Open access peer-reviewed chapter

Current Approaches for Exploration of Nanoparticles as Antibacterial Agents

Written By

Didem Şen Karaman, Suvi Manner, Adyary Fallarero and Jessica M. Rosenholm

Submitted: 31 October 2016 Reviewed: 24 February 2017 Published: 31 May 2017

DOI: 10.5772/68138

From the Edited Volume

Antibacterial Agents

Edited by Ranjith N. Kumavath

Chapter metrics overview

3,750 Chapter Downloads

View Full Metrics

Abstract

The ascending anxiety regarding antimicrobial resistance as well as the recalcitrant nature of biofilm-associated infections call for the development of alternative strategies to treat bacterial diseases. Nanoparticles have been considered as one of the emerging and promising platforms in this respect. Their unique physical and chemical properties may lead to fine-tuned interactions between them and bacteria. In this chapter, we aim to provide an overview on the use of nanoparticles as antimicrobial agents. Both antibacterial and anti-biofilm activities of nanoparticles and their current approaches will be reviewed. The in vitro methods that are used to evaluate the potency of nanoparticles as antimicrobial agents will be discussed in detail.

Keywords

  • antibacterial agents
  • nanoparticles
  • antibacterial resistance (AMR)
  • anti-biofilm agents
  • in vitro methods

1. Introduction

The term antimicrobial was derived from the Greek words anti (against), micro (little), and bios (life), and it refers to all agents that act against microorganisms. Thus, antimicrobials include agents that act against bacteria (antibacterial), viruses (antiviral), fungi (antifungal), and protozoa (antiprotozoal). Among these, antibacterial agents are by far the most widely known and studied class of antimicrobials. Nowadays, the emergence of antimicrobial resistance (AMR) among the microbial pathogens greatly increases the threat generated by bacterial infections. Drug-resistant bacteria lead to poor clinical outcomes increasing health care costs and mortality. In the US, the estimated health care costs associated with the treatment of infectious diseases are annually more than 120 billion dollars, and further, treatment of infections caused by resistant pathogens costs 5 billion dollars per year [1]. According to the US Center for Disease Control and Prevention (CDC), more than two million antibiotic-resistant infections occur every year in the US and they lead to 23,000 deaths [2]. In the European Union, antibiotic-resistant infections are responsible for 25,000 deaths every year [3]. Both Gram-positive, especially methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-resistant Enterococcus faecium (VRE), and drug-resistant Streptococcus pneumoniae, and Gram-negative bacteria, namely multidrug-resistant Acinetobacter baumannii (MRAB), carbapenem-resistant Enterobacteriaceae (CRE), and Pseudomonas aeruginosa, display resistance to multiple drugs and are of serious concern [4]. In addition, biofilm formation complicates treatment of various infections. Biofilm-related infections, such as chronic wounds and urinary tract infections, pneumonia in cystic fibrosis patients, and infections related to the use of medical devices, comprise up to 80% of all human bacterial infections, and affect millions of people resulting in up to 550,000 deaths every year [57]. The emergence of AMR has been mainly attributed to the inappropriate and excessive use of antimicrobials in humans and animals. In many countries, unregulated availability of antibiotics without prescription results in promotion of overuse [8]. Further, inappropriate prescription of antibiotics also contributes to the promotion of resistant bacteria. The reports have shown that the diagnosis, choice of treatment agent, or duration of antibiotic therapy were incorrect in 30–50% of the cases [9], and 30–60% of antibiotics prescribed in intensive care units have been inappropriate or unnecessary [10]. Extensive agricultural usage of antibiotics in livestock as growth supplement is another reason of excessive consumption of antibiotics by humans, through the intake of resistant bacteria in the food supplies.

It is estimated that more than 70% of all pathogenic bacteria are resistant to at least one of the conventional antibiotics [11]. Antimicrobial resistance is acquired on both cellular and community levels [12]. Acquirement and dissemination of resistance genes is a process that occurs over time. Nevertheless, the evolution of bacterial resistance is substantially accelerated by the dispensable use of antibiotics [13]. Further dissemination of resistance genes between bacterial species has led to the emergence of multidrug-resistant (MDR) bacteria [14]. Community level of resistance is caused by biofilm formation [15]. However, when it comes to biofilms, the genetically transferable, conventional resistance mechanisms are not the leading cause of decreased antimicrobial susceptibility [7]. Bacterial biofilms are structured communities of bacteria embedded in a matrix of extracellular polymeric substances (EPS) that can be formed on variety of surfaces, such as tissues and medical devices [16]. Biofilm is a transient phenotype that makes even sensitive bacteria without known genetic basis for resistance to display remarkably reduced susceptibility to antimicrobials and host immune responses [17]. Many factors contribute to the antimicrobial tolerance of biofilms. First, biofilm matrix can restrict penetration of antibiotics and protects the cells from detrimental insults [18]. Secondly, biofilms comprise a heterogeneous population of cells that are in different physiological states due to decreasing oxygen and nutrient gradients existing between the surface and deeper layers of biofilms. For example, cells located in the deepest part of the biofilm tend to display a slower growth rate and, therefore, are less susceptible to antibiotics that are developed against dividing cells [19, 20]. Further, the non-dividing, dormant population of bacteria, referred also as persister cells, is in well-protected mode and highly tolerant antibiotics. These cells survive even from prolonged antibiotic treatment and serve as reservoirs for infections [21]. Consequently, biofilm bacteria can be up to 1000 times more tolerant to antimicrobial agents than planktonic cells of the same species [22].

Despite this, pharmaceutical companies have substantially declined investments in antimicrobial drug discovery during the past few decades [23]. Antimicrobial drug discovery is not economically attractive, and regulatory requirements have become very challenging [3, 24]. The need of novel bactericidal agents has increased due to the emergence of multi-drug resistant bacterial strains and biofilm-associated infections. Consequently, attention has been especially devoted to emerging nanoparticle-based materials in the field of antimicrobial therapies. In this chapter, the existing nanoparticles as antimicrobial means and the current in vitro test methods that will ease clinical translation of nanomaterials by establishing in vivo relevant data will be described and discussed.

Advertisement

2. Nanotechnology-based antibacterial therapies

Antibacterial applications of nanotechnology are gaining importance to prevent the catastrophic consequences of antibiotic resistance. Nanotechnology can be implemented as preventives, diagnostics, drug carriers, and synergetics in the antibacterial therapies.

The unique properties of nanomaterials compared to its bulk form make them favourable for antibacterial therapies. Many inorganic and organic nanomaterials represent inherent antibacterial properties that are not expressed in their bulk form. Fast and sensitive bacterial detection can be provided with nanoparticle-based approaches. Furthermore, nanoparticles offer discrete advantages as antibacterial drug delivery systems. They can be designed as targeted, environmentally responsive, combinatorial delivery systems [25]. Another approach of nanomaterials for the antibacterial therapy is as vaccine that contains nanoparticles as adjuvants or delivery vehicles, which provoke immune responses against bacterial infection. In the following parts in section 2.1 and 2.2, the existing nanotechnologies for the antibacterial delivery systems and inherently antibacterial nanoparticles will be discussed in detail.

2.1. Nanomaterials as antibacterial delivery systems

The existing disadvantages of conventional antibiotics can be solved to some extent by using nanomaterial-based antimicrobial delivery systems. In such approaches, the conventional antibiotics can be loaded into the nanoparticles through physical encapsulation, adsorption, or chemical conjugation. By this way, the pharmacokinetics and therapeutic index of the drug can ideally be improved compared to the free form of the drug. The aimed-for advantages are provided by the improved serum solubility, prolonged systemic circulation lifetime of the drug, targeted delivery of the drug to the site of infection, sustained and controlled release of the drug, and also combinatorial drug delivery to the site of interest that could be reached by virtue of the nanoscopic delivery system [2628]. This rationale of nanotherapeutics in this case aimed to enhance the therapeutic effect and minimize the side effects of antibiotics, starts with the appropriate design of nanoparticles. In nanoparticles design, the particle size, surface properties, and the release profile of the therapeutic agent have vital impact on the success of the therapeutic approach. Various nanoparticles-based drug delivery systems have been designed and investigated for improving the efficacy of antibiotics of the administered drugs, the most common of which shall be outlined in the following.

Lipid-based nanoparticles are widely used for the delivery of antibacterial agents. They can be designed as liposomes, solid lipid nanoparticles (SLN), and nanostructured lipid carriers (NLC). Liposomes are one of the most studied nanosystems for antimicrobial therapy in various diseases. Liposomes are spherical lipid vesicles with bilayered membrane structure, consisting of amphiphilic lipid molecules. Since their structure is similar to the bacterial cell membrane, efficacious interaction between liposomes and cells can be obtained. These interactions may create adsorption, endocytosis, lipid exchange, and fusion of the liposomes. Especially, the design of liposomes that cause fusion and is known as fusogenic is the most attractive one in the sense of efficiency. Fusogenic liposomes are able to destabilize the bacterial membrane and release their therapeutic content inside the cells [29, 30]. The structure of liposomes, where an aqueous cavity is surrounded by lipid membranes, empowers them to transport both lipophilic and hydrophilic drugs (in lipid bilayers and aqueous compartments, respectively) without chemical modification, protecting them from degradation [31]. SLNs are composed of a solid lipid core stabilized by surfactants and are moderately amorphous structures in which bilayers are not distinguished. They can provide long-term stability and better incorporation efficacy for hydrophobic drugs and can be easily scaled-up in production. NLCs were developed in order to overcome the limitations of SLNs regarding low-loading capacity for nonhydrophobic drugs and their stability issues. In the NLC structure, liquid lipids are used to stabilize the construct, which allows a biphasic drug release profile with initial burst release continued with sustained drug release. Liposomes have shown to be successful in combating resistant pathogens. Especially, their modified designs are used to improve the potency of formulations in bacterial resistance and clearance [32]. Additionally, researchers have confirmed the feasibility of SLN and NCL as drug carriers, however, their advantages over liposomes have not been confirmed with human data [33]. Most of the research on SLN and NCL as antimicrobial carriers are still in the preclinical stage.

To date, a significant number of reports on the activity of antibiotic-conjugated polymeric nanoparticles against various infections, including those caused by drug-resistant pathogens, have been published [34]. Notably, high biocompatibility of these structures, additional to improved pharmacokinetic properties, supports the potential of these nanosystems as new tools to treat infections. Polymeric nanoparticles can be prepared from natural and synthetic polymers with the prerequisite of biocompatibility and biodegradability. In the polymeric antibacterial drug delivery systems, drug molecules can be incorporated in the internal part of the particles, on the surface of polymeric nanocarriers with covalent or non-covalent bonds, imprinted in the polymeric nanoparticles or encapsulated in the stimuli-responsive shell of polymeric nanoparticles [34]. The encapsulation route of the drug into the polymeric nanoparticle drug delivery system plays a key role in the nanocarriers’ pharmacokinetic profile. The action mechanism of the polymeric nanoparticles is defined by the physicochemical properties and the composition of the particles. Polymeric nanoparticles may interact with the bacterial cell wall via passive or active targeting. Passive targeting is based on particle size and the ability of particles to disturb the structure of bacterial membrane leading to pore formation in the membrane. For active targeting of polymeric nanoparticles, the surface of polymeric nanoparticles is usually functionalized with specific antibodies and aptamer bacteriophage proteins providing specific identification for the detection of pathogens and interaction between the particles and pathogens. The reported studies reveal that both the active and passive targeting strategies to deliver antimicrobial agents with polymeric nanoparticles improve their activities compared to their free forms [3537].

Dendrimers are highly branched macromolecules employed as antibacterial drug delivery systems. The unique properties of dendrimers, such as well-defined 3D structures, available functional groups, and their ability to mimic cell membranes, make them potential drug carriers. Both hydrophobic and hydrophilic drug molecules can be incorporated separately or at the same time into dendrimer structures. Lipophilic molecules can be incorporated inside the cavity of dendrimers, and hydrophilic agents can be covalently or physically attached to the surfaces of dendrimers. The antibacterial can be used in the building of dendrimer blocks, whereby the synthesized dendrimers themselves become potent antimicrobials. Dendrimers aid to improve the solubility, penetration, and controlled release of the drug molecules. Currently, the existing research in the design of dendrimers as antibacterial drug delivery systems also focuses on species-selective dendrimer biocide formulations. For instance, peptide, glycol, and glycopeptide dendrimer designs provide effective therapy for the bacterial infections.

An inorganic nanomaterial, in contrast to the organic materials listed above, which has also shown promise for antibacterial therapies is mesoporous silica nanoparticles (MSNs). In the design of MSN-based drug delivery systems, their advantageous characteristics (biocompatibility, high surface area, tunable particle diameter, mesoporous structure, and ease of functionalization) have been exploited. The designs with targeted and sustained release mechanisms make them powerful candidates also for antibacterial therapies. In the use of MSNs as drug delivery vehicles for antibacterial therapeutics, their surface functionality along with the size and shape are crucial parameters to improve and optimize the efficacy [38]. Their surface functionalities can be modified to target both planktonic bacteria and biofilms [39]. In recent studies, the utility of MSN for efficient antibiotic delivery [36, 40, 41] and hybrid antibacterial materials preparation by incorporating antibacterial enzymes [42], peptides [43], metal ions/particles [44], and polymers (surface modifiers) [45] to MSNs has been reported.

For rational and efficient utilization of these nanomaterial-based drug delivery systems, systematic investigation of pharmacokinetics and biodistribution should be carried out. The pharmacokinetics and biodistribution of nanoparticles are defined by their physicochemical properties [38]. Apart from their physicochemical properties, the administration routes and their elimination from the body need to be systematically evaluated. Hence, thorough evaluation of the current nanoparticle-based drug delivery systems in antibacterial therapies is important for their translation into the clinic. To date, four liposomal/lipid complex drug delivery systems for antibiotic delivery have been approved for use in human patients, including Abelcet, AmBisome, Amphotec, and Fungisome [46]. This should come as no surprise with regard that a liposomal formulation was the first nanodrug to hit the market in 1995 (Doxil®), and they have been studied since the early 1980s.

2.2. Nanomaterials as active antibacterial agents

Various types of inorganic and organic nanoparticles have been utilized as antibacterial agents. The inherent antibacterial properties of some metals and metal oxides have been known for centuries. An important advantage of antibacterial metal and metal oxide nanoparticles is that they have multiple modes of action, which is why microbes can scarcely develop resistance to them.

Among the inorganic antibacterial particles, silver nanoparticles are the most intensively investigated ones and capable to kill both Gram-positive and Gram-negative bacteria, having even shown to be effective against drug-resistant species [46]. Besides silver nanoparticles, other metal nanomaterials have also been studied for antimicrobial treatment, including gold [47], copper [48, 49], tellurium [50, 51], and bismuth [52]. Moreover, many studies have revealed the antibacterial activity of metal oxide nanomaterials, such as zinc oxide (ZnO) [53], copper oxide (CuO) [54, 55], magnesium oxide (MgO), nitric oxide (NO) [56], titanium dioxide (TiO2) [57], aluminum oxide (Al2O3) [58], magnetic iron oxide (α-Fe2O3) [59], and cerium oxide (CeO2) [60] nanoparticles. The toxic mode of metal and metal oxide nanoparticles against bacterial cells has been associated with ROS generation and membrane disruption [61]. According to literature findings, the release of ions is designated as the driving force behind the antimicrobial properties of antibacterial nanoparticles.

Cationic polymeric nanoparticles have been considered as promising organic antibacterial nanoparticles. They can be composed of natural or synthetic cationic polymers. The antibacterial polymeric nanoparticles kill microorganisms upon their contact with bacterial cells due to the strong interaction of their cationic surfaces with the bacterial cells [62]. The mechanisms of action have been proposed for how these cationic groups are able to disrupt the bacterial cell membrane, with some requiring hydrophobic chains of certain lengths to penetrate and burst the bacterial membrane. Moreover, different polymeric nanosized antibacterials with long-term antibacterial activity, chemically stable, and ability to bind to surfaces of interest have been reported. These include lipid nanoparticles, quaternary ammonium polyethyleneimine-incorporated polymeric nanoparticles [63, 64], chitosan [65], and self-assembled peptide nanoparticles [66]. In addition to the above-mentioned metallic and polymeric nanoparticles, carbon-based nanostructures have shown antibacterial effects. For instance, the antibacterial activity of fullerene [67] and carbon nanotubes [68, 69] (single-walled or multi-walled) derivatives have been observed. However, the antibacterial mechanism of carbon-based nanostructures is still under debate and has not received particular attention, possibly due to the difficulties of their dispersion in water, especially in case of the carbon nanotubes [70].

Advertisement

3. Advantages and challenges of nanotechnology-based antibacterial treatments

Promising approaches for the effective delivery of therapeutic compounds can be provided by the use of nanoparticles as drug carriers. Literature findings and clinical results have surely presented several clinical advantages of antimicrobial nanoparticles and their utilization as drug carrier systems. Antimicrobial nanoparticles are of great interest as they provide a number of benefits over free antimicrobial agents. In detail, nanocarriers can conquer the solubility and stability issues and reduce side effects [62]. With the use of nanocarriers in the delivery of drugs, combination drug therapy can be achieved by incorporating two or more drugs or different therapeutic modalities into the carrier matrix. The surface modifications can be carried out by conjugating targeting ligands on the nanocarriers that are not known by the immune system and specifically targeted to special microorganisms. Administration of antimicrobial agents using nanoparticles can increase the overall pharmacokinetics by progressing therapeutic index, extending drug circulation, and providing controlled drug release. Multiple mechanisms of action can be provided by the antibacterial nanoparticles, which prevent the development of antibacterial resistance by many pathogenic bacteria. Several routes of administration, including oral, nasal, parenteral, intraocular, and so on, can be employed with the nanotechnology-based antibacterial treatments.

The significant advantages of nanomaterials as antimicrobial agents are their modularity in design, enabling a multimodal approach that makes it especially difficult for bacteria to develop resistance mechanisms against these. Namely, a nanotechnology-based antibacterial agent can be constructed out of several components that possess antimicrobial activities in themselves, such as, for instant, be composed of an antibacterial core material (e.g. metal or metal oxide) surrounded with an antibacterial polymeric shell or coating, in which antibiotic drugs could be incorporated [71]. The core material could further be “prickly,” which physically can destroy the bacterial cell wall by a “nano-piercing” process once the polymeric shell has been dissolved, leading to the disruption of bacterial integrity and lysis, as presented in a recent study by Wu et al. where zinc-doped copper oxide prickly nanoparticles exhibited high bacterial killing efficiency owing to the provided core particle nanostructure [72]. Furthermore, varying possibilities for combination therapy together with existing (commercial) antibiotics to reach synergistic effects are evident [14, 73, 74].

Although nanoparticle-based antibacterial treatments promise significant benefits and advances in addressing the key hurdles in treating infectious diseases, there are challenges in translating this exciting technology for clinical use. These include thoroughly evaluating the interactions of nanoparticles with cells, tissues, and organs, which accordingly recalibrates doses and identifies proper administration routes to obtain desired therapeutic effects. Hence, to provide a clinical translation of nanomaterials, standardized in vitro experimentations that will provide in vivo relevant data should be established [75]. In section 4, we describe existing in vitro methods for testing antimicrobial activity. In addition, the current methods commonly employed in testing the antibacterial and anti-biofilm activity of nanoparticles are discussed with the relevancy and pitfalls.

Advertisement

4. Methods to study antimicrobial activity in vitro

Traditionally, antimicrobial research has focused on planktonic bacteria, and there is a variety of test methods available for evaluation of antimicrobial activity against planktonic cells [76]. The Clinical and Laboratory Standards Institute (CLSI) and European Committee on Antimicrobial Susceptibility Testing (EUCAST) are the major contributors to harmonized antimicrobial susceptibility testing (AST) providing uniform procedures for testing and analysis of antibacterial activity. CLSI standards clearly define the specific and essential requirements for materials, methods, and methodologies that need to be followed without any modifications because deviations from the approved procedures might affect the experimental outcome [77]. All the critical elements for the testing, such as the culture medium, inoculum density, and incubation conditions, are listed. Unlike the antibacterial assays, standardized methods for anti-biofilm studies are scarce. No standard methods have been approved by CLSI or EUCAST for evaluation of antimicrobial activity against biofilms. Altogether, five standards (ASTM E2196, ASTM E2647, ASTM E2562, ASTM E2799, and ASTM E2871) set by the American Society for Testing and Materials (ASTM) exist, and they all are applicable as such only for Pseudomonas aeruginosa biofilms [78]. Moreover, only one standard is intended for susceptibility testing. Further, due to distinct phenotype and heterogeneity of biofilm bacteria, conventional in vitro methods used for assessment of bacterial susceptibility to antimicrobials are not appropriate for biofilm-growing bacteria [79]. Lack of standardization makes a comparison of the test results difficult, and further, can lead to the generation of conflicting data between studies since the experimental outcome is strongly dependent on the assay conditions and materials employed in the testing.

4.1. Antimicrobial susceptibility testing

In vitro susceptibility assays are performed to assess the antimicrobial susceptibility of microorganisms in order to provide efficient treatment for infections [80]. Moreover, they are used for resistance surveillance and in research laboratories to study antimicrobial activity of antimicrobial agents. Determination of the minimum inhibitory concentration (MIC) is the most widely used measure of the antimicrobial susceptibility of microorganisms. The MIC is defined as the lowest concentration of an antimicrobial (expressed as mg/L or μg/μL) required to inhibit the visible growth after overnight incubation. In addition, the minimum bactericidal concentration (MBC) is frequently used. The MBC refers to the lowest concentration of an antimicrobial that kills 99.9% of the original inoculum after a certain incubation period (Figure 1). These measures can be easily achieved and compared with each other when standardized methodologies and protocols are followed. A number of guidelines and standardized protocols for MIC and MBC determinations exist that include information on each step of the testing, ranging from storage and preparation of antibiotic stocks to interpretation of the results against particular microorganisms [81].

4.1.1. Dilution methods

Dilution methods are used to determine the MIC values of the antimicrobial agents. Moreover, they serve as reference methods for antimicrobial susceptibility testing. The minimum inhibitory concentration can be determined both on agar (agar dilution) and in broth (broth dilution). Standards for agar and broth dilution techniques used to assess the in vitro susceptibility of aerobically grown bacteria are described in CLSI document M07-A9. The document covers the performance, experimental conditions, reporting of the minimal inhibitory concentration (MIC), quality control procedures, and limitations of the recommended methods.

Figure 1.

Minimum inhibitory concentration (MIC) versus minimum bactericidal concentration (MBC).

When conducted on agar, a two-fold diluted series of the antimicrobial agent is incorporated into agar medium followed by inoculation of standardized suspension of the given organism onto the agar plate. Broth dilution can be performed in tubes (macrodilution, volume 2 ml) or in microtiter well plates (microdilution, volume ≤500 μl) containing a two-fold diluted series of antimicrobial agent prepared in the liquid growth medium that is inoculated with a standard inoculum of bacteria followed by a defined incubation period under particular conditions. After the incubation, the outcome is read based on turbidity or growth zones, and the MIC is defined. The MIC value can also be utilized to distinguish between bactericidal and bacteriostatic activities. Alternatively, when using microdilution, the MIC can be determined spectrophotometrically according to the EUCAST protocol [81]. In addition to CLSI and EUCAST standards, ISO-20776-1 standard proposes acceptable performance criteria for microdilution method. After broth dilutions, the MBC can be determined by sub-culturing the samples from tubes or wells and plating on agar to determine the number of cells (CFU/ml) after incubation for 24 h. Then, MBC is defined as the lowest concentration at which 99.9% of the final inoculum is killed (Figure 1). The main advantage of the dilution method is a generation of quantitative data, the MIC value. Moreover, the assay is overall reproducible, and small amounts of antimicrobials are needed when the microdilution method is utilized. By contrast, large amounts of antimicrobial agents are needed in macrodilution testing. The main disadvantage of the method is several steps in sample preparation, which in turn, increases the possibility of errors.

4.1.2. Diffusion methods

Standards for antimicrobial disk susceptibility test are proposed in the document M02-A12 by the CLSI. Specifications for the agar (type, depth), concentration range of the test antimicrobial, concentration of the microorganism and incubation conditions (time, temperature and atmosphere) are included. Also, interpretation of the results, quality control procedures, and limitations of the methods when used for susceptibility testing of aerobically growing bacteria are described. Agar disk diffusion method is routinely applied to the in vitro susceptibility testing of bacteria. The Kirby-Bauer test is the most thoroughly described disk diffusion method for which interpretive standards exist (Figure 2). A filter paper disc impregnated with antimicrobials at different concentrations is placed on an agar plate, and the antimicrobial diffuses from the disc into the agar around the disc. Thereafter, the plate is inoculated with a standardized suspension of a microorganism followed by incubation. After the incubation period, the growth inhibition zones around the discs are measured. The diameter of the zone is depending on the antimicrobial susceptibility of microorganism. The disc diffusion test is simple to perform, but it only provides qualitative data and categorizes microorganisms as susceptible, intermediate, and resistant based on the susceptibility. Thus, it cannot be used to distinguish between bactericidal and bacteriostatic effects. Commercially available zone reader systems can be utilized to calculate an approximate MIC value by comparing the zone size and standard curve of the bacteria and antibiotic stored in an algorithm [82].

Figure 2.

Disk diffusion test and antimicrobial gradient diffusion method. On the left, agar plate showing zone of inhibition by different antimicrobials of diameter of zones of inhibition refers to the susceptibility of a microorganism.

4.1.3. Combined dilution and diffusion method

The antimicrobial gradient diffusion method is based on the establishment of an antimicrobial concentration gradient in the agar medium to measure the antimicrobial susceptibility. Thin plastic test strips marked with concentration scale and impregnated with antibiotic concentration gradient are placed on agar plates that have been inoculated with a standardized inoculum (Figure 2). After incubation overnight, the experimental outcome is read, and the MIC can be determined by the intersection of the lowest part of the ellipse-shaped growth inhibition area with the test strip. E-Test is the commercially available test for this purpose.

4.1.4. Time-kill test

Time-kill assay is complementary to MIC and MBC determinations. It provides information on the dynamic interaction between the antimicrobial and microorganism, thus revealing whether the antimicrobial effect is time or concentration dependent. Such activities can be investigated utilizing the standard protocol M26-A by CLSI and ASTM2315. These protocols are frequently modified. Time-kill assay is usually conducted at a concentration twice or four times the MIC. Standardized inoculum is added to a nutrient broth containing the antimicrobial at various concentrations. A sample is taken from each concentration at the inoculation time and after selected time points. Samples are serially diluted and viable plate counts are performed. The kill curves are constructed by plotting the log CFU against time. A 3-log reduction in cell counts corresponding the killing of 99.9% is considered as significant antimicrobial activity [83]. Alternatively, measurement of luminescence can be utilized to determine the time-kill relationship. Luminescence is detected by the ATP assay, in which adenosine triphosphate (ATP), an indicator for bacterial viability is quantified, and the number of viable cells is determined based on the amount of ATP. The assay uses luciferase reaction, in which luciferin is converted to oxyluciferin in the presence of molecular oxygen and ATP, and generates light by luminescence. Luminescent signal is proportional to the number of viable cells [84].

4.1.5. Methods to study antimicrobial susceptibility of biofilms

Since conventional susceptibility testing methods are not applicable for biofilms, and the MIC values do not provide a valid estimation of the antibiotic concentration needed to treat biofilm-related infections, the minimum biofilm inhibitory concentration (MBIC) has been determined instead. The MBIC determines the susceptibility of bacteria when the biofilm is forming and refers to the lowest concentration of an antimicrobial, in which no visible growth occurs after exposure to antimicrobial after the incubation period, and it can be recorded by optical reading [85]. Based upon the viable plate counts, the MBIC is defined as the lowest antimicrobial concentration in which there is no time-dependent increase in the mean number of viable cells between two exposure times [86]. Moreover, the Calgary Biofilm Device (CBD) can be used for determination of MBIC, as well as the minimum biofilm eradication concentration (MBEC), which is defined as the lowest concentration of an antimicrobial required to eradicate the established biofilms [87] along with susceptibility of planktonic bacteria (MIC) [22]. The commercially available CBD consists of 96 pegs mounted on the lid of a 96-microtiter well plate. Biofilms are first formed on the pegs for a defined time period. After the incubation period, the lid is transferred to another 96-well plate containing antimicrobials in fresh culture media at various concentrations. The MBEC is defined as the concentration of antimicrobial in which no visible growth can be detected [88]. ASTM 2799 standard describes the operational parameters required to grow, treat, sample, and analyze Pseudomonas aeruginosa biofilms using MBEC assay. In this assay, the experimental outcome is reported as log10 colony forming units (CFUs) per surface area, and the antimicrobial efficacy is assessed as the log10 reduction of viable cells. The experimental outcome can be evaluated using the CLSI guidelines for interpretation. However, breakpoints for resistance are not available, and those for planktonic bacteria are not applicable for biofilm bacteria [86]. Even though the ASTM protocol describes the specific experimental conditions only for P. aeruginosa biofilms, it can be used for other species with some modifications.

4.2. Methods to assess the anti-biofilm activity

Several assays with distinct endpoints are essential for the determination of the antimicrobial activity against biofilms. These assays rely on quantification of (i) viable cells in the biofilm, (ii) total biomass and (iii) biofilm matrix. An ideal anti-biofilm agent would target biofilm viability, biomass, and the matrix. Most of the assays are based on various staining methods. Several models have been proposed for evaluation of antimicrobial activity on biofilms. Depending on the flow of nutrients and bypass the waste products, biofilm models can be classified as closed and open systems [89]. Microtiter well plate-based assays are the most commonly used, while the Calgary biofilm device, substratum suspending reactors, and the flow cell system are the most widely used biofilm models for in vitro susceptibility testing [79]. Because the experimental outcome is affected by the choice of the model system, it is utmost important to select a model, in which biofilms can be formed in conditions close to in vivo settings.

4.2.1. Crystal violet staining

Crystal violet (Hexamethyl pararosaniline chloride, CV) assay is not only one of the oldest but also most widely used staining methods applied to biofilm quantification [90, 91]. Crystal violet is an inexpensive and basic dye that is used to measure the effects on total biomass of biofilms. Crystal violet binds indifferently to both negatively charged bacteria and polysaccharides present in the EPS. After staining, adsorbed dye is eluted in a solvent, such as acetic acid or ethanol. The amount of the dye solubilized by the solvent is measured spectrophotometrically, and it is directly proportional to biofilm biomass [92, 93]. Disadvantages of the method are shortcomings in its dynamic range, laboriousness, and low reproducibility [92]. Experimental conditions, bacterial species, concentration, and nature of the solvent used, as well as incubation time are crucial steps that affect the experimental outcome. Furthermore, the assay does not sort out living or dead cells or biofilm matrix, thus not providing any information on the number of living bacterial cells [93] and, more importantly, imprecise information on the antimicrobial activity. However, the method can be used for both Gram-negative and Gram-positive bacteria and fungi, but the optimal assay conditions, such as temperature, incubation time, and solvent, vary between species [94, 95].

4.2.2. Resazurin staining

Resazurin, also known as alamar blue (7-hydroxy-3H-phenoxazin-3-one-10-oxide), is a non-invasive, non-fluorescent dye, which is reduced to resorufin, a pink, fluorescent dye as a result of metabolically active cells and bacterial viability. Resorufin is detected spectrophotometrically to determine the viable cells [96]. Resazurin staining assay can be used to assess the antimicrobial activity based on the effects on viability of various microorganisms grown in suspensions or as biofilms [97]. However, time of resazurin reduction is species and strain specific. Consequently, the experimental conditions, such as incubation time and resazurin concentration, need to be optimized to obtain reproducible data [93, 98].

4.2.3. Fluorescein diacetate (FDA) assay

The assay is used to measure nonspecific esterase activity of viable microbial cells that converts colorless, nonfluorescent FDA to fluorescein, which is a green fluorescent compound that can be quantified fluorometrically [88]. The assay is not widely used because the dye rapidly leaks from the cells and is unstable. Moreover, hydrolysis of FDA to fluorescein in the absence of live cells and quenching of fluorescence by assay solutions may also occur under certain conditions, thus making the reliability of the assay questionable. However, the assay enables biofilm quantification without removing biofilm from the place where it has been formed, allowing the quantification of entire biofilm [84, 99].

4.2.4. Other viability assays

Biofilm viability can be assessed using tetrazolium salt reduction assay, in which tetrazolium salts, such as MTT (3-[4, 5-dimethylthiazol-2-yl]-2,5-diphenyltetrazolium bromide), XTT (2,3-bis-[2-methoxy-4-nitro-5-sulfophenyl]-2H-tetrazolium-5-carboxanilide), and MTS (3-[4, 5-dimethylthiazol-2-yl]-5-[3-carboxymethoxyphenyl]-2-[4-sulfophenyl]-2H-tetrazolium), are reduced to formazan dyes. The color change can be quantified spectrophotometrically. The amount of formazan dye produced is proportional to the number of viable cells. However, the amount of reductase enzyme produced by cells is dependent on metabolic activity [100, 101]. In other words, cells that are metabolically less active when encased in a biofilm produce also reduced amount of reductase enzyme that can lead to identification of artificially low quantities of cells when viable cells are detected using this assay. Moreover, high densities of bacteria may produce the maximum amount of formazan product leading to optical densities comparable with even higher cell densities [102].

4.2.5. Viable plate counts (colony forming units)

Viable plate counts (colony forming unit counting) are used to assess the biofilm viability based on cell counting. The assay can be used to evaluate the efficacy of antimicrobials to prevent biofilm formation or eradicate pre-formed biofilms, respectively [103]. Depending on the assay mode, bacteria and antimicrobials are added simultaneously to the microtiter well plates or biofilms are first allowed to form followed by the exposure to antimicrobial. After an incubation period, biofilms are dislodged and disaggregated. The resulting suspensions are carefully homogenized, for example, by vortexing or sonication [93]. Subsequently, suspensions are serially diluted and plated or spread on agar. Colony forming units (CFUs) per surface area or volume are counted after an incubation period. The experimental outcome can be evaluated as reduction in CFUs compared to untreated control biofilms. The method gives accurate information on bacterial viability [94]. However, it is regarded as a time-consuming and laborious methodology because of serial dilutions and plating. Additionally, special attention needs to be addressed to the detachment and disaggregation steps to avoid false negative/positive results. The complete recovery and disaggregation of biofilm need to be ensured by applying methods that do not affect viability of the biofilm cells [93, 104].

4.3. Testing of antimicrobial activities of nanoparticles

With respect to the evaluation of antimicrobial activity of nanoparticles, only one specific standard is set by the International Organization for Standardization (ISO), and it is intended for determination of the antimicrobial potency of silver nanoparticles against Staphylococcus aureus by measuring the release of muramic acid using gas chromatography-mass spectrometry (GC-MS) [105]. Thus, conventional methodologies for antimicrobial susceptibility testing have been not only adapted but also modified for investigation of antimicrobial activity of various nanoparticles [102]. In antimicrobial testing, experimental outcome is affected by many factors, such as solvent, inoculum preparation, type of the culture media, and incubation conditions, and these factors have been also found to be influential when testing antimicrobial activities of nanoparticles [76, 106]. For example, choice of the culture media can have a substantial impact on antimicrobial activity of nanoparticles. Media composition and its pH may affect the experimental outcome because of the impact on physicochemical properties of nanoparticles, and further, media type (solid vs. liquid) has found to profoundly influence the antimicrobial activity of nanoparticles [1, 107, 108].

Both standardized and modified microdilution and macrodilution methods have been applied to the determination of MIC and MBC values to evaluate the antimicrobial susceptibility of se-veral microorganisms to nanoparticles [76, 109, 110116]. Additionally, resazurin staining assay has been employed to determine the MIC. The MIC was recorded as the lowest concentration at which color change from blue to pink occurred [117]. Even though standardized antimicrobial susceptibility testing protocols can be followed, no standards describing the synthesis of nanoparticles exist. Differences in the synthesis methodology are known to impact the particle size and chemical composition of the nanoparticles, which, in turn, can further affect the antimicrobial activity and cause variability in the experimental outcome [1, 118]. Hence, the impact of such factors has to be taken into consideration when results between studies are compared.

Diffusion methods can be applied alongside dilution assays to confirm the antimicrobial susceptibility of microorganisms [108, 119, 120]. Agar disk diffusion tests performed both according to the standardized protocols and with modifications are frequently used for susceptibility assessment and evaluation of antimicrobial activity of nanoparticles [109, 121124]. Paper disc method has been employed as an alternative to standardized single disk method [112]. Further, agar well diffusion assays have been successfully utilized for the evaluation of antimicrobial effects of nanoparticles [120, 125, 126].

Antimicrobial effects can also be determined by reading the optical density [125]. Even though measurement of optical density is a straightforward method, it is not the most suitable for measuring the activity of nanoparticles because nanoparticles as such can also interfere with the optical density [46, 127]. Viable plate counts have been frequently performed according to the various protocols to assess the antimicrobial efficacy of nanoparticles against both planktonic and biofilm-growing bacteria [109, 128, 110112, 116]. Samples can either be spread or pipetted on agar plates followed by overnight incubation and determination of the number of CFUs [116, 111, 113115]. Especially, when quantifying the biofilm bacteria, efficient disaggregation of samples is of great importance to avoid false positive results.

Crystal violet staining is the most widely applied staining assay to investigate the antimicrobial activity of the various nanoparticles against biofilm-growing bacteria [129, 130, 121, 116, 131, 132]. By combining assays that quantify different features of biofilms, more relevant information on the activity of nanoparticles can be obtained. In that context, effects of nanoparticles on biofilm inhibition have been studied using viable plate counts and crystal violet staining in parallel [133, 134]. Additionally, LIVE/DEAD and crystal violet staining has been combined for the same purpose [133, 134]. Crystal violet staining has been also used together with resazurin staining assay to assess the impact of nanoparticles on total biomass, including the matrix components and biofilm viability, respectively [135]. Further, crystal violet along with phenol has been applied to quantify the effects on biomass and EPS [136]. Antimicrobial agents displaying bactericidal effects have usually an impact on both viability and biomass, while antimicrobials acting like detergents affect only the biomass [137]. Further, to distinguish between bactericidal and bacteriostatic activities, LIVE and DEAD staining of bacterial biofilms can be conducted using a combination of fluorescein diacetate (FDA) and propidium iodide (PI) dyes or by commercially available LIVE/DEAD kit containing propidium iodide (PI) and SYTO9 fluorescent dyes [136, 138]. Commonly used microtiter well plate-based assays are summarized in Table 1.

Assays Endpoint Read-out Planktonic bacteria Biofilm Ref.
Measurement of optical density Growth inhibition, MIC Absorbance X X [119, 125]
Resazurin staining assay Viability, MIC Fluorescence X X [117, 136]
Tetrazolium salt reduction assay (MTT, XTT, MTS) Viability Fluorescence X X [111, 132, 142, 143]
Crystal violet staining assay Biofilm biomass, MBIC Absorbance X [132]
Fluorescein diacetate (FDA) assay Viability Fluorescence X X [143, 144]

Table 1.

Microtiter well plate–based assays used to investigate antimicrobial activity of various nanoparticles.

In order to gain information on the mechanistic action of nanoparticles, antimicrobial assays can be conducted in two modes, prior to and post biofilm formation. In the pre-exposure mode, nanoparticles and bacteria are simultaneously added, whereas in post-exposure mode, biofilms are first allowed to form, followed by the exposure to nanoparticles. Crystal violet staining has been used to evaluate the impact of nanoparticles on biofilm formation and eradication, respectively [139, 140, 127, 141], and viable plate counts have been utilized in the investigation of the antimicrobial efficacy of nanoparticles in the prevention of biofilm formation and eradication of pre-formed biofilms [111, 112].

Advertisement

5. Conclusion

There is a strong demand to develop novel antimicrobial materials, and the emergence of nano-technology is creating a variety of options in this respect. Numerous nanoparticles exhibit antibacterial activity against several bacterial species. Today, nanomaterials are a promising platform to control bacterial infections in a broad range of applications. However, the absence of standardizations in testing methods leads to inconsistency in results. The foremost requirement of the assays applied to the evaluation of antimicrobial activity is reproducibility. Antimicrobial activity should be tested against various microorganisms, preferably against representatives of both Gram-negative and Gram-positive species. Moreover, a combination of several assays is preferred to confirm the activity. Several standardized methodologies exist for testing the antibacterial activity of conventional agents against planktonic bacteria. These methods are not applicable for biofilms, and further, they do not allow the prediction of the in vivo activity against biofilm-growing bacteria. The majority of the bacterial infections are nowadays attributed to biofilm formation, standardized test methods are urgently needed for more accurate evaluation of antimicrobial activity against biofilms. Even though nanoparticles represent a prominent approach to combat both multi-drug resistant and biofilm-related infections, lack of standardization of synthesis and testing methodologies is a significant concern. Several assays have been reported so far to test the anti-biofilm activity of nanoparticles-containing formulations. However, since antimicrobial assays are sensitive to variation in assay conditions, only standardization of these methods enables comparative analysis between studies.

References

  1. 1. Rizzello L, Pompa PP. Nanosilver-based antibacterial drugs and devices: Mechanisms, methodological drawbacks, and guidelines. Chem Soc Rev. 2014;43(5):1501–18.
  2. 2. Prestinaci F, Pezzotti P, Pantosti A. Antimicrobial resistance: a global multifaceted phenomenon. Pathog Glob Health. 2015;109(7):309–18.
  3. 3. Renwick MJ, Brogan DM, Mossialos E. A systematic review and critical assessment of incentive strategies for discovery and development of novel antibiotics. J Antibiot (Tokyo). 2016;69(2):73–88.
  4. 4. Tor Y, Fair R. Antibiotics and bacterial resistance in the 21st century. Perspect Med Chem. 2014;6:25–64.
  5. 5. Davies D. Understanding biofilm resistance to antibacterial agents. Nat Rev Drug Discov. 2003;2(2):114–22.
  6. 6. Worthington RJ, Richards JJ, Melander C. Small molecule control of bacterial biofilms. Org Biomol Chem. 2012;10(37):7457.
  7. 7. Bjarnsholt T. The role of bacterial biofilms in chronic infections. APMIS. 2013;121:1–58.
  8. 8. Michael CA, Dominey-Howes D, Labbate M. The antimicrobial resistance crisis: causes, consequences, and management. Front Public Health. 2014;2(145) 1–8.
  9. 9. Ventola CL. The antibiotic resistance crisis: part 1: causes and threats. Pharm Ther. 2015;40(4):277–83.
  10. 10. Luyt C-E, Bréchot N, Trouillet J-L, Chastre J. Antibiotic stewardship in the intensive care unit. Crit Care. 2014;18(480)1–12.
  11. 11. Katz ML, Mueller LV, Polyakov M, Weinstock SF. Where have all the antibiotic patents gone? Nat Biotechnol. 2006;24(12):1529–31.
  12. 12. Penesyan A, Gillings M, Paulsen I. Antibiotic discovery: combatting bacterial resistance in cells and in biofilm communities. Molecules. 2015;20(4):5286–98.
  13. 13. Levy SB. Factors impacting on the problem of antibiotic resistance. J Antimicrob Chemother. 2002;49(1):25–30.
  14. 14. Rogers SA, Huigens RW, Cavanagh J, Melander C. Synergistic effects between conventional antibiotics and 2-aminoimidazole-derived antibiofilm agents. Antimicrob Agents Chemother. 2010;54(5):2112–8.
  15. 15. Cheng G, Dai M, Ahmed S, Hao H, Wang X, Yuan Z. antimicrobial drugs in fighting against antimicrobial resistance. Front Microbiol. 2016;7(470):1–11.
  16. 16. Donlan RM, Costerton JW. Biofilms: survival mechanisms of clinically relevant microorganisms. Clin Microbiol Rev. 2002;15(2):167–93.
  17. 17. Stewart PS, William Costerton J. Antibiotic resistance of bacteria in biofilms. Lancet. 2001;358(9276):135–8.
  18. 18. Bordi C, de Bentzmann S. Hacking into bacterial biofilms: a new therapeutic challenge. Ann Intensive Care. 2011;1(1):19.
  19. 19. Stewart PS, Franklin MJ. Physiological heterogeneity in biofilms. Nat Rev Microbiol. 2008;6(3):199–210.
  20. 20. Landini P, Antoniani D, Burgess JG, Nijland R. Molecular mechanisms of compounds affecting bacterial biofilm formation and dispersal. Appl Microbiol Biotechnol. 2010;86(3):813–23.
  21. 21. Stewart PS. Mechanisms of antibiotic resistance in bacterial biofilms. Int J Med Microbiol. 2002;292(2):107–13.
  22. 22. Ceri H, Olson ME, Stremick C, Read RR, Morck D, Buret A. The Calgary Biofilm Device: new technology for rapid determination of antibiotic susceptibilities of bacterial biofilms. J Clin Microbiol. 1999;37(6):1771–6.
  23. 23. Piddock LJ. The crisis of no new antibiotics—what is the way forward? Lancet Infect Dis. 2012;12(3):249–53.
  24. 24. Silver LL. Challenges of antibacterial discovery. Clin Microbiol Rev. 2011;24(1):71–109.
  25. 25. Gao W, Thamphiwatana S, Angsantikul P, Zhang L. Nanoparticle approaches against bacterial infections: Nanoparticle against bacterial infections. Wiley Interdiscip Rev Nanomed Nanobiotechnol. 2014;6(6):532–47.
  26. 26. Rosenholm JM, Mamaeva V, Sahlgren C, Lindén M. Nanoparticles in targeted cancer therapy: mesoporous silica nanoparticles entering preclinical development stage. Nanomed. 2012;7(1):111–20.
  27. 27. Davis ME, Chen Z (Georgia), Shin DM. Nanoparticle therapeutics: an emerging treatment modality for cancer. Nat Rev Drug Discov. 2008;7(9):771–82.
  28. 28. Zhang L, Gu F, Chan J, Wang A, Langer R, Farokhzad O. Nanoparticles in medicine: therapeutic applications and developments. Clin Pharmacol Ther. 2008;83(5):761–9.
  29. 29. Nicolosi D, Scalia M, Nicolosi VM, Pignatello R. Encapsulation in fusogenic liposomes broadens the spectrum of action of vancomycin against Gram-negative bacteria. Int J Antimicrob Agents. 2010;35(6):553–8.
  30. 30. Watarai S, Iwase T, Tajima T, Yuba E, Kono K, Sekiya Y. Application of pH-sensitive fusogenicpolymer-modified liposomes for development of mucosal vaccines. Vet Immunol Immunopathol. 2014;158(1-2):62–72.
  31. 31. Hallaj-Nezhadi S, Hassan M. Nanoliposome-based antibacterial drug delivery. Drug Deliv. 2015;22(5):581–9.
  32. 32. Yang K, Gitter B, Rüger R, Wieland GD, Chen M, Liu X, et al. Antimicrobial peptide-modified liposomes for bacteria targeted delivery of temoporfin in photodynamic antimicrobial chemotherapy. Photochem Photobiol Sci. 2011;10(10):1593.
  33. 33. Boukherroub R. Functionalized nanomaterials for the management of microbial infection: a strategy to address microbial drug resistance. Boston, MA: Elsevier; 2017.
  34. 34. Michalak G, Głuszek K, Piktel E, Deptuła P, Puszkarz I, Niemirowicz K, et al. Polymeric nanoparticles – a novel solution for delivery of antimicrobial agents. Med Stud. 2016;1:56–62.
  35. 35. Scott C, Abdelghany, Quinn, Ingram, Gilmore, Donnelly, et al. Gentamicin-loaded nanoparticles show improved antimicrobial effects towards Pseudomonas aeruginosa infection. Int J Nanomedicine. 2012;4053.
  36. 36. Kavruk M, Celikbicak O, Ozalp VC, Borsa BA, Hernandez FJ, Bayramoglu G, et al. Antibiotic loaded nanocapsules functionalized with aptamer gates for targeted destruction of pathogens. Chem Commun. 2015;51(40):8492–5.
  37. 37. Barreras US, Méndez FT, Martínez REM, Valencia CS, Rodríguez PRM, Rodríguez JPL. Chitosan nanoparticles enhance the antibacterial activity of chlorhexidine in collagen membranes used for periapical guided tissue regeneration. Mater Sci Eng C. 2016;58:1182–7.
  38. 38. Gupta A, Landis RF, Rotello VM. Nanoparticle-based antimicrobials: surface functionality is critical. F1000Research. 2016;5:364.
  39. 39. Hetrick EM, Shin JH, Paul HS, Schoenfisch MH. Anti-biofilm efficacy of nitric oxide-releasing silica nanoparticles. Biomaterials. 2009;30(14):2782–9.
  40. 40. Tamanna T, Bulitta JB, Yu A. Controlling antibiotic release from mesoporous silica nano drug carriers via self-assembled polyelectrolyte coating. J Mater Sci Mater Med. 2015;26(2).
  41. 41. Lee B-Y, Li Z, Clemens DL, Dillon BJ, Hwang AA, Zink JI, et al. Redox-triggered release of moxifloxacin from mesoporous silica nanoparticles functionalized with disulfide snap-tops enhances efficacy against pneumonic tularemia in mice. Small. 2016;12(27):3690–702.
  42. 42. Li L, Wang H. Enzyme-coated mesoporous silica nanoparticles as efficient antibacterial agents in vivo. Adv Healthc Mater. 2013;2(10):1351–60.
  43. 43. Braun K, Pochert A, Lindén M, Davoudi M, Schmidtchen A, Nordström R, et al. Membrane interactions of mesoporous silica nanoparticles as carriers of antimicrobial peptides. J Colloid Interface Sci. 2016;475:161–70.
  44. 44. Tian Y, Qi J, Zhang W, Cai Q, Jiang X. Facile, One-pot synthesis, and antibacterial activity of mesoporous silica nanoparticles decorated with well-dispersed silver nanoparticles. ACS Appl Mater Interfaces. 2014;6(15):12038–45.
  45. 45. Şen Karaman D, Sarwar S, Desai D, Björk E. M, Odén M, Chakrabarti P, Rosenholm J.M, Chakraborti S. Shape engineering boosts antibacterial activity of chitosan coated mesoporous silica nanoparticle doped with silver: a mechanistic investigation. Journal of Materials Chemistry B. 2016, 4, 3292–3304.
  46. 46. Zhu X, Radovic-Moreno AF, Wu J, Langer R, Shi J. Nanomedicine in the management of microbial infection – Overview and perspectives. Nano Today. 2014;9(4):478–98.
  47. 47. Cui Y, Zhao Y, Tian Y, Zhang W, Lü X, Jiang X. The molecular mechanism of action of bactericidal gold nanoparticles on Escherichia coli. Biomaterials. 2012;33(7):2327–33.
  48. 48. Karlsson HL, Cronholm P, Hedberg Y, Tornberg M, De Battice L, Svedhem S, et al. Cell membrane damage and protein interaction induced by copper containing nanoparticles—importance of the metal release process. Toxicology. 2013;313(1):59–69.
  49. 49. Chatterjee AK, Chakraborty R, Basu T. Mechanism of antibacterial activity of copper nanoparticles. Nanotechnology. 2014;25(13):135101.
  50. 50. Zare B, Faramarzi MA, Sepehrizadeh Z, Shakibaie M, Rezaie S, Shahverdi AR. Biosynthesis and recovery of rod-shaped tellurium nanoparticles and their bactericidal activities. Mater Res Bull. 2012;47(11):3719–25.
  51. 51. Lin Z-H, Lee C-H, Chang H-Y, Chang H-T. Antibacterial activities of tellurium nanomaterials. Chem—Asian J. 2012;7(5):930–934.
  52. 52. Claudio C-R, Chellam S. Bismuth nanoparticles: antimicrobials of broad-spectrum, low cost and safety. In: Seifalian A, de Mel A, Kalaskaer DM, editors Nanomedicine. One Central Press; 2014. Chapter 17 p. 430–438.
  53. 53. Ansari MA, Khan HM, Khan AA, Sultan A, Azam A. Characterization of clinical strains of MSSA, MRSA and MRSE isolated from skin and soft tissue infections and the antibacterial activity of ZnO nanoparticles. World J Microbiol Biotechnol. 2012;28(4):1605–13.
  54. 54. Meghana S, Kabra P, Chakraborty S, Padmavathy N. Understanding the pathway of antibacterial activity of copper oxide nanoparticles. RSC Adv. 2015;5(16):12293–9.
  55. 55. Hans M, Erbe A, Mathews S, Chen Y, Solioz M, Mücklich F. Role of copper oxides in contact killing of bacteria. Langmuir. 2013;29(52):16160–6.
  56. 56. Schairer DO, Chouake JS, Nosanchuk JD, Friedman AJ. The potential of nitric oxide releasing therapies as antimicrobial agents. Virulence. 2012;3(3):271–9.
  57. 57. Allahverdiyev AM, Abamor ES, Bagirova M, Rafailovich M. Antimicrobial effects of TiO2 and Ag2O nanoparticles against drug-resistant bacteria and leishmania parasites. Future Microbiol. 2011;6(8):933–40.
  58. 58. Ansari MA, Khan HM, Khan AA, Pal R, Cameotra SS. Antibacterial potential of Al2O3 nanoparticles against multidrug resistance strains of Staphylococcus aureus isolated from skin exudates. J Nanoparticle Res. 2013;15(10):1970–1982.
  59. 59. Ismail RA, Sulaiman GM, Abdulrahman SA, Marzoog TR. Antibacterial activity of magnetic iron oxide nanoparticles synthesized by laser ablation in liquid. Mater Sci Eng C. 2015;53:286–97.
  60. 60. Pelletier DA, Suresh AK, Holton GA, McKeown CK, Wang W, Gu B, et al. Effects of engineered cerium oxide nanoparticles on bacterial growth and viability. Appl Environ Microbiol. 2010;76(24):7981–9.
  61. 61. Lemire JA, Harrison JJ, Turner RJ. Antimicrobial activity of metals: mechanisms, molecular targets and applications. Nat Rev Microbiol. 2013;11(6):371–84.
  62. 62. R K, Pv K. Dendrimeric biocides—a tool for effective antimicrobial therapy. J Nanomed Nanotechnol. 2016;07(02).
  63. 63. Beyth N, Yudovin-Farber I, Bahir R, Domb AJ, Weiss EI. Antibacterial activity of dental composites containing quaternary ammonium polyethylenimine nanoparticles against Streptococcus mutans. Biomaterials. 2006;27(21):3995–4002.
  64. 64. Beyth N, Houri-Haddad Y, Baraness-Hadar L, Yudovin-Farber I, Domb AJ, Weiss EI. Surface antimicrobial activity and biocompatibility of incorporated polyethylenimine nanoparticles. Biomaterials. 2008;29(31):4157–63.
  65. 65. Qi L, Xu Z, Jiang X, Hu C, Zou X. Preparation and antibacterial activity of chitosan nanoparticles. Carbohydr Res. 2004;339(16):2693–700.
  66. 66. Gazit E. Self-assembled peptide nanostructures: the design of molecular building blocks and their technological utilization. Chem Soc Rev. 2007;36(8):1263.
  67. 67. Lyon DY, Adams LK, Falkner JC, Alvarez PJJ. Antibacterial activity of fullerene water suspensions: effects of preparation method and particle size. Environ Sci Technol. 2006;40(14):4360–6.
  68. 68. Kang S, Pinault M, Pfefferle LD, Elimelech M. Single-walled carbon nanotubes exhibit strong antimicrobial activity. Langmuir. 2007;23(17):8670–3.
  69. 69. Choi J, Seo Y, Hwang J, Kim J, Jeong Y, Hwang M. Antibacterial activity and cytotoxicity of multi-walled carbon nanotubes decorated with silver nanoparticles. Int J Nanomedicine. 2014;4621.
  70. 70. Li Q, Mahendra S, Lyon DY, Brunet L, Liga MV, Li D, et al. Antimicrobial nanomaterials for water disinfection and microbial control: potential applications and implications. Water Res. 2008;42(18):4591–602.
  71. 71. Lam SJ, O’Brien-Simpson NM, Pantarat N, Sulistio A, Wong EHH, Chen Y-Y, et al. Combating multidrug-resistant Gram-negative bacteria with structurally nanoengineered antimicrobial peptide polymers. Nat Microbiol. 2016;1(11):16162.
  72. 72. Wu R, Zhang H, Pan J, Zhu H, Ma Y, Cui W, et al. Spatio-design of multidimensional prickly Zn-doped CuO nanoparticle for efficient bacterial killing. Adv Mater Interfaces. 2016;3(18):1600472.
  73. 73. Smekalova M, Aragon V, Panacek A, Prucek R, Zboril R, Kvitek L. Enhanced antibacterial effect of antibiotics in combination with silver nanoparticles against animal pathogens. Vet J. 2016;209:174–9.
  74. 74. Li P, Li J, Wu C, Wu Q, Li J. Synergistic antibacterial effects of β-lactam antibiotic combined with silver nanoparticles. Nanotechnology. 2005;16(9):1912–7.
  75. 75. Huh AJ, Kwon YJ. “Nanoantibiotics”: a new paradigm for treating infectious diseases using nanomaterials in the antibiotics resistant era. J Controlled Release. 2011;156(2):128–45.
  76. 76. Balouiri M, Sadiki M, Ibnsouda SK. Methods for in vitro evaluating antimicrobial activity: a review. J Pharm Anal. 2016;6(2):71–9.
  77. 77. Jacqueline C, Caillon J. Impact of bacterial biofilm on the treatment of prosthetic joint infections. J Antimicrob Chemother. 2014;69(suppl 1):i37–40.
  78. 78. Malone M, Goeres DM, Gosbell I, Vickery K, Jensen S, Stoodley P. Approaches to biofilm-associated infections: the need for standardized and relevant biofilm methods for clinical applications. Expert Rev Anti Infect Ther. 2017;15(2):147–56.
  79. 79. Lebeaux D, Chauhan A, Rendueles O, Beloin C. From in vitro to in vivo models of bacterial biofilm-related infections. Pathogens. 2013;2(2):288–356.
  80. 80. Govan JRW. Multidrug-resistant pulmonary infection in cystic fibrosis—what does “resistant” mean? J Med Microbiol. 2006;55(12):1615–7.
  81. 81. Andrews JM. Determination of minimum inhibitory concentrations. J Antimicrob Chemother. 2001;48(suppl 1):5–16.
  82. 82. Mermel LA, Allon M, Bouza E, Craven DE, Flynn P, O’Grady NP, et al. Clinical practice guidelines for the diagnosis and management of intravascular catheter‐related infection: 2009 update by the infectious diseases society of America. Clin Infect Dis. 2009;49(1):1–45.
  83. 83. Pankey GA, Sabath LD. Clinical relevance of bacteriostatic versus bactericidal mechanisms of action in the treatment of gram‐positive bacterial infections. Clin Infect Dis. 2004;38(6):864–70.
  84. 84. Weir E, Lawlor A, Whelan A, Regan F. The use of nanoparticles in anti-microbial materials and their characterization. The Analyst. 2008;133(7):835.
  85. 85. Nostro A, Procopio F, Pizzimenti FC, Cannatelli MA, Bisignano G, Marino A, et al. Effects of oregano, carvacrol and thymol on Staphylococcus aureus and Staphylococcus epidermidis biofilms. J Med Microbiol. 2007;56(4):519–23.
  86. 86. Macia MD, Rojo-Molinero E, Oliver A. Antimicrobial susceptibility testing in biofilm-growing bacteria. Clin Microbiol Infect. 2014;20(10):981–90.
  87. 87. LaPlante KL, Mermel LA. In vitro activity of daptomycin and vancomycin lock solutions on staphylococcal biofilms in a central venous catheter model. Nephrol Dial Transplant. 2007;22(8):2239–46.
  88. 88. Ali L, Khambaty F, Diachenko G. Investigating the suitability of the Calgary Biofilm Device for assessing the antimicrobial efficacy of new agents. Bioresour Technol. 2006;97(15):1887–93.
  89. 89. Coenye T, Nelis HJ. In vitro and in vivo model systems to study microbial biofilm formation. J Microbiol Methods. 2010;83(2):89–105.
  90. 90. Stepanovic S, Vukovic D, Dakic I, Savic B, Svabic-Vlahovic M. A modified microtiter-plate test for quantification of staphylococcal biofilm formation. J Microbiol Methods. 2000;40(2):175–9.
  91. 91. Peeters E, Nelis HJ, Coenye T. Comparison of multiple methods for quantification of microbial biofilms grown in microtiter plates. J Microbiol Methods. 2008;72(2):157–65.
  92. 92. Junker LM, Clardy J. High-throughput screens for small-molecule inhibitors of pseudomonas aeruginosa biofilm development. Antimicrob Agents Chemother. 2007;51(10):3582–90.
  93. 93. Pantanella F, Valenti P, Natalizi T. Analytical techniques to study microbial biofilm on abiotic surfaces: pros and cons of the main techniques currently in use. Ann Ig Med Prev E Comunità. 2013;25(1):31–42.
  94. 94. Pitts B, Hamilton MA, Zelver N, Stewart PS. A microtiter-plate screening method for biofilm disinfection and removal. J Microbiol Methods. 2003;54(2):269–76.
  95. 95. O’Toole GA. Microtiter dish biofilm formation assay. J Vis Exp. 2011;47:2437.
  96. 96. Van den Driessche F, Rigole P, Brackman G, Coenye T. Optimization of resazurin-based viability staining for quantification of microbial biofilms. J Microbiol Methods. 2014;98:31–4.
  97. 97. Mariscal A, Lopez-Gigosos RM, Carnero-Varo M, Fernandez-Crehuet J. Fluorescent assay based on resazurin for detection of activity of disinfectants against bacterial biofilm. Appl Microbiol Biotechnol. 2009;82(4):773–83.
  98. 98. Sandberg ME, Schellmann D, Brunhofer G, Erker T, Busygin I, Leino R, et al. Pros and cons of using resazurin staining for quantification of viable Staphylococcus aureus biofilms in a screening assay. J Microbiol Methods. 2009;78(1):104–6.
  99. 99. Hannig C, Follo M, Hellwig E, Al-Ahmad A. Visualization of adherent micro-organisms using different techniques. J Med Microbiol. 2010;59(1):1–7.
  100. 100. Manavathu EK, Vager DL, Vazquez JA. Development and antimicrobial susceptibility studies of in vitro monomicrobial and polymicrobial biofilm models with Aspergillus fumigatus and Pseudomonas aeruginosa. BMC Microbiol. 2014;14(1):53.
  101. 101. Viganor L, Galdino ACM, Nunes APF, Santos KRN, Branquinha MH, Devereux M, et al. Anti-Pseudomonas aeruginosa activity of 1,10-phenanthroline-based drugs against both planktonic- and biofilm-growing cells. J Antimicrob Chemother. 2016;71(1):128–34.
  102. 102. Webster TJ, Seil I. Antimicrobial applications of nanotechnology: methods and literature. Int J Nanomedicine. 2012;2767.
  103. 103. Manner S, Skogman M, Goeres D, Vuorela P, Fallarero A. Systematic exploration of natural and synthetic flavonoids for the inhibition of Staphylococcus aureus biofilms. Int J Mol Sci. 2013;14(10):19434–51.
  104. 104. Toté K, Berghe DV, Maes L, Cos P. A new colorimetric microtitre model for the detection of Staphylococcus aureus biofilms: new S. aureus biofilm model. Lett Appl Microbiol. 2007;46(2):249–54.
  105. 105. ISO/TS 16550:2014 Determination of silver nanoparticles potency by release of muramic acid from Staphylococcus aureus. British Standards Institution; 2014.
  106. 106. Kim JS, Kuk E, Yu KN, Kim J-H, Park SJ, Lee HJ, et al. Antimicrobial effects of silver nanoparticles. Nanomedicine Nanotechnol Biol Med. 2007;3(1):95–101.
  107. 107. Sondi I, Salopek-Sondi B. Silver nanoparticles as antimicrobial agent: a case study on E. coli as a model for Gram-negative bacteria. J Colloid Interface Sci. 2004;275(1):177–82.
  108. 108. Cremonini E, Zonaro E, Donini M, Lampis S, Boaretti M, Dusi S, et al. Biogenic selenium nanoparticles: characterization, antimicrobial activity and effects on human dendritic cells and fibroblasts. Microb Biotechnol. 2016;9(6):758–71.
  109. 109. Sodagar A, Akhavan A, Hashemi E, Arab S, Pourhajibagher M, Sodagar K, et al. Evaluation of the antibacterial activity of a conventional orthodontic composite containing silver/hydroxyapatite nanoparticles. Prog Orthod. 2016;17(1).
  110. 110. Díez-Martínez R, García-Fernández E, Manzano M, Martínez Á, Domenech M, Vallet-Regí M, et al. Auranofin-loaded nanoparticles as a new therapeutic tool to fight streptococcal infections. Sci Rep. 2016;6:19525.
  111. 111. Takahashi C, Akachi Y, Ogawa N, Moriguchi K, Asaka T, Tanemura M, et al. Morphological study of efficacy of clarithromycin-loaded nanocarriers for treatment of biofilm infection disease. Med Mol Morphol. 2016;50(1):9–16.
  112. 112. Barros J, Grenho L, Fontenente S, Manuel CM, Nunes OC, Melo LF, et al. Staphylococcus aureus and Escherichia coli dual-species biofilms on nanohydroxyapatite loaded with CHX or ZnO nanoparticles: DUAL-SPECIES BIOFILMS ON NANOHA-CHX OR NANOHA-ZNO. J Biomed Mater Res A. 2017;105(2):491–497.
  113. 113. Khan ST, Ahmad J, Ahamed M, Musarrat J, Al-Khedhairy AA. Zinc oxide and titanium dioxide nanoparticles induce oxidative stress, inhibit growth, and attenuate biofilm formation activity of Streptococcus mitis. JBIC J Biol Inorg Chem. 2016;21(3):295–303.
  114. 114. Krausz AE, Adler BL, Cabral V, Navati M, Doerner J, Charafeddine RA, et al. Curcumin-encapsulated nanoparticles as innovative antimicrobial and wound healing agent. Nanomed Nanotechnol Biol Med. 2015;11(1):195–206.
  115. 115. Wu J, Xu H, Tang W, Kopelman R, Philbert MA, Xi C. Eradication of bacteria in suspension and biofilms using methylene blue-loaded dynamic nanoplatforms. Antimicrob Agents Chemother. 2009;53(7):3042–3048.
  116. 116. Du J, Singh H, Yi T-H. Antibacterial, anti-biofilm and anticancer potentials of green synthesized silver nanoparticles using benzoin gum (Styrax benzoin) extract. Bioprocess Biosyst Eng. 2016;39(12):1923–1931.
  117. 117. Krishnamoorthy K, Manivannan G, Kim SJ, Jeyasubramanian K, Premanathan M. Antibacterial activity of MgO nanoparticles based on lipid peroxidation by oxygen vacancy. J Nanoparticle Res. 2012;14(9).
  118. 118. Flores CY, Miñán AG, Grillo CA, Salvarezza RC, Vericat C, Schilardi PL. Citrate-capped silver nanoparticles showing good bactericidal effect against both planktonic and sessile bacteria and a low cytotoxicity to osteoblastic cells. ACS Appl Mater Interfaces. 2013;5(8):3149–59.
  119. 119. Ramasamy M, Lee J-H, Lee J. Potent antimicrobial and antibiofilm activities of bacteriogenically synthesized gold-silver nanoparticles against pathogenic bacteria and their physiochemical characterizations. J Biomater Appl. 2016;31(3):366–78.
  120. 120. Liakos I, Grumezescu A, Holban A, Florin I, D’Autilia F, Carzino R, et al. Polylactic acid—lemongrass essential oil nanocapsules with antimicrobial properties. Pharmaceuticals. 2016;9(3):42.
  121. 121. Vijayakumar S, Vaseeharan B, Malaikozhundan B, Shobiya M. Laurus nobilis leaf extract mediated green synthesis of ZnO nanoparticles: characterization and biomedical applications. Biomed Pharmacother. 2016;84:1213–22.
  122. 122. Manju S, Malaikozhundan B, Vijayakumar S, Shanthi S, Jaishabanu A, Ekambaram P, et al. Antibacterial, antibiofilm and cytotoxic effects of Nigella sativa essential oil coated gold nanoparticles. Microb Pathog. 2016;91:129–35.
  123. 123. Konwar A, Kalita S, Kotoky J, Chowdhury D. Chitosan–iron oxide coated graphene oxide nanocomposite hydrogel: a robust and soft antimicrobial biofilm. ACS Appl Mater Interfaces. 2016;8(32):20625–20634.
  124. 124. Cihalova K, Chudobova D, Michalek P, Moulick A, Guran R, Kopel P, et al. Staphylococcus aureus and MRSA growth and biofilm formation after treatment with antibiotics and SeNPs. Int J Mol Sci. 2015;16(10):24656–24672.
  125. 125. Ramalingam V, Rajaram R, PremKumar C, Santhanam P, Dhinesh P, Vinothkumar S, et al. Biosynthesis of silver nanoparticles from deep sea bacterium Pseudomonas aeruginosa JQ989348 for antimicrobial, antibiofilm, and cytotoxic activity: biosynthesis of silver nanoparticles. J Basic Microbiol. 2014;54(9):928–36.
  126. 126. Kalishwaralal K, BarathManiKanth S, Pandian SRK, Deepak V, Gurunathan S. Silver nanoparticles impede the biofilm formation by Pseudomonas aeruginosa and Staphylococcus epidermidis. Colloids Surf B Biointerfaces. 2010;79(2):340–344.
  127. 127. Akhil K, Jayakumar J, Gayathri G, Khan SS. Effect of various capping agents on photocatalytic, antibacterial and antibiofilm activities of ZnO nano particles. J Photochem Photobiol B. 2016;160:32–42.
  128. 128. Sonkusre P, Singh Cameotra S. Biogenic selenium nanoparticles inhibit Staphylococcus aureus adherence on different surfaces. Colloids Surf B Biointerfaces. 2015;136:1051–1057.
  129. 129. Aliasghari A, Rabbani Khorasgani M, Vaezifar S, Rahimi F, Younesi H, Khoroushi M. Evaluation of antibacterial efficiency of chitosan and chitosan nanoparticles on cariogenic streptococci: an in vitro study. Iran J Microbiol. 2016;8(2):93–100.
  130. 130. Yeroslavsky G, Lavi R, Alishaev A, Rahimipour S. Sonochemically-produced metal-containing polydopamine nanoparticles and their antibacterial and antibiofilm activity. Langmuir. 2016;32(20):5201–12.
  131. 131. Ahmed A, Khan AK, Anwar A, Ali SA, Shah MR. Biofilm inhibitory effect of chlorhexidine conjugated gold nanoparticles against Klebsiella pneumoniae. Microb Pathog. 2016;98:50–56.
  132. 132. Shi S, Jia J, Guo X, Zhao Y, Chen D, Guo Y, et al. Reduced Staphylococcus aureus biofilm formation in the presence of chitosan-coated iron oxide nanoparticles. Int J Nanomed. 2016;11:6499–6506.
  133. 133. Fazly Bazzaz BS, Khameneh B, Zarei H, Golmohammadzadeh S. Antibacterial efficacy of rifampin loaded solid lipid nanoparticles against Staphylococcus epidermidis biofilm. Microb Pathog. 2016;93:137–144.
  134. 134. Giri K, Rivas Yepes L, Duncan B, Kolumam Parameswaran P, Yan B, Jiang Y, et al. Targeting bacterial biofilms via surface engineering of gold nanoparticles. RSC Adv. 2015;5(128):105551–105559.
  135. 135. Boda SK, Broda J, Schiefer F, Weber-Heynemann J, Hoss M, Simon U, et al. Cytotoxicity of ultrasmall gold nanoparticles on planktonic and biofilm encapsulated gram-positive Staphylococci. Small. 2015;11(26):3183–3193.
  136. 136. Skogman ME, Vuorela PM, Fallarero A. Combining biofilm matrix measurements with biomass and viability assays in susceptibility assessments of antimicrobials against Staphylococcus aureus biofilms. J Antibiot (Tokyo). 2012;65(9):453–459.
  137. 137. Wirth SM, Lowry GV, Tilton RD. Natural organic matter alters biofilm tolerance to silver nanoparticles and dissolved silver. Environ Sci Technol. 2012;46(22):12687–12696.
  138. 138. Mu H, Tang J, Liu Q, Sun C, Wang T, Duan J. Potent antibacterial nanoparticles against biofilm and intracellular bacteria. Sci Rep. 2016;6:18877.
  139. 139. Loo C-Y, Rohanizadeh R, Young PM, Traini D, Cavaliere R, Whitchurch CB, et al. Combination of silver nanoparticles and curcumin nanoparticles for enhanced anti-biofilm activities. J Agric Food Chem. 2016;64(12):2513–2522.
  140. 140. Niemirowicz K, Piktel E, Wilczewska A, Markiewicz K, Durnaś B, Wątek M, et al. Core-shell magnetic nanoparticles display synergistic antibacterial effects against Pseudomonas aeruginosa and Staphylococcus aureus when combined with cathelicidin LL-37 or selected ceragenins. Int J Nanomed. 2016;11:5443–55.
  141. 141. Ionescu AC, Brambilla E, Travan A, Marsich E, Donati I, Gobbi P, et al. Silver–polysaccharide antimicrobial nanocomposite coating for methacrylic surfaces reduces Streptococcus mutans biofilm formation in vitro. J Dent. 2015;43(12):1483–1490.
  142. 142. Gahlawat G, Shikha S, Chaddha BS, Chaudhuri SR, Mayilraj S, Choudhury AR. Microbial glycolipoprotein-capped silver nanoparticles as emerging antibacterial agents against cholera. Microb Cell Factories. 2016;15:25.
  143. 143. Suppi S, Kasemets K, Ivask A, Künnis-Beres K, Sihtmäe M, Kurvet I, et al. A novel method for comparison of biocidal properties of nanomaterials to bacteria, yeasts and algae. J Hazard Mater. 2015;286:75–84.
  144. 144. Kar S. Development of nano mullite based mesoporous silica biocer with incorporated bacteria for arsenic remediation. Ceram Silik. 2016;60(3):1–10.

Written By

Didem Şen Karaman, Suvi Manner, Adyary Fallarero and Jessica M. Rosenholm

Submitted: 31 October 2016 Reviewed: 24 February 2017 Published: 31 May 2017