossRef]
  38. Weinstein, R.A.; Darouiche, R.O. Device-Associated Infections: A Macroproblem That Starts with Microadherence. Clin. Infect. Dis. 2001, 33, 1567–1572. [Google Scholar] [CrossRef]
  39. Perry, E.K.; Tan, M.W. Bacterial biofilms in the human body: Prevalence and impacts on health and disease. Front. Cell Infect Microbiol. 2023, 13, 1237164. [Google Scholar] [CrossRef]
  40. Dewasthale, S.; Mani, I.; Vasdev, K. Microbial biofilm: Current challenges in health care industry. J. Appl. Biotechnol. Bioeng. 2018, 5, 160–164. [Google Scholar] [CrossRef]
  41. Mirzaei, R.; Mohammadzadeh, R.; Alikhani, M.Y.; Shokri Moghadam, M.; Karampoor, S.; Kazemi, S.; Barfipoursalar, A.; Yousefimashouf, R. The biofilm-associated bacterial infections unrelated to indwelling devices. IUBMB Life 2020, 72, 1271–1285. [Google Scholar] [CrossRef] [PubMed]
  42. Lu, S.; Yuan, Z.; He, X.; Du, Z.; Wang, Y. The impact of negative pressure wound therapy on surgical wound infection, hospital stay and postoperative complications after spinal surgery: A meta-analysis. Int. Wound J. 2024, 21, e14378. [Google Scholar] [CrossRef]
  43. Zimmerli, W.; Moser, C. Pathogenesis and treatment concepts of orthopaedic biofilm infections. FEMS Immunol. Med. Microbiol. 2012, 65, 158–168. [Google Scholar] [CrossRef]
  44. Lerche, C.J.; Schwartz, F.; Theut, M.; Fosbøl, E.L.; Iversen, K.; Bundgaard, H.; Høiby, N.; Moser, C. Anti-biofilm Approach in Infective Endocarditis Exposes New Treatment Strategies for Improved Outcome. Front. Cell Dev. Biol. 2021, 9, 643335. [Google Scholar] [CrossRef] [PubMed]
  45. Kolpen, M.; Kragh, K.N.; Enciso, J.B.; Faurholt-Jepsen, D.; Lindegaard, B.; Egelund, G.B.; Jensen, A.V.; Ravn, P.; Mathiesen, I.H.M.; Gheorge, A.G.; et al. Bacterial biofilms predominate in both acute and chronic human lung infections. Thorax 2022, 77, 1015–1022. [Google Scholar] [CrossRef]
  46. Khan, F.; Tabassum, N.; Kim, Y.-M. A strategy to control colonization of pathogens: Embedding of lactic acid bacteria on the surface of urinary catheter. Appl. Microbiol. Biotechnol. 2020, 104, 9053–9066. [Google Scholar] [CrossRef] [PubMed]
  47. Staats, A.; Li, D.; Sullivan, A.C.; Stoodley, P. Biofilm formation in periprosthetic joint infections. Ann. Jt. 2021, 6, 43. [Google Scholar] [CrossRef]
  48. Zorgani, A.; Abofayed, A.; Glia, A.; Albarbar, A.; Hanish, S. Prevalence of Device-associated Nosocomial Infections Caused by Gram-negative Bacteria in a Trauma Intensive Care Unit in Libya. Oman Med. J. 2015, 30, 270–275. [Google Scholar] [CrossRef] [PubMed]
  49. Beam, E.; Osmon, D. Prosthetic Joint Infection Update. Infect. Dis. Clin. N. Am. 2018, 32, 843–859. [Google Scholar] [CrossRef]
  50. Danin, P.-E.; Girou, E.; Legrand, P.; Louis, B.; Fodil, R.; Christov, C.; Devaquet, J.; Isabey, D.; Brochard, L. Description and Microbiology of Endotracheal Tube Biofilm in Mechanically Ventilated Subjects. Respir. Care 2015, 60, 21–29. [Google Scholar] [CrossRef]
  51. Urwin, L.; Okurowska, K.; Crowther, G.; Roy, S.; Garg, P.; Karunakaran, E.; MacNeil, S.; Partridge, L.J.; Green, L.R.; Monk, P.N. Corneal Infection Models: Tools to Investigate the Role of Biofilms in Bacterial Keratitis. Cells 2020, 9, 2450. [Google Scholar] [CrossRef] [PubMed]
  52. Selby, L.M.; Rupp, M.E.; Cawcutt, K.A. Prevention of Central-Line Associated Bloodstream Infections. Infect. Dis. Clin. N. Am. 2021, 35, 841–856. [Google Scholar] [CrossRef] [PubMed]
  53. Klug, D.; Balde, M.; Pavin, D.; Hidden-Lucet, F.o.; Clementy, J.; Sadoul, N.; Rey, J.L.; Lande, G.; Lazarus, A.; Victor, J.; et al. Risk Factors Related to Infections of Implanted Pacemakers and Cardioverter-Defibrillators. Circulation 2007, 116, 1349–1355. [Google Scholar] [CrossRef] [PubMed]
  54. Rebic, V.; Masic, N.; Teskeredzic, S.; Aljicevic, M.; Abduzaimovic, A.; Rebic, D. The Importance of Acinetobacter Species in the Hospital Environment. Med. Arch. 2018, 72, 330. [Google Scholar] [CrossRef] [PubMed]
  55. Galdys, A.L.; Marsh, J.W.; Delgado, E.; Pasculle, A.W.; Pacey, M.; Ayres, A.M.; Metzger, A.; Harrison, L.H.; Muto, C.A. Bronchoscope-associated clusters of multidrug-resistant Pseudomonas aeruginosa and carbapenem-resistant Klebsiella pneumoniae. Infect. Control. Hosp. Epidemiol. 2019, 40, 40–46. [Google Scholar] [CrossRef] [PubMed]
  56. Zou, J.; Peng, B.; Qu, J.; Zheng, J. Are Bacterial Persisters Dormant Cells Only? Front. Microbiol. 2022, 12, 708580. [Google Scholar] [CrossRef] [PubMed]
  57. McGuire, C.N.; Walter, D.J. Cryptococcus neoformans endocarditis in an immunocompetentpatient a case report. BMC Cardiovasc. Disord. 2022, 22, 565. [Google Scholar] [CrossRef]
  58. Garvey, M. Medical Device-Associated Healthcare Infections: Sterilization and the Potential of Novel Biological Approaches to Ensure Patient Safety. Int. J. Mol. Sci. 2023, 25, 201. [Google Scholar] [CrossRef] [PubMed]
  59. Josephs-Spaulding, J.; Singh, O.V. Medical Device Sterilization and Reprocessing in the Era of Multidrug-Resistant (MDR) Bacteria: Issues and Regulatory Concepts. Front. Med. Technol. 2021, 2, 587352. [Google Scholar] [CrossRef] [PubMed]
  60. Giles, C.; Lamont-Friedrich, S.J.; Michl, T.D.; Griesser, H.J.; Coad, B.R. The importance of fungal pathogens and antifungal coatings in medical device infections. Biotechnol. Adv. 2018, 36, 264–280. [Google Scholar] [CrossRef]
  61. 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]
  62. Rismanchian, M.; Babashahi, A.; Goroohi, H.; Shahabouee, M.; Yaghini, J.; Badrian, H. Microflora around teeth and dental implants. Dent. Res. J. 2012, 9, 215. [Google Scholar] [CrossRef]
  63. Arciola, C.R.; Campoccia, D.; Montanaro, L. Implant infections: Adhesion, biofilm formation and immune evasion. Nat. Rev. Microbiol. 2018, 16, 397–409. [Google Scholar] [CrossRef]
  64. Abdallah, M.; Benoliel, C.; Drider, D.; Dhulster, P.; Chihib, N.-E. Biofilm formation and persistence on abiotic surfaces in the context of food and medical environments. Arch. Microbiol. 2014, 196, 453–472. [Google Scholar] [CrossRef]
  65. Rosenthal, V.D.; Bat-Erdene, I.; Gupta, D.; Belkebir, S.; Rajhans, P.; Zand, F.; Myatra, S.N.; Afeef, M.; Tanzi, V.L.; Muralidharan, S.; et al. International Nosocomial Infection Control Consortium (INICC) report, data summary of 45 countries for 2012-2017: Device-associated module. Am. J. Infect. Control 2020, 48, 423–432. [Google Scholar] [CrossRef]
  66. Al Bataineh, M.T.; Alazzam, A. Transforming medical device biofilm control with surface treatment using microfabrication techniques. PLoS ONE 2023, 18, e0292647. [Google Scholar] [CrossRef]
  67. Lourenço, B.N.; Marchioli, G.; Song, W.; Reis, R.L.; van Blitterswijk, C.A.; Karperien, M.; van Apeldoorn, A.; Mano, J.F. Wettability Influences Cell Behavior on Superhydrophobic Surfaces with Different Topographies. Biointerphases 2012, 7, 46. [Google Scholar] [CrossRef]
  68. Vetrivel, A.; Ramasamy, M.; Vetrivel, P.; Natchimuthu, S.; Arunachalam, S.; Kim, G.-S.; Murugesan, R. Pseudomonas aeruginosa Biofilm Formation and Its Control. Biologics 2021, 1, 312–336. [Google Scholar] [CrossRef]
  69. Thi, M.T.T.; Wibowo, D.; Rehm, B.H.A. Pseudomonas aeruginosa Biofilms. Int. J. Mol. Sci. 2020, 21, 8671. [Google Scholar] [CrossRef]
  70. Crivello, G.; Fracchia, L.; Ciardelli, G.; Boffito, M.; Mattu, C. In Vitro Models of Bacterial Biofilms: Innovative Tools to Improve Understanding and Treatment of Infections. Nanomaterials 2023, 13, 904. [Google Scholar] [CrossRef]
  71. Kurmoo, Y.; Hook, A.L.; Harvey, D.; Dubern, J.-F.; Williams, P.; Morgan, S.P.; Korposh, S.; Alexander, M.R. Real time monitoring of biofilm formation on coated medical devices for the reduction and interception of bacterial infections. Biomater. Sci. 2020, 8, 1464–1477. [Google Scholar] [CrossRef] [PubMed]
  72. Wi, Y.M.; Patel, R. Understanding Biofilms and Novel Approaches to the Diagnosis, Prevention, and Treatment of Medical Device-Associated Infections. Infect. Dis. Clin. N. Am. 2018, 32, 915–929. [Google Scholar] [CrossRef]
  73. Li, X.; Sun, L.; Zhang, P.; Wang, Y. Novel Approaches to Combat Medical Device-Associated BioFilms. Coatings 2021, 11, 294. [Google Scholar] [CrossRef]
  74. Mitra, D.; Kang, E.-T.; Neoh, K.G. Polymer-Based Coatings with Integrated Antifouling and Bactericidal Properties for Targeted Biomedical Applications. ACS Appl. Polym. Mater. 2021, 3, 2233–2263. [Google Scholar] [CrossRef]
  75. Chug, M.K.; Brisbois, E.J. Recent Developments in Multifunctional Antimicrobial Surfaces and Applications toward Advanced Nitric Oxide-Based Biomaterials. ACS Mater. Au 2022, 2, 525–551. [Google Scholar] [CrossRef]
  76. Khan, S.A.; Shakoor, A. Recent Strategies and Future Recommendations for the Fabrication of Antimicrobial, Antibiofilm, and Antibiofouling Biomaterials. Int. J. Nanomed. 2023, 18, 3377–3405. [Google Scholar] [CrossRef]
  77. Uneputty, A.; Dávila-Lezama, A.; Garibo, D.; Oknianska, A.; Bogdanchikova, N.; Hernández-Sánchez, J.F.; Susarrey-Arce, A. Strategies applied to modify structured and smooth surfaces: A step closer to reduce bacterial adhesion and biofilm formation. Colloid Interface Sci. Commun. 2022, 46, 100560. [Google Scholar] [CrossRef]
  78. Yang, W.J.; Cai, T.; Neoh, K.-G.; Kang, E.-T.; Teo, S.L.-M.; Rittschof, D. Barnacle Cement as Surface Anchor for “Clicking” of Antifouling and Antimicrobial Polymer Brushes on Stainless Steel. Biomacromolecules 2013, 14, 2041–2051. [Google Scholar] [CrossRef]
  79. Yang, W.J.; Cai, T.; Neoh, K.-G.; Kang, E.-T.; Dickinson, G.H.; Teo, S.L.-M.; Rittschof, D. Biomimetic Anchors for Antifouling and Antibacterial Polymer Brushes on Stainless Steel. Langmuir 2011, 27, 7065–7076. [Google Scholar] [CrossRef]
  80. Zhang, B.; Yan, Q.; Yuan, S.; Zhuang, X.; Zhang, F. Enhanced Antifouling and Anticorrosion Properties of Stainless Steel by Biomimetic Anchoring PEGDMA-Cross-Linking Polycationic Brushes. Ind. Eng. Chem. Res. 2019, 58, 7107–7119. [Google Scholar] [CrossRef]
  81. Xu, G.; Liu, P.; Pranantyo, D.; Xu, L.; Neoh, K.-G.; Kang, E.-T. Antifouling and Antimicrobial Coatings from Zwitterionic and Cationic Binary Polymer Brushes Assembled via “Click” Reactions. Ind. Eng. Chem. Res. 2017, 56, 14479–14488. [Google Scholar] [CrossRef]
  82. Wang, M.; Zheng, Y.; Yin, C.; Dai, S.; Fan, X.; Jiang, Y.; Liu, X.; Fang, J.; Yi, B.; Zhou, Q.; et al. Recent Progress in antibacterial hydrogel coatings for targeting biofilm to prevent orthopedic implant-associated infections. Front. Microbiol. 2023, 14, 1343202. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, N.; Ma, Y.; Shi, H.; Song, Y.; Guo, S.; Yang, S. Mg-, Zn-, and Fe-Based Alloys With Antibacterial Properties as Orthopedic Implant Materials. Front. Bioeng. Biotechnol. 2022, 10, 888084. [Google Scholar] [CrossRef]
  84. Egghe, T.; Morent, R.; Hoogenboom, R.; De Geyter, N. Substrate-independent and widely applicable deposition of antibacterial coatings. Trends Biotechnol. 2023, 41, 63–76. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, J.; Zheng, X.; Jian, R.; Bai, W.; Zheng, G.; Xie, Z.; Lin, Q.; Lin, F.; Xu, Y. In Situ Reduction of Silver Nanoparticles/Urushiol-Based Polybenzoxazine Composite Coatings with Enhanced Antimicrobial and Antifouling Performances. Polymers 2024, 16, 1167. [Google Scholar] [CrossRef]
  86. Simovic, S.; Barnes, T.J.; Tan, A.; Prestidge, C.A. Assembling nanoparticle coatings to improve the drug delivery performance of lipid based colloids. Nanoscale 2012, 4, 1220–1230. [Google Scholar] [CrossRef]
  87. Luchini, A.; Vitiello, G. Understanding the Nano-bio Interfaces: Lipid-Coatings for Inorganic Nanoparticles as Promising Strategy for Biomedical Applications. Front. Chem. 2019, 7, 343. [Google Scholar] [CrossRef] [PubMed]
  88. Jiménez-Jiménez, C.; Manzano, M.; Vallet-Regí, M. Nanoparticles Coated with Cell Membranes for Biomedical Applications. Biology 2020, 9, 406. [Google Scholar] [CrossRef]
  89. Mashaghi, S.; Jadidi, T.; Koenderink, G.; Mashaghi, A. Lipid Nanotechnology. Int. J. Mol. Sci. 2013, 14, 4242–4282. [Google Scholar] [CrossRef]
  90. Kang, M.G.; Khan, F.; Tabassum, N.; Cho, K.J.; Jo, D.M.; Kim, Y.M. Inhibition of Biofilm and Virulence Properties of Pathogenic Bacteria by Silver and Gold Nanoparticles Synthesized from Lactiplantibacillus sp. Strain C1. ACS Omega 2023, 8, 9873–9888. [Google Scholar] [CrossRef]
  91. Tabassum, N.; Khan, F.; Kang, M.G.; Jo, D.M.; Cho, K.J.; Kim, Y.M. Inhibition of Polymicrobial Biofilms of Candida albicans-Staphylococcus aureus/Streptococcus mutans by Fucoidan-Gold Nanoparticles. Mar. Drugs 2023, 21, 123. [Google Scholar] [CrossRef] [PubMed]
  92. Kang, M.G.; Khan, F.; Jo, D.M.; Oh, D.; Tabassum, N.; Kim, Y.M. Antibiofilm and Antivirulence Activities of Gold and Zinc Oxide Nanoparticles Synthesized from Kimchi-Isolated Leuconostoc sp. Strain C2. Antibiotics 2022, 11, 1524. [Google Scholar] [CrossRef] [PubMed]
  93. Khan, F.; Kang, M.G.; Jo, D.M.; Chandika, P.; Jung, W.K.; Kang, H.W.; Kim, Y.M. Phloroglucinol-Gold and -Zinc Oxide Nanoparticles: Antibiofilm and Antivirulence Activities towards Pseudomona saeruginosa PAO1. Mar. Drugs 2021, 19, 601. [Google Scholar] [CrossRef]
  94. Guerrero Correa, M.; Martínez, F.B.; Vidal, C.P.; Streitt, C.; Escrig, J.; de Dicastillo, C.L. Antimicrobial metal-based nanoparticles: A review on their synthesis, types and antimicrobial action. Beilstein J. Nanotechnol. 2020, 11, 1450–1469. [Google Scholar] [CrossRef] [PubMed]
  95. Kang, X.; Yang, X.; He, Y.; Guo, C.; Li, Y.; Ji, H.; Qin, Y.; Wu, L. Strategies and materials for the prevention and treatment of biofilms. Mater. Today Bio 2023, 23, 100827. [Google Scholar] [CrossRef]
  96. Wang, Y.; Wei, T.; Qu, Y.; Zhou, Y.; Zheng, Y.; Huang, C.; Zhang, Y.; Yu, Q.; Chen, H. Smart, Photothermally Activated, Antibacterial Surfaces with Thermally Triggered Bacteria-Releasing Properties. ACS Appl. Mater. Interfaces 2020, 12, 21283–21291. [Google Scholar] [CrossRef]
  97. Yang, Z.; Tu, Q.; Shen, X.; Liu, Y.; Zhang, Q.; Zhao, X.; Maitz, M.; Liu, T.; Hua, Q.; Wang, J.; et al. A facile metal-phenolic-amine strategy for dual-functionalization of blood-contacting devices with antibacterial and anticoagulant properties. Mater. Chem. Front. 2018, 3, 265–275. [Google Scholar] [CrossRef]
  98. Yuan, Y.; Zhang, Y. Enhanced biomimic bactericidal surfaces by coating with positively-charged ZIF nano-dagger arrays. Nanomed. Nanotechnol. Biol. Med. 2017, 13, 2199–2207. [Google Scholar] [CrossRef] [PubMed]
  99. Qu, Y.; Wei, T.; Zhao, J.; Jiang, S.; Yang, P.; Yu, Q.; Chen, H. Regenerable smart antibacterial surfaces: Full removal of killed bacteria via a sequential degradable layer. J. Mater. Chem. B 2018, 6, 3946–3955. [Google Scholar] [CrossRef]
  100. Zou, Y.; Lu, K.; Lin, Y.; Wu, Y.; Wang, Y.; Li, L.; Huang, C.; Zhang, Y.; Brash, J.L.; Chen, H.; et al. Dual-Functional Surfaces Based on an Antifouling Polymer and a Natural Antibiofilm Molecule: Prevention of Biofilm Formation without Using Biocides. ACS Appl. Mater. Interfaces 2021, 13, 45191–45200. [Google Scholar] [CrossRef]
  101. Lin, J.; Hu, J.; Wang, W.; Liu, K.; Zhou, C.; Liu, Z.; Kong, S.; Lin, S.; Deng, Y.; Guo, Z. Thermo and light-responsive strategies of smart titanium-containing composite material surface for enhancing bacterially anti-adhesive property. Chem. Eng. J. 2021, 407, 125783. [Google Scholar] [CrossRef]
  102. Yu, M.; Ding, X.; Zhu, Y.; Wu, S.; Ding, X.; Li, Y.; Yu, B.; Xu, F.-J. Facile Surface Multi-Functionalization of Biomedical Catheters with Dual-Microcrystalline Broad-Spectrum Antibacterial Drugs and Antifouling Poly(ethylene glycol) for Effective Inhibition of Bacterial Infections. ACS Appl. Bio Mater. 2019, 2, 1348–1356. [Google Scholar] [CrossRef]
  103. Vaterrodt, A.; Thallinger, B.; Daumann, K.; Koch, D.; Guebitz, G.M.; Ulbricht, M. Antifouling and Antibacterial Multifunctional Polyzwitterion/Enzyme Coating on Silicone Catheter Material Prepared by Electrostatic Layer-by-Layer Assembly. Langmuir 2016, 32, 1347–1359. [Google Scholar] [CrossRef] [PubMed]
  104. Patel, K.; Kushwaha, P.; Kumar, S.; Kumar, R. Lysine and α-Aminoisobutyric Acid Conjugated Bioinspired Polydopamine Surfaces for the Enhanced Antibacterial Performance of the Foley Catheter. ACS Appl. Bio Mater. 2019, 2, 5799–5809. [Google Scholar] [CrossRef] [PubMed]
  105. Fuchs, A.V.; Ritz, S.; Pütz, S.; Mailänder, V.; Landfester, K.; Ziener, U. Bioinspired phosphorylcholine containing polymer films with silver nanoparticles combining antifouling and antibacterial properties. Biomater. Sci. 2013, 1, 470. [Google Scholar] [CrossRef] [PubMed]
  106. Sun, Y.; Zhao, Y.-Q.; Zeng, Q.; Wu, Y.-W.; Hu, Y.; Duan, S.; Tang, Z.; Xu, F.-J. Dual-Functional Implants with Antibacterial and Osteointegration-Promoting Performances. ACS Appl. Mater. Interfaces 2019, 11, 36449–36457. [Google Scholar] [CrossRef] [PubMed]
  107. Chu, G.; Zhang, C.; Liu, Y.; Cao, Z.; Wang, L.; Chen, Y.; Zhou, W.; Gao, G.; Wang, K.; Cui, D. A Gold Nanocluster Constructed Mixed-Metal Metal–Organic Network Film for Combating Implant-Associated Infections. ACS Nano 2020, 14, 15633–15645. [Google Scholar] [CrossRef] [PubMed]
  108. Silva, D.; Sousa, H.C.d.; Gil, M.H.; Santos, L.F.; Moutinho, G.M.; Serro, A.P.; Saramago, B. Antibacterial layer-by-layer coatings to control drug release from soft contact lenses material. Int. J. Pharm. 2018, 553, 186–200. [Google Scholar] [CrossRef] [PubMed]
  109. Abdulkareem, E.H.; Memarzadeh, K.; Allaker, R.P.; Huang, J.; Pratten, J.; Spratt, D. Anti-biofilm activity of zinc oxide and hydroxyapatite nanoparticles as dental implant coating materials. J. Dent. 2015, 43, 1462–1469. [Google Scholar] [CrossRef]
  110. Lim, D.J.; Skinner, D.; McLemore, J.; Rivers, N.; Elder, J.B.; Allen, M.; Koch, C.; West, J.; Zhang, S.; Thompson, H.M.; et al. In-vitro evaluation of a ciprofloxacin and azithromycin sinus stent for Pseudomonas aeruginosa biofilms. Int. Forum Allergy Rhinol. 2020, 10, 121–127. [Google Scholar] [CrossRef]
  111. Zarghami, V.; Ghorbani, M.; Bagheri, K.P.; Shokrgozar, M.A. Prevention the formation of biofilm on orthopedic implants by melittin thin layer on chitosan/bioactive glass/vancomycin coatings. J. Mater. Sci. Mater. Med. 2021, 32, 75. [Google Scholar] [CrossRef]
  112. Wang, X.; Tan, L.; Liu, X.; Cui, Z.; Yang, X.; Yeung, K.W.K.; Chu, P.K.; Wu, S. Construction of perfluorohexane/IR780@liposome coating on Ti for rapid bacteria killing under permeable near infrared light. Biomater. Sci. 2018, 6, 2460–2471. [Google Scholar] [CrossRef]
  113. Nablo, B.J.; Prichard, H.L.; Butler, R.D.; Klitzman, B.; Schoenfisch, M.H. Inhibition of implant-associated infections via nitric oxide release. Biomaterials 2005, 26, 6984–6990. [Google Scholar] [CrossRef]
  114. Xiong, Z.; Zhang, X.; Zhang, S.; Lei, L.; Ma, W.; Li, D.; Wang, W.; Zhao, Q.; Xing, B. Bacterial toxicity of exfoliated black phosphorus nanosheets. Ecotoxicol. Environ. Saf. 2018, 161, 507–514. [Google Scholar] [CrossRef]
  115. Yazici, H.; O’Neill, M.B.; Kacar, T.; Wilson, B.R.; Oren, E.E.; Sarikaya, M.; Tamerler, C. Engineered Chimeric Peptides as Antimicrobial Surface Coating Agents toward Infection-Free Implants. ACS Appl. Mater. Interfaces 2016, 8, 5070–5081. [Google Scholar] [CrossRef]
  116. Monte-Serrano, M.; Fernandez-Saiz, P.; Ortí-Lucas, R.M. Effective Antimicrobial Coatings Containing Silver-Based Nanoclays and Zinc Pyrithione. J. Microb. Biochem. Technol. 2015, 7, 398–403. [Google Scholar] [CrossRef]
  117. Cruz, A.; Condinho, M.; Carvalho, B.; Arraiano, C.M.; Pobre, V.; Pinto, S.N. The Two Weapons against Bacterial Biofilms: Detection and Treatment. Antibiotics 2021, 10, 1482. [Google Scholar] [CrossRef]
  118. Xu, T.; Xiao, Y.; Wang, H.; Zhu, J.; Lee, Y.; Zhao, J.; Lu, W.; Zhang, H. Characterization of Mixed-Species Biofilms Formed by Four Gut Microbiota. Microorganisms 2022, 10, 2332. [Google Scholar] [CrossRef]
  119. Elias, S.; Banin, E. Multi-species biofilms: Living with friendly neighbors. FEMS Microbiol. Rev. 2012, 36, 990–1004. [Google Scholar] [CrossRef]
  120. Zou, Y.; Liu, C.; Zhang, H.; Wu, Y.; Lin, Y.; Cheng, J.; Lu, K.; Li, L.; Zhang, Y.; Chen, H.; et al. Three lines of defense: A multifunctional coating with anti-adhesion, bacteria-killing and anti-quorum sensing properties for preventing biofilm formation of Pseudomonas aeruginosa. Acta Biomater. 2022, 151, 254–263. [Google Scholar] [CrossRef]
Figure 2. Mechanisms of resistance to antibiotics and their effects. Reprinted from the [27], Copyright © 2021 by the authors and Licensee, Frontiers in Microbiology (Lausanne. Switzerland).
Figure 2. Mechanisms of resistance to antibiotics and their effects. Reprinted from the [27], Copyright © 2021 by the authors and Licensee, Frontiers in Microbiology (Lausanne. Switzerland).
Antibiotics 13 00623 g002
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

Mishra, A.; Aggarwal, A.; Khan, F. Medical Device-Associated Infections Caused by Biofilm-Forming Microbial Pathogens and Controlling Strategies. Antibiotics 2024, 13, 623. https://doi.org/10.3390/antibiotics13070623

AMA Style

Mishra A, Aggarwal A, Khan F. Medical Device-Associated Infections Caused by Biofilm-Forming Microbial Pathogens and Controlling Strategies. Antibiotics. 2024; 13(7):623. https://doi.org/10.3390/antibiotics13070623

Chicago/Turabian Style

Mishra, Akanksha, Ashish Aggarwal, and Fazlurrahman Khan. 2024. "Medical Device-Associated Infections Caused by Biofilm-Forming Microbial Pathogens and Controlling Strategies" Antibiotics 13, no. 7: 623. https://doi.org/10.3390/antibiotics13070623

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