Next Article in Journal
Therapeutic Candidates for Alzheimer’s Disease: Saponins
Next Article in Special Issue
Theoretical Analysis of Riboflavin Adsorption on Hexagonal Boron Nitride for Drug Delivery Applications: Unveiling the Influence of Point Defects
Previous Article in Journal
Environmental Air Pollutants Affecting Skin Functions with Systemic Implications
Previous Article in Special Issue
Features of DNA–Montmorillonite Binding Visualized by Atomic Force Microscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Toxicity Mechanisms of Copper Nanoparticles and Copper Surfaces on Bacterial Cells and Viruses

by
Javiera Ramos-Zúñiga
,
Nicolás Bruna
and
José M. Pérez-Donoso
*
BioNanotechnology and Microbiology Laboratory, Center for Bioinformatics and Integrative Biology (CBIB), Faculty of Life Sciences, Universidad Andres Bello, Santiago 8370186, Chile
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 10503; https://doi.org/10.3390/ijms241310503
Submission received: 24 April 2023 / Revised: 8 June 2023 / Accepted: 13 June 2023 / Published: 22 June 2023

Abstract

:
Copper is a metal historically used to prevent infections. One of the most relevant challenges in modern society are infectious disease outbreaks, where copper-based technologies can play a significant role. Currently, copper nanoparticles and surfaces are the most common antimicrobial copper-based technologies. Despite the widespread use of copper on nanoparticles and surfaces, the toxicity mechanism(s) explaining their unique antimicrobial properties are not entirely known. In general, toxicity effects described in bacteria and fungi involve the rupture of membranes, accumulation of ions inside the cell, protein inactivation, and DNA damage. A few studies have associated Cu-toxicity with ROS production and genetic material degradation in viruses. Therefore, understanding the mechanisms of the toxicity of copper nanoparticles and surfaces will contribute to developing and implementing efficient antimicrobial technologies to combat old and new infectious agents that can lead to disease outbreaks such as COVID-19. This review summarizes the current knowledge regarding the microbial toxicity of copper nanoparticles and surfaces and the gaps in this knowledge. In addition, we discuss potential applications derived from discovering new elements of copper toxicity, such as using different molecules or modifications to potentiate toxicity or antimicrobial specificity.

1. Copper Properties and Applications

Copper is a heavy metal with three oxidation states: metallic copper, cuprous ion, and cupric ion; the last two being stable [1]. This element is essential for life in all prokaryotic and eukaryotic organisms due to its structural and functional role as an enzymatic cofactor of proteins involved in metabolic pathways such as cell respiration, photosynthesis, ion transport, hormone production, growth, and cell development [2,3,4,5]. Cellular requirements of this metal vary depending on the species, environment, and metabolic state [5]. Despite being an essential element, high concentrations of copper are toxic for most organisms. Copper toxicity is mainly associated with the ability to generate hydroxyl radicals (ROS) through the Fenton and Haber–Weiss reactions, triggering damage to lipids, proteins, and DNA and affecting cell viability [4,6].
Copper is widely used worldwide due to its versatility and antimicrobial activity. Copper-based technologies are currently applied in construction, textiles, agriculture, and health, among others (Table 1). For example, copper ions are used as a photoluminescence activator due to their conductive capacity [7]. In agriculture, solutions based on copper compounds are used in fruit and vegetable crops to prevent bacterial and fungal infections [8,9]. In the livestock industry, copper has been used as a diet supplement to promote pigs’ growth and prevent bacterial diseases such as Escherichia coli-mediated mastitis [9,10,11]. The use of copper in hospital settings to prevent healthcare-associated infections (HAIs) by E. coli, methicillin-resistant Staphylococcus aureus (MRSA), Listeria monocytogenes, Clostridium difficile, and influenza A virus, among many other microorganisms, has been reported [12,13,14,15,16,17]. Copper has also been used as nanoparticles (NPs): Rai et al. (2018) discussed the different uses of copper nanoparticles for food preservation. It has been reported that using bimetal nanoparticles composed of Ag-NPs and Cu-NPs added to an agar film inhibits the growth of microorganisms in food packaging [2,3,4]. The use of copper and Cu-NPs in water treatment processes has also been described where copper solutions are used for their antimicrobial activity in water and copper-based nanoparticles as an adsorbent material [18,19].
This review focuses on copper as an antimicrobial and antiviral agent, with special interest in the toxicity mechanism of copper nanoparticles and surfaces. Due to the widespread application and investigation of copper in recent years, it is essential to deeply understand how this metal works in terms of their toxicity mechanisms and its effects on cells. This knowledge will contribute to potentiate the capabilities of this material as an antimicrobial. Based on this, the terms used to search and select the literature were copper toxicity, antimicrobial copper, antiviral copper, copper nanoparticles, copper nanoparticle toxicity mechanisms, copper surface mechanisms, copper surfaces, and contact killing. Based on these criteria, literature articles that discussed or proposed mechanisms of toxicity and articles reporting copper-based clinical applications were included in this review.
Table 1. Copper-based technologies.
Table 1. Copper-based technologies.
ApplicationReferences
Agriculture and Livestock
Fungicide, algicide, nutritional supplement[8,9,20,21,22]
Building
Electricity, wiring, plumbing, cooling, roofing[1,9]
Industrial
Water treatment, dye manufacturing, oil refining, wood preservation[8,9,23,24]
Health
Intrauterine devices, dental crowns, antimicrobial surfaces, face masks[12,25,26,27]
Others
Photoluminescence[7]
Textile[25,28]
Food Industry[9]
Carbon–copper Nanotubes[29]
Transformation of solar energy, batteries, and gas sensors[30,31]

2. Use of Copper as an Antimicrobial

Copper has been used as an antimicrobial agent throughout history. In ancient Egypt, copper was used for medical purposes to disinfect wounds and drinking water [32,33]. Subsequently, other civilizations adopted this idea; Greeks, Romans, and Aztecs used copper and derived compounds for treating burns, intestinal worms, and infections [32,34]. The use of this metal extended until the commercial development of antibiotics in 1945 [35]. However, because of the spread of antibiotic-resistant microorganisms, during the last few decades, the use of copper as an alternative to dealing with resistant microorganisms has emerged as a powerful strategy [35,36]. Copper is a flexible material used in different formats, such as on surfaces, as nanoparticles, and in solutions, to prevent the spread of pathogens. It is durable over time and does not require continuous maintenance or treatment to be effective, making it a long-term investment. Moreover, the constant use of copper has been considered safe for humans, as in the case of intrauterine devices, copper pipes, and dental amalgam alloys [37]. Studies of copper textiles showed no skin irritation or adverse reactions in animals. In 2008, the US Environmental Protection Agency (EPA) declared copper and all its alloys as the first effective metallic antimicrobial agent, spreading its application in different industries [36]. Currently, various industries use copper’s antimicrobial properties in two forms: solid (surfaces and nanoparticles) and ionic (salts). Table 2 describes some of the main applications of copper as an antimicrobial. The advantages of using copper as an antimicrobial agent are associated with the low dose required to eliminate the microorganisms and the infrequent evolution of resistance mechanisms [27].
In addition, copper can act effectively against persistent and dormant cells [24,38]. Compared to antibiotics, this metal damages cells by affecting multiple processes, including membrane damage, genotoxicity, ROS production, protein dysfunction, and nutrient assimilation [24].
The main objective of using copper on surfaces is to prevent the spread of infections from contaminated surfaces. This type of material, with optimal surface disinfection, reduces the presence of microorganisms [39]. Copper surfaces are highly effective in preventing the formation of bacterial biofilms. For example, their use in the poultry industry to prevent biofilm formation by Salmonella enteritidis has been recently reported [40]. Copper surfaces also prevent the transfer of genes involved in antibiotic resistance in different organisms because they eradicate the genetic material [39]. In this way, it is possible to prevent the horizontal transfer of antibiotic-resistance genes between microorganisms on surfaces [41]. Using copper as an antimicrobial is a promising alternative to fight antibiotic-resistant microorganisms such as methicillin-resistant Staphylococcus aureus (MRSA) or viruses such as SARS-CoV-2. Copper inactivates a broad spectrum of microorganisms such as bacteria, fungi, viral, and spores [42], and most clinical microorganisms present low copper-resistance levels [43]. Unlike antibiotics, copper affects various cellular or viral targets, causing membrane and capsid destabilization, ROS production, protein damage, enzyme inhibition, displacement of Fe–S centers, and genotoxicity [3,4,5,42,44,45,46]. In recent years, the use of copper in medical devices has increased, with 30% of the articles cited in Table 2 reporting the use of copper surfaces in hospitals or clinical scenarios. Copper surfaces reduce the microbial load and HAI infection rates within health facilities [47,48,49]. In turn, the use of copper has positive effects when adhered to titanium surfaces and in Ti/Cu alloys used in medical implants [50,51]. Although the price of pure copper is higher than other metals or alloys used in clinical settings, such as aluminum or bronze, which are commonly used in medical equipment, copper is a less-expensive metal than titanium, gold, silver, or stainless steel, which are also widely used in the clinical field [52]. The scientific evidence provided in this review indicates that copper has antimicrobial properties that may favor its application in hospitals or public spaces. However, more robust scientific evidence still needs to be developed to accelerate the widespread use of copper in these settings.
The use of copper-based nanoparticles (Cu-NPs)—CuO, Cu2O, Cu0, CuS, and CuI—with antimicrobial properties is also important because of the intrinsic characteristics of these materials at the nanometric scale: their surface area, mechanical, chemical, electrical, optical, and magnetic characteristics are different from the ionic form [53]. Some of the uses of copper as an antimicrobial in nanotechnology applications involve the incorporation of Cu-NPs in cotton fibers for medical dressings and impregnated supplies [25,54]. These are also used in the textile industry, for example, socks with Cu-NPs are manufactured to prevent infections by “athlete’s foot” fungus and diabetic foot [25,37]. During the last few years, the applications of Cu-NPs for preventing the spread of HAIs have increased, and they have been implemented in different hospital environments. Table 2 provides several articles describing the use of copper, where more than 50% of the articles correspond to clinical applications of copper-based NPs. These NPs have also been used in anti-folding catheters to prevent the formation of biofilms by pathogens or multi-resistant organisms [55,56,57], prostheses [58,59,60,61], air filters [34,57,58], composite films with chitosan and copper oxide nanoparticles, and starch hydrogels with copper nanoparticles for wound-dressing applications [62,63]. Furthermore, copper-based NPs have been used in dentistry in long-lasting adhesive restoration materials [64], adhesive resins [65,66], and implant coatings [67]. In addition, Cu-NPs incorporated into the cover and paints of waiting room chairs have contributed to reducing the bacterial load in hospitals [68]. Cu-NPs are also used in respiratory face masks against human influenza A (H1N1) [25]. This application is relevant in the context of the current pandemic to prevent the spread of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Recently, in 2021, the EPA registered copper surfaces for their use against coronaviruses (Environmental Protection Agency, 2021). As a result of the current situation, many studies of copper as an antiviral to prevent the spread of SARS-CoV-2 have been implemented. Among them, the use of copper surfaces in combination with nanoparticles to generate a synergistic effect against viruses has been recently reported [69,70]. Technologies based on Cu-NPs or copper solutions have also been used to improve the composition of personal protective equipment (PPE) [49,71,72].
Table 2. Examples of copper application as an antimicrobial.
Table 2. Examples of copper application as an antimicrobial.
ApplicationFunctionType of CopperReferences
Copper surfacesNorovirus inactivationCu-alloy surfaces[73]
Copper surfacesElimination of vancomycin-resistant EnterococciCu0 and Cu-alloy surfaces[39,74]
Antimicrobial surfaces in spaceflightPrevention of S. cohnii and E. coli infectionsCu0 and CuO/Cu2O [75]
Surfaces in pediatric intensive care unitPrevention of infections associated with hospitalsCu0 surfaces[55,76]
Copper-doped silica xerogelsAvoidance of skin infectionsCu ions[56]
Wound dressingsPrevention of S. aureus and K. pneumoniae infectionsCuO/Cu2O NPs[25]
Application of nanotechnology in textiles, latex gloves, and polymersPrevention of microorganisms in medical suppliesCuO/Cu2O NPs[8,25,30]
Antiviral filtersDialysis pumps, blood banks and air filtersCuO/Cu2O NPs[34,57,58]
Cotton fibersPrevention of bacteria, fungi and virusesCu ions and NPs[54]
Surfaces with nanoparticle inlayAntimicrobial activityCu0 NPs[31,59]
Office, kindergarten, retirement home, hospital facilitiesAntimicrobial activityCu/brass surfaces[60]
Face masksAntiviral activity against influenza ACuO/Cu2O NPs[25]
Dental materialsBacterial plaque preventionCuO/Cu2O NPs[64,65,66,67]
ProsthesesAntibacterial 3D filamentsCu-NPs[77]
Polyester surface-coating paintAntimicrobial activityCu-NPs[68]
Wound dressingAntibacterial activityBacterial celullose/Cu-NPs[78]
NanostructuresAntimicrobial–
Antibiofilm activity
Cu-NPs/Carbon nanotubes[79]
3D face masks, connector for ventilatorsAntiviral activity against SARS-CoV-2Cu-NPs[80]
Antibacterial nanomaterialsMultidrug-resistant bacterial treatmentCuS NPs/
Graphene oxide nanosheets
[81]
Antibacterial coatingsAntibacterial activity
Antibiofilm activity
CuS nanoparticles[82]
Poultry industryAntibiofilm formation by Salmonella enteritidisCu surface[40]
Nanostructured copper surface/Prototypes for air and water cleaningAntiviral activity against SARS-CoV-2Copper surface and copper nanoparticles[70]
Transparent surface coatingAntiviral activity against Influenza ACu-NPs–
Graphene
[71]
Spray coatingAntiviral activity against SARS-CoV-2Cu-NPs[72]
Nanocomposite surfacesInhibition of SARS-CoV-2 and HuNV virusesCu–Ag nanocomposite surfaces[69]
Textile coatingSARS-CoV-2 inactivationHybrid alginate–copper sulfate[49]
Photothermal treatmentInhibition of bacteria in infected skinCuS NPs[83]

3. Copper Toxicity Mechanisms

3.1. Copper-Based Nanoparticles

Interest in copper-based nanoparticles (Cu-NPs) has grown over the last few years because of their antimicrobial properties against bacteria, fungi, viruses, and algae [84]. To date, different Cu-NPs with antimicrobial properties have been described, including CuO/Cu2O, CuS, and Cu0 NPs. Concerning the former, Giannousi et al. (2014) described the degradation of plasmid DNA in bacterial strains using CuO/Cu2O NPs (Figure 1) [85]. This degradation occurs depending on the dose of nanoparticles used. Furthermore, they showed that the concentration of ions released was below that needed for the inhibition of bacterial growth (Minimum Inhibitory Concentration). Other authors have evaluated the antimicrobial activity of copper oxide nanoparticles and their size effect. Applerot et al. (2012) and Padil et al. (2013) reported that copper oxide nanoparticles of small size (~5 and ~4.8 nm, respectively) had better antimicrobial activity than nanoparticles of larger size (~7.8 nm) [30,86]. This phenomenon occurs due to the capacity of smaller nanoparticles to penetrate the cells (Figure 1) [87,88,89,90].
On the other hand, the antimicrobial effect of CuS nanoparticles has been scarcely reported, even though these NPs are much less toxic for microorganisms than copper oxide or Cu0 nanoparticles. ROS production mediated by the Cu ions released from the nanostructure is the main cause of the cellular effects of CuS NPs (Figure 1) [91,92]. However, the antimicrobial effect of these nanoparticles could also be due to their nanometric size, composition, and structure [93,94,95].
Regarding Cu0 NPs, several articles have contributed to elucidating their antimicrobial mechanism (Figure 1). The toxicity mechanism of these NPs involves the interaction, accumulation, and dissolution of the NPs in the membrane, changing their permeability. Consequently, the proton motive force of the membrane decreases [96]. The generation of ROS by ions released from the nanoparticles causes lipid peroxidation, protein oxidation, and DNA degradation [96,97]. In turn, the metal ions inside the cell damage the DNA and affect ATP production. Copper ions can interact with phosphate and -SH groups of proteins and DNA, causing denaturation and other structural perturbations [84,96,98,99]. These antecedents suggest that the toxicity of Cu0 NPs is mediated by the release of Cu+2 ions when metal nanoparticles are oxidized.
In general, it has been described that the toxicity of Cu-NPs is associated with the adhesion of the NPs to the cell through electrostatic interactions [100,101,102]. As mentioned above, adhesion to the membrane could involve depolarization and cell death (Figure 1) [96,103]. The interaction of Cu-NPs with the cell membrane and the formation of holes in the membrane that allow the entry of Cu-NPs inside cells (Figure 1) has been described [104]. However, despite the evaluated background, the specific toxicity mechanism for each type of copper NPs has not been fully explained.
The toxicity of Cu-NPs depends on many factors such as temperature, pH, concentration, number of bacterial cells, and intrinsic physicochemical properties of the nanostructures, such as shape, size, and surface composition [53,103,105]. The interaction of small-sized nanoparticles and plasma membranes has been described; this interaction could generate a large contact surface between the metal ions and the cell, generating cell damage [30,86,93,94,94,95]. Shape and surface charge also influence the cell contact area, thus increasing its efficiency [95,106]. This background leaves a window for future research that may contribute to elucidating the mechanism of toxicity of CuO, Cu-S, and Cu0 nanoparticles.
Figure 1. General mechanism of toxicity for copper-based nanoparticles. (1) Accumulation, dissolution, and entry of nanoparticles into the cell. (2) Depolarization and membrane rupture due to the accumulation of nanoparticles on the cell surface [24,107,108,109]. (3) Cell entry of the nanoparticles and release of Cu (II) ions because of NP destabilization. The release of ions causes the following effects: (4) ROS production, lipid peroxidation, and protein oxidation; (5) Decrease in ATP production; (6) Protein damage due to oxidation and iron replacement by copper in proteins with Fe–S centers; and (7) DNA damage and degradation [84,98,99,108,110].
Figure 1. General mechanism of toxicity for copper-based nanoparticles. (1) Accumulation, dissolution, and entry of nanoparticles into the cell. (2) Depolarization and membrane rupture due to the accumulation of nanoparticles on the cell surface [24,107,108,109]. (3) Cell entry of the nanoparticles and release of Cu (II) ions because of NP destabilization. The release of ions causes the following effects: (4) ROS production, lipid peroxidation, and protein oxidation; (5) Decrease in ATP production; (6) Protein damage due to oxidation and iron replacement by copper in proteins with Fe–S centers; and (7) DNA damage and degradation [84,98,99,108,110].
Ijms 24 10503 g001

3.2. Copper Surfaces

The toxicity mechanisms of metallic copper surfaces and copper alloys are not totally known. However, contact killing has been proposed as the main mechanism involved in the elimination of microorganisms present on surfaces [36]. Contact killing is a mechanism that explains the cellular damage associated with the direct contact of a microorganism with surfaces composed of metals, surfaces coated with biocidal substances, or antimicrobial substances [35,45,46,111,112]. This mechanism of toxicity involves direct physical interaction with the organism, has immediate action and long-term effects, and does not require repeated applications for more effectiveness.
In general, two methods have been regularly used to evaluate the toxicity of copper surfaces: “wet” and “dry” [5,111]. The “wet” application refers to the inoculation technique, where a suspension of bacteria is added to the copper coupons (Figure 2A). The “dry” method uses a small amount of liquid spread through a cotton cane (Figure 2B). The “wet” mechanism begins with the dissolution of copper ions from the surface, generating cellular damage resulting from the entry of copper ions [32,36,45,112]. As shown in Figure 2A, the contact-killing mechanism involves damage to the plasma membrane due to the stress caused by the dissolution of copper ions from the surface, the admission and accumulation of high concentrations of Cu+1 and Cu+2 inside the cell [3,6,32], and the intracellular production of hydroxyl radicals as a result of Fenton-like reactions between Cu+1 and Cu+2 in the presence of H2O2 [5,6,32,46]. Copper ions released from the surface enter the cell and compete for metal-binding sites in enzymes (Figure 2A), generating the inhibition of essential enzymes and the destruction of iron–sulfur centers. In this way, copper displaces iron by coupling to sulfur and inactivating enzymes [5,6,32,44]. Finally, the DNA structure is broken or degraded due to the ROS generated in the cell (Figure 2A) [5,32,36,113].
The effect of copper surfaces on bacterial respiration has also been reported on vancomycin-resistant Enterococcus species in a process associated with cytochrome inhibition by copper ions released from the surface [39]. Warnes et al. (2012) reported that in E. coli O157: H7 and Salmonella, the damage caused by hydroxyl radicals is not the primary toxicity mechanism of copper surfaces [114]. The authors mainly refer to the damage occurring on the outer membrane since it is the cellular structure that interacts the most with the surface. Concerning this hypothesis, some studies demonstrate the high affinity of lipopolysaccharides with divalent cations in the extracellular space [115]. This phenomenon could influence the ion concentration in the membrane and its consequent rupture.
On the other hand, in the “dry” method, the liquid evaporates immediately, allowing direct contact of the cells with the metal surface [32]. Microorganisms exposed to the dry-surface method present higher sensitivity to copper surfaces. The time of death is shorter in bacteria and yeasts exposed to dry surfaces [116,117,118]. To date, it is still unknown why the “dry” or dissolution-independent method has higher antimicrobial efficacy than the “wet” method, and it seems that the cellular toxicity mediated by dry surfaces could involve an independent mechanism of copper solubilization (Figure 2B). The mechanism involved in the “dry” conditions is unknown. It has been suggested that it could involve ROS production. Warnes et al. (2011) indicate that exposure to copper on “dry” surfaces could involve the accumulation of ions immediately upon contact with the surface. This effect decreases the membrane potential in the cell, causing its destabilization. Santo et al. (2011) determined that “dry” toxicity in E. coli causes membrane damage due to ion accumulation, this being the primary mechanism of the previously described genotoxicity [42]. In addition, the inhibitory effect of copper on LD-transpeptidases associated with peptidoglycan has been reported in E. coli and E. faecium [119]. Using quantitative proteomics, Nandakumar et al. (2011) identified some of the main changes in cells exposed to “dry” conditions. This study supports the hypothesis that the cell envelope is mainly affected by copper surfaces due to the increase in the expression of proteins associated with the shell and the polysaccharide biogenesis of the capsule [119,120]. The overexpression of this machinery could be related to the repair or response to the damage caused by copper. Regarding the genotoxicity associated with copper on surfaces, the presence of DNA-repair proteins has not been reported [41]. In addition, proteins with iron-sulfur centers, such as isopropyl malate dehydratase, have been described as being inactivated in the presence of copper. This phenomenon results from the displacement of iron–sulfur centers [44].
We have previously mentioned some mechanisms involved in the toxicity of copper surfaces. However, most of these processes are exclusively related to the dissolution of ions from the metal surface. In both conditions (wet and dry), microorganisms cannot grow on the surface and the inactivation time is short: minutes to hours in the case of “wet” conditions and minutes in “dry” conditions. Finally, the key steps that influence contact killing in microorganisms are the dissolution of copper on surfaces, the bacteria–metal contact, the surface structure, and the nobility of the metal alloys [42,45,112,121,122,123]. Among these, bacteria–metal contact is the most critical step. Mathews et al. demonstrated the role of the physical contact of bacteria with the metal surface by blocking the copper surface with an inert polymer. Their study showed a decrease in the toxicity of copper surfaces in the absence of direct contact [122].
Considering the current knowledge regarding copper surface toxicity and their increasing applications that impact millions of humans daily in public environments, it is necessary to understand the effect of copper surfaces on microbial communities, to determine the microorganisms most capable of surviving on them, and to elucidate if these surfaces generate a selection pressure favoring microorganisms of health interest. Santo et al. (2010) isolated microorganisms present in European copper-composite currencies [117]. The authors determined that Gram-positive bacteria are more predominant in these samples. It would be ideal to evaluate the population associated with these surfaces by genomic-based analysis to determine the influence of copper on microbial communities.

4. Antiviral Effect of Copper Nanoparticles and Surfaces

Viruses are highly susceptible to copper-mediated toxicity because viruses lack a metal response or repair mechanisms such as those that exist in microorganisms [34,58]. Table 3 summarizes studies on copper toxicity on surfaces, as nanoparticles, and in solutions, its effect on different viruses (enveloped or non-enveloped), and the associated mechanism. To date, no correlation between copper susceptibility and virus type has been determined. Virus envelope characteristics, family, and genetic material are not directly related to copper toxicity [57].

4.1. Copper Nanoparticles as an Antiviral Agent

Many articles evaluating the effect of different Cu-NPs—CuO, Cu2O, Cu0, CuS, and CuI—against viruses have been reported to date. Table 3 summarizes some examples of the effect of these Cu-NPs on enveloped and non-enveloped viruses. In this section of the review, we will focus on the mechanisms described for each type of Cu-NP.
To date, most reports describing the use of Cu-NPs as antiviral agents study the effect of copper oxide NPs (CuO/Cu2O) over coated and uncoated viruses. Tavakoli and Hashemzadeh (2020) described that the mechanisms associated with the antiviral activity of CuO nanoparticles against Herpes Simplex Virus (HSV-1) involve the interaction with the virus surface (Table 3). In this case, nanoparticles affect viral attachment and cell entry [124]. The same process occurs when copper oxide nanoparticles interact with the Hepatitis C virus (HCVpp), blocking its entry into the cells [125]. In other viruses such as respiratory syncytial virus (RSV), Cytomegalovirus (CMV), Adenovirus Type I (HAdV), and Rhinovirus 2 (HRV-2), CuO NPs cause a reduction in viral titers [57].
Antiviral activity has also been reported in CuS NPs (Table 3). In the HuNoV virus, CuS NPs could disrupt capsid proteins, causing protein degradation [126]. Recently, Guo et al. (2021) described the use of CuS NPs to prevent Hepatitis B virus infections (HBV) [127]. The authors synthesized ultra-small Cu2S nanoparticles that blocked HBV assembly after photoinduction. Their study showed that the size of the nanoparticles also influences the antiviral effect. Further studies will be necessary to evaluate future applications of this type of nanomaterials as an antiviral.
On the other hand, Fujimori et al. (2012) determined that the antiviral mechanism of copper(i) iodide (CuI) particles against influenza H5N1 virus involves the generation of hydroxyl radicals. This radical could degrade functional viral proteins, such as hemagglutinin (HA) and neuraminidase (NA). Furthermore, the oxidation of lipids from the virus coating was also reported [128].
In general, the toxicity mechanisms of copper-based NPs on viruses have been associated with damage to the viral genome, inhibition of replication, protein synthesis, and the assembly and release of virions [124,125,126,128,129]. Figure 3 shows the main effects of Cu-NPs on viruses. There are two possible mechanisms explaining toxicity of NPs on viruses. One is an unknown dissolution-independent mechanism where the NPs remain stable and “accumulate” high concentrations of copper that can “attack” the virus. The second mechanism involves NP destabilization, where high concentrations of copper ions are generated. These generate capsid degradation, protein inactivation, and viral genome damage [126]. The common factor described in all Cu-NPs corresponds to “virus neutralization” in the case of RSV, CMV, HadV, and HRV-2 [57]. Specifically, it has been reported that Cu-NPs can block the attachment and cell entry of the HCV virus [125]. In the HSV-1 and HuNoV viruses, Cu-NPs can inactivate viral proteins via oxidation and degradation of the viral genome (Figure 3). It will be essential to consider these antecedents to optimize the use of copper NPs as antiviral agents in future applications.

4.2. Antiviral Effect of Copper Surfaces

In viruses, the toxicity mechanism is not associated with metabolic processes. In this regard, surface mechanisms could be associated with virus immobilization and genetic material inactivation. It has also been related to blocking or destroying the viral receptors for host cells [130]. The antiviral mechanism of copper surfaces has been associated with the destruction of viral genetic material and damage to the capsid surface, which could be disintegrated by the effect of copper ions released from the metal surface (Figure 4) [36,73].
The mechanism of virus inactivation on copper surfaces is shown in Figure 4. First, this process involves surface adhesion, such as that previously described in bacteria. The toxicity of copper surfaces against viruses could be a dependent or independent effect of the dissolution of copper ions from the surface. The dependent mechanism is associated with capsid damage and the degradation of proteins and DNA produced by the release of high concentrations of copper ions from the surface [73,131]. Table 3 summarizes different antiviral applications of copper surfaces on enveloped and non-enveloped viruses and the mechanisms described for these. RNA damage was reported when the influenza A virus was exposed to copper surfaces [15]. It has been reported that copper exposition involves capsid damage, loss of spike proteins, and viral genome degradation [132,133]. RNA degradation and capsid damage was observed when the non-enveloped virus human norovirus (HuNoV) was exposed to dry surfaces and copper alloys [73,131] The mechanism of toxicity of copper surfaces is independent of the virus type, whether enveloped or non-enveloped. However, most studies are associated with enveloped viruses such as Influenza A, Human coronavirus 229E, and SARS-CoV-2 [15,73].
Figure 4. Toxicity mechanisms of copper surfaces against viruses. (1) The viral particle adheres to the copper surface, and (2) the virus is exposed to the copper surface. Two mechanisms have been proposed. Left: the virus is damaged by the surface independently of copper ion release; this mechanism is unknown. Right: the virus is damaged by copper ions released by the metal surface. (3) Exposure to copper ions causes capsid damage and protein inactivation associated with ROS. (4–5) Genetic material is exposed and degraded [73,131,133].
Figure 4. Toxicity mechanisms of copper surfaces against viruses. (1) The viral particle adheres to the copper surface, and (2) the virus is exposed to the copper surface. Two mechanisms have been proposed. Left: the virus is damaged by the surface independently of copper ion release; this mechanism is unknown. Right: the virus is damaged by copper ions released by the metal surface. (3) Exposure to copper ions causes capsid damage and protein inactivation associated with ROS. (4–5) Genetic material is exposed and degraded [73,131,133].
Ijms 24 10503 g004
Table 3. Studies of copper as an antiviral agent and the toxicity mechanisms proposed.
Table 3. Studies of copper as an antiviral agent and the toxicity mechanisms proposed.
VirusDiseaseNucleic Acid TypeFamilyCopper TypeProposed MechanismReferences
Enveloped
Influenza ARespiratory tract infectionSingle-stranded RNAOrthomyxoviridaeCu surfaceRNA damage[15]
CuI NPsViral protein degradation[128]
Hepatitis C virus (HCV)Hepatic cirrhosis and hepatocellular carcinomaomaSingle-stranded RNAFlaviviridaeCuO NPsBlockage of attachment and entry stages[125]
Hepatitis B virus (HBV)Acute and chronic viral hepatitisDouble-stranded DNAHepadnaviridaeCu2S NPs [47]
Herpes Simplex Virus (HSV-1)Labial lesions, encephalitis, and peripheral nervous system disordersDouble-stranded DNAHerpesviridaeCuO NPsInactivation via oxidation of viral proteins, degradation of viral genome[124]
Human Immunodeficiency Virus (HIV)Acquired immunodeficiency syndromeSingle-stranded RNARetroviridaeCu ionsHIV-1 protease neutralization[134]
CuO filtersHIV-1 protease inactivation, nucleic acid denaturation[58]
Cupric ionsNucleic acid degradation[135]
Respiratory Syncytial Virus (RSV)Respiratory illnessSingle-stranded RNAParamyxoviridaeCuO NPsViral neutralization[57]
Cytomegalovirus (CMV)MononucleosisDouble-stranded DNAHerpesviridaeCuO NPsViral neutralization[57]
Bronchitis VirusInflammation of the bronchial tubesSingle-stranded RNACoronaviridaeCu ionsVirus inactivation[136]
Junin VirusArgentine hemorrhagic feverSingle-stranded RNAArenaviridaeCopper ions + peroxideInactivation by ROS[135]
Human coronavirus 229E (HuCoV-229E)Respiratory illnessSingle-stranded RNACoronaviridaeCu Surfaces and AlloysCapsid damage, loss of surface spikes and viral genome inactivation[133]
Non-enveloped
Human norovirus (HuNoV)Acute gastroenteritisSingle-stranded RNACaliciviridaeAu/CuS NPsCapsid protein degradation, capsid damage[126]
Dry Copper Surfaces and AlloysRNA degradation and capsid damage[73,131]
Bacteriophages MS2 and F2-Single-stranded RNALeviviridaeCupric ionsPhage inactivation[137]
Bacteriophage R17-Single-stranded RNALeviviridaeCupric ionsDegradation of phage genome[138]
Bacteriophage Qß -Single-stranded RNALeviviridaeParticles Cu2O, CuO and AgROS, leached copper ions and surface contact[139]
PoliovirusPoliomyelitis, paralysisSingle-stranded RNAPicornaviridaeCu ionsCapsid proteins and RNA attack[140]
Cu ionsInactivation of viral RNA [141]
Adenovirus Type I (HAdV)KeratoconjunctivitisDouble-stranded DNAAdenoviridaeCuO NPsViral neutralization[57]
Rhinovirus 2 (HRV-2)Respiratory illnessSingle-stranded RNAPicornaviridaeCuO NPsViral neutralization[57]
NPs—nanoparticles; Cu0, CuO, Cu2O; Copper ions—CuSO4, CI—Copper(i) iodide.

4.3. Copper and SARS-CoV-2

In the current COVID-19 pandemic caused by the SARS-CoV-2 virus, using copper surfaces and nanoparticles has become an alternative approach to prevent the spread of the virus. SARS-CoV-2 infection is spread by drops exhaled from the nose or mouth of people infected with the virus. These droplets can be in the air or reach commonly used surfaces [142]. The use of copper in surfaces and nanoparticles has been proposed to prevent the spread of the virus. Copper surfaces have been used in hospitals and public spaces [60,143]. Other studies have modified copper surfaces to prevent infection and the spread of the virus by using a surface coating composed of copper particles and polyurethane [144]. Unlike plastic and metal surfaces, copper surfaces can inactivate the virus by 96–99% after 2 h of exposure [144,145,146]. In addition, copper-based nanoparticles have been used in face masks, PPE, and disinfectants [80,147,148,149]. The effect of disinfectants based on copper iodine (CuI) nanoparticles has been evaluated for the disinfection of PPE; however, the impact has been assessed only in liquid solutions [145].
Previously, Warnes et al. (2015) evaluated the effect of copper surfaces on the human coronaviruses causing severe acute respiratory syndrome (SARS) and the Middle East Respiratory Syndrome (MERS). The authors assessed the human coronavirus 229E on five joint-touch surfaces, including copper and copper alloy surfaces. Rapid inactivation of the coronavirus on copper surfaces depends on the composition of the surface. Inactivation was associated with copper ions released from the surface, ROS generation, and viral genome fragmentation [133]. Aboubakr et al. (2020) and Kampf et al. (2020) studied the persistence of coronaviruses over 2 h to 9 days in different environments, commonly touched surfaces, weather conditions, and biocidal agents [150,151]. Specifically, the authors evaluated surfaces such as steel, aluminum, metal, wood, paper, glass, ceramic, and copper and biocidal agents such as ethanol, 2-propanol, sodium hypochlorite, and hydrogen peroxide. The lowest persistence of SARS-CoV-2 occurs on copper, latex, and soft-porosity surfaces [150,151]. In 2020, Van Doremalen et al. determined the stability of SARS-CoV-1 and SARS-CoV-2 on surfaces and aerosols. Five environmental conditions were assessed (aerosols, stainless steel, plastic, copper, and cardboard), showing that SARS-CoV-2 was not viable after 4 h of copper exposure. The half-life on all surfaces studied was similar for both viruses [132]. The mechanism associated with the toxicity of copper surfaces or nanoparticles against viruses has not been fully defined. Based on the knowledge described above for other viruses, it is most likely that the production of ROS inactivates the viral genetic material.

5. Concluding Remarks

As has been seen throughout this review, the use of copper surfaces and nanomaterials as antimicrobials has strongly increased during the last few decades. Since these metallic materials are potential tools for combating future microbial agents, it is crucial to thoroughly explore the pathways involved in their toxicity. The toxic effects of copper are mainly associated with the dissolution of ions from the surface or the destabilization of nanoparticles. These ions cause the rupture and depolarization of the plasma membrane, the production of ROS, the inhibition of essential metabolic enzymes, genotoxicity, and cell death. Considering the literature, there are still many unanswered questions associated with the copper toxicity mechanism, a situation that represents an enormous challenge for the scientific community.
Regarding limitations, Bisht et al. (2022) suggest that one of the main problems limiting the development of new copper applications is that most trials are conducted using in vitro studies [152]; however, it is essential to validate copper applications in field studies before implementing these antimicrobial materials on a large scale. Many factors influence the costs of using copper on a large scale, such as manufacturing, durability, cleaning methods, obtaining certifications, surface maintenance, treatment or coatings, material, and replacement costs. Moreover, through the cleaning methods, copper may lose its aesthetic properties due to metal reactivity, and it is crucial to evaluate if cleaning methods affect its antimicrobial properties [153]. In this sense, it will be helpful for future studies to focus on the efficacy of applying these materials to enhance such properties by finding appropriate cleaning methods and complementing them with coating technologies, and even enhancing synergistic approaches to develop practical and durable surfaces that allow their application in hospitals and public areas [52]. Concerning the use of copper to control the spread of viruses such as SARS-CoV-2, Hewawaduge et al. (2021) reported that the viral load in CuS-impregnated masks was significantly reduced in in vitro assays [154]. These copper-based applications could be vital in reducing the spread of SARS-CoV-2 or other viruses.
In addition, it is necessary to understand how the use of copper surfaces and NPs affects microbial communities. For example, the effect of masks or clothing with copper fibers or the effect of NPs on the skin microbiota should be determined. On the other hand, due to their use in hospitals, it is relevant to determine what changes are produced by copper surfaces among the abundance of pathogens associated with HAIs. This point is appropriate considering that this metal could generate selection pressure among the organisms present in the surface community. In turn, it will be necessary to find elements that could enhance the antimicrobial effects of copper to make it more effective and specific to defined microbial genera/species, especially in the future, when new antimicrobial agents will be required in the face of multidrug-resistant organisms.
Another relevant point to consider is the effect of copper on human health and the environment. Copper NPs can be toxic in different organisms when used in inappropriate concentrations, causing inflammation in the pulmonary epithelium and accumulating in organs such as liver and kidney [155,156,157]. The indiscriminate and unregulated use of copper nanoparticles can be harmful to the environment, since the waste associated with the research and application of copper nanoparticle-related products ends up in landfills without any management [84,158]. Copper-rich wastes can pollute soils and water, negatively affecting the ecosystem. It is important for future research to address these limitations and to understand the mechanisms of copper toxicity, and then use that knowledge to improve NP properties and their interaction with the environment.
After analyzing the literature regarding Cu-based NPs, one limitation is observed that is related to the composition of the nanoparticles used in many studies. Fifty percent of the articles cited in Table 2 describe the use of nanoparticles as antimicrobials. However, 64.7% of these articles do not indicate the composition of the nanoparticles used (CuxS, CuO, CuxSx, or others), mainly due to problems associated with their characterization. This limitation affects the standardization and comparison between the cited articles. In future research, it will be essential to consider the characterization of nanoparticles in terms of composition, size, solubility, and stability. In turn, the methods used to determine the antimicrobial properties of copper NPs differ among the studies cited in Table 2. Inhibition halos, viable cell counts, DNA degradation studies, ROS production, lipid peroxidation, and membrane potential determination have been used [30,96]. In the case of viruses, different methodologies are regularly used to study the effect of NPs. Most studies evaluate the infective capacity of the virus in cell cultures exposed to copper-based nanoparticles; however, other studies evaluate the effect of NPs by mixing the viral particles, antigens, or proteins with the nanomaterials [15,47,73,125,128,131,133,134].
In conclusion, despite the enormous advance of the last few decades in understanding the effect of Cu surfaces and NPs on microorganisms, many unanswered questions about their antimicrobial effect remain. A total understanding of this process is essential, particularly considering the wide use of copper as an antimicrobial in different applications and the urgent necessity for developing new alternatives to deal with multidrug-resistant microorganisms.

Author Contributions

Investigation, J.R.-Z., N.B. and J.M.P.-D.; writing—original draft preparation, J.R.-Z., N.B. and J.M.P.-D.; writing—review and editing, J.R.-Z., N.B. and J.M.P.-D.; figure preparation J.R.-Z., supervision, project administration, and funding acquisition, J.M.P.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Erika Elcira Donoso Lopez and Fondecyt 1200870 (JMP-D).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

In loving memory of Claudio Vásquez Guzmán, an excellent friend, mentor, and scientist, but an even better human being. Thanks for all the adventures and for showing us the beauty of science and friendship.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Barceloux, D.G. Copper. J. Toxicol. Clin. Toxicol. 1999, 37, 217–230. [Google Scholar] [CrossRef]
  2. Flemming, C.A.; Trevors, J.T. Copper Toxicity and Chemistry in the Environment: A Review. Water Air Soil Pollut. 1989, 44, 143–158. [Google Scholar] [CrossRef]
  3. Hong, R.; Kang, T.Y.; Michels, C.A.; Gadura, N. Membrane Lipid Peroxidation in Copper Alloy-Mediated Contact Killing of Escherichia coli. Appl. Environ. Microbiol. 2012, 78, 1776–1784. [Google Scholar] [CrossRef] [Green Version]
  4. Puig, S.; Thiele, D.J. Molecular Mechanisms of Copper Uptake and Distribution. Curr. Opin. Chem. Biol. 2002, 6, 171–180. [Google Scholar] [CrossRef]
  5. Stevenson, J.; Barwinska-Sendra, A.; Tarrant, E.; Waldron, K.J. Mechanism of Action and Applications of the Antimicrobial Properties of Copper. Microb. Pathog. Strateg. Combat. Sci. Technol. Educ. 2013, 2, 468–479. [Google Scholar]
  6. Solioz, M. Copper and Bacteria: Evolution, Homeostasis and Toxicity; SpringerBriefs in Molecular Science; Springer International Publishing: Cham, Switzerland, 2018; ISBN 978-3-319-94438-8. [Google Scholar]
  7. Xu, C.X.; Sun, X.W.; Zhang, X.H.; Ke, L.; Chua, S.J. Photoluminescent Properties of Copper-Doped Zinc Oxide Nanowires. Nanotechnology 2004, 15, 856–861. [Google Scholar] [CrossRef]
  8. Borkow, G.; Gabbay, J. Copper as a Biocidal Tool. Curr. Med. Chem. 2005, 12, 2163–2175. [Google Scholar] [CrossRef] [Green Version]
  9. Elguindi, J.; Hao, X.; Lin, Y.; Alwathnani, H.A.; Wei, G.; Rensing, C. Advantages and Challenges of Increased Antimicrobial Copper Use and Copper Mining. Appl. Microbiol. Biotechnol. 2011, 91, 237–249. [Google Scholar] [CrossRef]
  10. Edmonds, M.S.; Izquierdo, O.A.; Baker, D.H. Feed Additive Studies with Newly Weaned Pigs: Efficacy of Supplemental Copper, Antibiotics and Organic Acids. J. Anim. Sci. 1985, 60, 462–469. [Google Scholar] [CrossRef] [Green Version]
  11. Scaletti, R.W.; Trammell, D.S.; Smith, B.A.; Harmon, R.J. Role of Dietary Copper in Enhancing Resistance to Escherichia coli Mastitis. J. Dairy Sci. 2003, 86, 1240–1249. [Google Scholar] [CrossRef] [Green Version]
  12. Casey, A.L.; Adams, D.; Karpanen, T.J.; Lambert, P.A.; Cookson, B.D.; Nightingale, P.; Miruszenko, L.; Shillam, R.; Christian, P.; Elliott, T.S.J. Role of Copper in Reducing Hospital Environment Contamination. J. Hosp. Infect. 2010, 74, 72–77. [Google Scholar] [CrossRef]
  13. Noyce, J.O.; Michels, H.; Keevil, C.W. Use of Copper Cast Alloys to Control Escherichia coli O157 Cross-Contamination during Food Processing. Appl. Environ. Microbiol. 2006, 72, 4239–4244. [Google Scholar] [CrossRef] [Green Version]
  14. Noyce, J.O.; Michels, H.; Keevil, C.W. Potential Use of Copper Surfaces to Reduce Survival of Epidemic Meticillin-Resistant Staphylococcus Aureus in the Healthcare Environment. J. Hosp. Infect. 2006, 63, 289–297. [Google Scholar] [CrossRef]
  15. Noyce, J.O.; Michels, H.; Keevil, C.W. Inactivation of Influenza A Virus on Copper versus Stainless Steel Surfaces. Appl. Environ. Microbiol. 2007, 73, 2748–2750. [Google Scholar] [CrossRef] [Green Version]
  16. Wheeldon, L.J.; Worthington, T.; Lambert, P.A.; Hilton, A.C.; Lowden, C.J.; Elliott, T.S.J. Antimicrobial Efficacy of Copper Surfaces against Spores and Vegetative Cells of Clostridium Difficile: The Germination Theory. J. Antimicrob. Chemother. 2008, 62, 522–525. [Google Scholar] [CrossRef] [Green Version]
  17. Wilks, S.A.; Michels, H.T.; Keevil, C.W. Survival of Listeria Monocytogenes Scott A on Metal Surfaces: Implications for Cross-Contamination. Int. J. Food Microbiol. 2006, 111, 93–98. [Google Scholar] [CrossRef]
  18. Kefeni, K.K.; Mamba, B.B.; Msagati, T.A.M. Application of Spinel Ferrite Nanoparticles in Water and Wastewater Treatment: A Review. Sep. Purif. Technol. 2017, 188, 399–422. [Google Scholar] [CrossRef]
  19. Blanc, D.S.; Carrara, P.; Zanetti, G.; Francioli, P. Water Disinfection with Ozone, Copper and Silver Ions, and Temperature Increase to Control Legionella: Seven Years of Experience in a University Teaching Hospital. J. Hosp. Infect. 2005, 60, 69–72. [Google Scholar] [CrossRef]
  20. Banik, S.; Pérez-de-Luque, A. In Vitro Effects of Copper Nanoparticles on Plant Pathogens, Beneficial Microbes and Crop Plants. Span. J. Agric. Res. 2017, 15, e1005. [Google Scholar] [CrossRef] [Green Version]
  21. Bishop, W.M.; Villalon, G.V.; Willis, B.E. Assessing Copper Adsorption, Internalization, and Desorption Following Algaecide Application to Control Lyngbya Wollei from Lake Gaston, NC/VA, USA. Water Air Soil Pollut. 2018, 229, 152. [Google Scholar] [CrossRef]
  22. Gopi, R.; Avasthe, R.K.; Kalita, H.; Yadav, A.; Das, S.K.; Rai, D. Eco-Friendly Management of Tomato Late Blight Using Botanicals, Bio-Control Agents, Compost Tea and Copper Fungicides. Indian J. Agric. Sci. 2020, 90, 35–39. [Google Scholar] [CrossRef]
  23. de Arquer, F.P.G.; Talapin, D.V.; Klimov, V.I.; Arakawa, Y.; Bayer, M.; Sargent, E.H. Semiconductor Quantum Dots: Technological Progress and Future Challenges. Science 2021, 373, eaaz8541. [Google Scholar] [CrossRef]
  24. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial Activity of Metals: Mechanisms, Molecular Targets and Applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef]
  25. Borkow, G.; Okon-Levy, N.; Gabbay, J. Copper Oxide Impregnated Wound Dressing: Biocidal and Safety Studies. Wounds 2010, 22, 301–310. [Google Scholar]
  26. Lucas, L.C.; Lemons, J.E. Biodegradation of Restorative Metallic Systems. Adv. Dent. Res. 1992, 6, 32–37. [Google Scholar] [CrossRef]
  27. Turner, R.J. Metal-Based Antimicrobial Strategies. Microb. Biotechnol. 2017, 10, 1062–1065. [Google Scholar] [CrossRef]
  28. McArthur, J.V.; Tuckfield, R.C.; Baker-Austin, C. Antimicrobial Textiles. In Antibiotic Resistance; Coates, A.R.M., Ed.; Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2012; Volume 211, pp. 135–152. ISBN 978-3-642-28950-7. [Google Scholar]
  29. Chai, G.; Sun, Y.; Sun, J.; Chen, Q. Mechanical Properties of Carbon Nanotube–Copper Nanocomposites. J. Micromech. Microeng. 2008, 18, 035013. [Google Scholar] [CrossRef]
  30. Applerot, G.; Lellouche, J.; Lipovsky, A.; Nitzan, Y.; Lubart, R.; Gedanken, A.; Banin, E. Understanding the Antibacterial Mechanism of CuO Nanoparticles: Revealing the Route of Induced Oxidative Stress. Small 2012, 8, 3326–3337. [Google Scholar] [CrossRef]
  31. Mahmoodi, S.; Elmi, A.; Nezhadi, S.H. Copper Nanoparticles as Antibacterial Agents. J. Mol. Pharm. Org. Process. Res. 2018, 6, 1. [Google Scholar] [CrossRef]
  32. Grass, G.; Rensing, C.; Solioz, M. Metallic Copper as an Antimicrobial Surface. Appl. Environ. Microbiol. 2011, 77, 1541–1547. [Google Scholar] [CrossRef] [Green Version]
  33. Hans, M.; Támara, J.C.; Mathews, S.; Bax, B.; Hegetschweiler, A.; Kautenburger, R.; Solioz, M.; Mücklich, F. Laser Cladding of Stainless Steel with a Copper–Silver Alloy to Generate Surfaces of High Antimicrobial Activity. Appl. Surf. Sci. 2014, 320, 195–199. [Google Scholar] [CrossRef]
  34. Borkow, G. Using Copper to Fight Microorganisms. CCB 2012, 6, 93–103. [Google Scholar] [CrossRef]
  35. Hans, M.; Erbe, A.; Mathews, S.; Chen, Y.; Solioz, M.; Mücklich, F. Role of Copper Oxides in Contact Killing of Bacteria. Langmuir 2013, 29, 16160–16166. [Google Scholar] [CrossRef]
  36. Vincent, M.; Duval, R.E.; Hartemann, P.; Engels-Deutsch, M. Contact Killing and Antimicrobial Properties of Copper. J. Appl. Microbiol. 2018, 124, 1032–1046. [Google Scholar] [CrossRef] [Green Version]
  37. Arendsen, L.P.; Thakar, R.; Sultan, A.H. The Use of Copper as an Antimicrobial Agent in Health Care, Including Obstetrics and Gynecology. Clin. Microbiol. Rev. 2019, 32, e00125-18. [Google Scholar] [CrossRef]
  38. San, K.; Long, J.; Michels, C.A.; Gadura, N. Antimicrobial Copper Alloy Surfaces Are Effective against Vegetative but Not Sporulated Cells of Gram-positive Bacillus subtilis. MicrobiologyOpen 2015, 4, 753–763. [Google Scholar] [CrossRef]
  39. Warnes, S.L.; Keevil, C.W. Mechanism of Copper Surface Toxicity in Vancomycin-Resistant Enterococci Following Wet or Dry Surface Contact. Appl. Environ. Microbiol. 2011, 77, 6049–6059. [Google Scholar] [CrossRef] [Green Version]
  40. Pontin, K.P.; Borges, K.A.; Furian, T.Q.; Carvalho, D.; Wilsmann, D.E.; Cardoso, H.R.P.; Alves, A.K.; Chitolina, G.Z.; Salle, C.T.P.; de Souza Moraes, H.L.; et al. Antimicrobial Activity of Copper Surfaces against Biofilm Formation by Salmonella Enteritidis and Its Potential Application in the Poultry Industry. Food Microbiol. 2021, 94, 103645. [Google Scholar] [CrossRef]
  41. Warnes, S.L.; Highmore, C.J.; Keevil, C.W. Horizontal Transfer of Antibiotic Resistance Genes on Abiotic Touch Surfaces: Implications for Public Health. mBio 2012, 3, e00489-12. [Google Scholar] [CrossRef] [Green Version]
  42. Santo, C.E.; Lam, E.W.; Elowsky, C.G.; Quaranta, D.; Domaille, D.W.; Chang, C.J.; Grass, G. Bacterial Killing by Dry Metallic Copper Surfaces. Appl. Environ. Microbiol. 2011, 77, 794–802. [Google Scholar] [CrossRef] [Green Version]
  43. Montero, D.A.; Arellano, C.; Pardo, M.; Vera, R.; Gálvez, R.; Cifuentes, M.; Berasain, M.A.; Gómez, M.; Ramírez, C.; Vidal, R.M. Antimicrobial Properties of a Novel Copper-Based Composite Coating with Potential for Use in Healthcare Facilities. Antimicrob. Resist. Infect. Control 2019, 8, 3. [Google Scholar] [CrossRef]
  44. Macomber, L.; Imlay, J.A. The Iron-Sulfur Clusters of Dehydratases Are Primary Intracellular Targets of Copper Toxicity. Proc. Natl. Acad. Sci. USA 2009, 106, 8344–8349. [Google Scholar] [CrossRef] [Green Version]
  45. Hans, M.; Mathews, S.; Mücklich, F.; Solioz, M. Physicochemical Properties of Copper Important for Its Antibacterial Activity and Development of a Unified Model. Biointerphases 2015, 11, 018902. [Google Scholar] [CrossRef]
  46. Pham, A.N.; Xing, G.; Miller, C.J.; Waite, T.D. Fenton-like Copper Redox Chemistry Revisited: Hydrogen Peroxide and Superoxide Mediation of Copper-Catalyzed Oxidant Production. J. Catal. 2013, 301, 54–64. [Google Scholar] [CrossRef]
  47. Guo, Z.; Huang, X.; Li, Z.; Shi, J.; Zhai, X.; Hu, X.; Zou, X. Employing CuInS2quantum Dots Modified with Vancomycin for Detecting: Staphylococcus Aureus and Iron(Iii). Anal. Methods 2021, 13, 1517–1526. [Google Scholar] [CrossRef]
  48. Wilks, S.A.; Michels, H.; Keevil, C.W. The Survival of Escherichia coli O157 on a Range of Metal Surfaces. Int. J. Food Microbiol. 2005, 105, 445–454. [Google Scholar] [CrossRef]
  49. Bataglioli, R.A.; Rocha Neto, J.B.M.; Calais, G.B.; Lopes, L.M.; Tsukamoto, J.; de Moraes, A.P.; Arns, C.W.; Beppu, M.M. Hybrid Alginate–Copper Sulfate Textile Coating for Coronavirus Inactivation. J. Am. Ceram. Soc. 2022, 105, 1748–1752. [Google Scholar] [CrossRef]
  50. Xu, W.; Hou, C.; Mao, Y.; Yang, L.; Tamaddon, M.; Zhang, J.; Qu, X.; Liu, C.; Su, B.; Lu, X. Characteristics of Novel Ti–10Mo-XCu Alloy by Powder Metallurgy for Potential Biomedical Implant Applications. Bioact. Mater. 2020, 5, 659–666. [Google Scholar] [CrossRef]
  51. Xiao, X.; Han, X.; Wang, L.-G.; Long, F.; Ma, X.-L.; Xu, C.-C.; Ma, X.-B.; Wang, C.-X.; Liu, Z.-Y. Anaerobically Photoreductive Degradation by CdS Nanocrystal: Biofabrication Process and Bioelectron-Driven Reaction Coupled with Shewanella Oneidensis MR-1. Biochem. Eng. J. 2020, 154, 107466. [Google Scholar] [CrossRef]
  52. Rubilar, O.; Rai, M.; Tortella, G.; Diez, M.C.; Seabra, A.B.; Durán, N. Biogenic Nanoparticles: Copper, Copper Oxides, Copper Sulphides, Complex Copper Nanostructures and Their Applications. Biotechnol. Lett. 2013, 35, 1365–1375. [Google Scholar] [CrossRef]
  53. Hajipour, M.J.; Fromm, K.M.; Akbar Ashkarran, A.; Jimenez de Aberasturi, D.; de Larramendi, I.R.; Rojo, T.; Serpooshan, V.; Parak, W.J.; Mahmoudi, M. Antibacterial Properties of Nanoparticles. Trends Biotechnol. 2012, 30, 499–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Grace, M.; Chand, N.; Bajpai, S.K. Copper Alginate-Cotton Cellulose (CACC) Fibers with Excellent Antibacterial Properties. J. Eng. Fibers Fabr. 2009, 4, 155892500900400. [Google Scholar] [CrossRef] [Green Version]
  55. Schmidt, M.G.; von Dessauer, B.; Benavente, C.; Benadof, D.; Cifuentes, P.; Elgueta, A.; Duran, C.; Navarrete, M.S. Copper Surfaces Are Associated with Significantly Lower Concentrations of Bacteria on Selected Surfaces within a Pediatric Intensive Care Unit. Am. J. Infect. Control 2016, 44, 203–209. [Google Scholar] [CrossRef] [Green Version]
  56. Wu, X.; Ye, L.; Liu, K.; Wang, W.; Wei, J.; Chen, F.; Liu, C. Antibacterial Properties of Mesoporous Copper-Doped Silica Xerogels. Biomed. Mater. 2009, 4, 045008. [Google Scholar] [CrossRef]
  57. Borkow, G.; Sidwell, R.W.; Smee, D.F.; Barnard, D.L.; Morrey, J.D.; Lara-Villegas, H.H.; Shemer-Avni, Y.; Gabbay, J. Neutralizing Viruses in Suspensions by Copper Oxide-Based Filters. Antimicrob. Agents Chemother. 2007, 51, 2605–2607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Borkow, G.; Lara, H.H.; Covington, C.Y.; Nyamathi, A.; Gabbay, J. Deactivation of Human Immunodeficiency Virus Type 1 in Medium by Copper Oxide-Containing Filters. Antimicrob. Agents Chemother. 2008, 52, 518–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ben-Sasson, M.; Zodrow, K.R.; Genggeng, Q.; Kang, Y.; Giannelis, E.P.; Elimelech, M. Surface Functionalization of Thin-Film Composite Membranes with Copper Nanoparticles for Antimicrobial Surface Properties. Environ. Sci. Technol. 2014, 48, 384–393. [Google Scholar] [CrossRef]
  60. Inkinen, J.; Mäkinen, R.; Keinänen-Toivola, M.M.; Nordström, K.; Ahonen, M. Copper as an Antibacterial Material in Different Facilities. Lett. Appl. Microbiol. 2017, 64, 19–26. [Google Scholar] [CrossRef]
  61. Wang, Z.; Shen, Y.; Haapasalo, M. Antibiofilm Peptides against Oral Biofilms. J. Oral Microbiol. 2017, 9, 1327308. [Google Scholar] [CrossRef] [Green Version]
  62. Jayaramudu, T.; Varaprasad, K.; Pyarasani, R.D.; Reddy, K.K.; Kumar, K.D.; Akbari-Fakhrabadi, A.; Mangalaraja, R.V.; Amalraj, J. Chitosan Capped Copper Oxide/Copper Nanoparticles Encapsulated Microbial Resistant Nanocomposite Films. Int. J. Biol. Macromol. 2019, 128, 499–508. [Google Scholar] [CrossRef]
  63. Villanueva, M.E.; Diez, A.M.d.R.; González, J.A.; Pérez, C.J.; Orrego, M.; Piehl, L.; Teves, S.; Copello, G.J. Antimicrobial Activity of Starch Hydrogel Incorporated with Copper Nanoparticles. ACS Appl. Mater. Interfaces 2016, 8, 16280–16288. [Google Scholar] [CrossRef]
  64. Renné, W.G.; Lindner, A.; Mennito, A.S.; Agee, K.A.; Pashley, D.H.; Willett, D.; Sentelle, D.; Defee, M.; Schmidt, M.; Sabatini, C. Antibacterial Properties of Copper Iodide-Doped Glass Ionomer-Based Materials and Effect of Copper Iodide Nanoparticles on Collagen Degradation. Clin. Oral Investig. 2017, 21, 369–379. [Google Scholar] [CrossRef]
  65. Argueta-Figueroa, L.; Scougall-Vilchis, R.J.; Morales-Luckie, R.A.; Olea-Mejía, O.F. An Evaluation of the Antibacterial Properties and Shear Bond Strength of Copper Nanoparticles as a Nanofiller in Orthodontic Adhesive. Aust. Orthod. J. 2015, 31, 42–48. [Google Scholar] [CrossRef]
  66. ALGhanem, A.; Fernandes, G.; Visser, M.; Dziak, R.; Renné, W.G.; Sabatini, C. Biocompatibility and Bond Degradation of Poly-Acrylic Acid Coated Copper Iodide-Adhesives. Dent. Mater. 2017, 33, e336–e347. [Google Scholar] [CrossRef]
  67. Gosau, M.; Haupt, M.; Thude, S.; Strowitzki, M.; Schminke, B.; Buergers, R. Antimicrobial Effect and Biocompatibility of Novel Metallic Nanocrystalline Implant Coatings. J. Biomed. Mater. Res. B Appl. Biomater. 2016, 104, 1571–1579. [Google Scholar] [CrossRef]
  68. Palza, H.; Nuñez, M.; Bastías, R.; Delgado, K. In Situ Antimicrobial Behavior of Materials with Copper-Based Additives in a Hospital Environment. Int. J. Antimicrob. Agents 2018, 51, 912–917. [Google Scholar] [CrossRef]
  69. Mosselhy, D.A.; Kareinen, L.; Kivistö, I.; Virtanen, J.; Loikkanen, E.; Ge, Y.; Maunula, L.; Sironen, T. Inhibition of SARS-CoV-2 Alpha Variant and Murine Noroviruses on Copper-Silver Nanocomposite Surfaces. Nanomaterials 2022, 12, 1037. [Google Scholar] [CrossRef] [PubMed]
  70. Yi, G.; Riduan, S.N.; Armugam, A.; Ong, J.T.; Hon, P.Y.; Abdad, M.Y.; Vasoo, S.; Ang, B.S.; Zhang, Y. Nanostructured Copper Surface Kills ESKAPE Pathogens and Viruses in Minutes. ChemMedChem 2021, 16, 3553–3558. [Google Scholar] [CrossRef] [PubMed]
  71. Das Jana, I.; Kumbhakar, P.; Banerjee, S.; Gowda, C.C.; Kedia, N.; Kuila, S.K.; Banerjee, S.; Das, N.C.; Das, A.K.; Manna, I.; et al. Copper Nanoparticle–Graphene Composite-Based Transparent Surface Coating with Antiviral Activity against Influenza Virus. ACS Appl. Nano Mater. 2021, 4, 352–362. [Google Scholar] [CrossRef]
  72. Foffa, I.; Losi, P.; Quaranta, P.; Cara, A.; Al Kayal, T.; D’Acunto, M.; Presciuttini, G.; Pistello, M.; Soldani, G. A Copper Nanoparticles-Based Polymeric Spray Coating: Nanoshield against Sars-Cov-2. J. Appl. Biomater. Funct. Mater. 2022, 20, 22808000221076330. [Google Scholar] [CrossRef] [PubMed]
  73. Warnes, S.L.; Summersgill, E.N.; Keevil, C.W. Inactivation of Murine Norovirus on a Range of Copper Alloy Surfaces Is Accompanied by Loss of Capsid Integrity. Appl. Environ. Microbiol. 2015, 81, 1085–1091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Prado, J.V.; Vidal, A.R.; Durán, T.C. Aplicación de la capacidad bactericida del cobre en la práctica médica. Rev. Méd. Chile 2012, 140, 1325–1332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Hahn, C.; Hans, M.; Hein, C.; Mancinelli, R.L.; Mücklich, F.; Wirth, R.; Rettberg, P.; Hellweg, C.E.; Moeller, R. Pure and Oxidized Copper Materials as Potential Antimicrobial Surfaces for Spaceflight Activities. Astrobiology 2017, 17, 1183–1191. [Google Scholar] [CrossRef] [PubMed]
  76. von Dessauer, B.; Navarrete, M.S.; Benadof, D.; Benavente, C.; Schmidt, M.G. Potential Effectiveness of Copper Surfaces in Reducing Health Care–Associated Infection Rates in a Pediatric Intensive and Intermediate Care Unit: A Nonrandomized Controlled Trial. Am. J. Infect. Control 2016, 44, e133–e139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zuniga, J.M. 3D Printed Antibacterial Prostheses. Appl. Sci. 2018, 8, 1651. [Google Scholar] [CrossRef] [Green Version]
  78. He, W.; Huang, X.; Zheng, Y.; Sun, Y.; Xie, Y.; Wang, Y.; Yue, L. In Situ Synthesis of Bacterial Cellulose/Copper Nanoparticles Composite Membranes with Long-Term Antibacterial Property. J. Biomater. Sci. Polym. Ed. 2018, 29, 2137–2153. [Google Scholar] [CrossRef]
  79. Seo, Y.; Hwang, J.; Lee, E.; Kim, Y.J.; Lee, K.; Park, C.; Choi, Y.; Jeon, H.; Choi, J. Engineering Copper Nanoparticles Synthesized on the Surface of Carbon Nanotubes for Anti-Microbial and Anti-Biofilm Applications. Nanoscale 2018, 10, 15529–15544. [Google Scholar] [CrossRef]
  80. Zuniga, J.M.; Cortes, A. The Role of Additive Manufacturing and Antimicrobial Polymers in the COVID-19 Pandemic. Expert Rev. Med. Devices 2020, 17, 477–481. [Google Scholar] [CrossRef]
  81. Wang, W.; Li, B.; Yang, H.; Lin, Z.; Chen, L.; Li, Z.; Ge, J.; Zhang, T.; Xia, H.; Li, L.; et al. Efficient Elimination of Multidrug-Resistant Bacteria Using Copper Sulfide Nanozymes Anchored to Graphene Oxide Nanosheets. Nano Res. 2020, 13, 2156–2164. [Google Scholar] [CrossRef]
  82. Gargioni, C.; Borzenkov, M.; D’Alfonso, L.; Sperandeo, P.; Polissi, A.; Cucca, L.; Dacarro, G.; Grisoli, P.; Pallavicini, P.; D’Agostino, A.; et al. Self-Assembled Monolayers of Copper Sulfide Nanoparticles on Glass as Antibacterial Coatings. Nanomaterials 2020, 10, 352. [Google Scholar] [CrossRef] [Green Version]
  83. Wang, Z.; Hou, Z.; Wang, P.; Chen, F.; Luo, X. CuS-PNIPAm Nanoparticles with the Ability to Initiatively Capture Bacteria for Photothermal Treatment of Infected Skin. Regen. Biomater. 2022, 9, rbac026. [Google Scholar] [CrossRef]
  84. Ingle, A.P.; Duran, N.; Rai, M. Bioactivity, Mechanism of Action, and Cytotoxicity of Copper-Based Nanoparticles: A Review. Appl. Microbiol. Biotechnol. 2014, 98, 1001–1009. [Google Scholar] [CrossRef]
  85. Giannousi, K.; Lafazanis, K.; Arvanitidis, J.; Pantazaki, A.; Dendrinou-Samara, C. Hydrothermal Synthesis of Copper Based Nanoparticles: Antimicrobial Screening and Interaction with DNA. J. Inorg. Biochem. 2014, 133, 24–32. [Google Scholar] [CrossRef]
  86. Thekkae Padil, V.V.; Černík, M. Green Synthesis of Copper Oxide Nanoparticles Using Gum Karaya as a Biotemplate and Their Antibacterial Application. Int. J. Nanomed. 2013, 8, 889–898. [Google Scholar] [CrossRef] [Green Version]
  87. Haldorai, Y.; Shim, J.-J. Novel Chitosan-TiO2 Nanohybrid: Preparation, Characterization, Antibacterial, and Photocatalytic Properties. Polym. Compos. 2014, 35, 327–333. [Google Scholar] [CrossRef]
  88. Lv, Q.; Zhang, B.; Xing, X.; Zhao, Y.; Cai, R.; Wang, W.; Gu, Q. Biosynthesis of Copper Nanoparticles Using Shewanella Loihica PV-4 with Antibacterial Activity: Novel Approach and Mechanisms Investigation. J. Hazard. Mater. 2018, 347, 141–149. [Google Scholar] [CrossRef]
  89. Meghana, S.; Kabra, P.; Chakraborty, S.; Padmavathy, N. Understanding the Pathway of Antibacterial Activity of Copper Oxide Nanoparticles. RSC Adv. 2015, 5, 12293–12299. [Google Scholar] [CrossRef]
  90. Zhang, S.; Wang, Y.; Song, H.; Lu, J.; Yuan, Z.; Guo, J. Copper Nanoparticles and Copper Ions Promote Horizontal Transfer of Plasmid-Mediated Multi-Antibiotic Resistance Genes across Bacterial Genera. Environ. Int. 2019, 129, 478–487. [Google Scholar] [CrossRef] [PubMed]
  91. Singaravelu, D.K.; Subramaniyan, S.B.; Tharunya, M.P.; Veerappan, A. Antimicrobial Lipid Capped Copper Sulfide Nanoparticles Display Enhanced Bactericidal Effect against Carbapenem-Resistant Acinetobacter Baumannii. Mater. Lett. 2022, 306, 130985. [Google Scholar] [CrossRef]
  92. Ahmed, K.B.A.; Subramaniyan, S.B.; Banu, S.F.; Nithyanand, P.; Veerappan, A. Jacalin-Copper Sulfide Nanoparticles Complex Enhance the Antibacterial Activity against Drug Resistant Bacteria via Cell Surface Glycan Recognition. Colloids Surf. B Biointerfaces 2018, 163, 209–217. [Google Scholar] [CrossRef] [PubMed]
  93. Midander, K.; Cronholm, P.; Karlsson, H.L.; Elihn, K.; Möller, L.; Leygraf, C.; Wallinder, I.O. Surface Characteristics, Copper Release, and Toxicity of Nano- and Micrometer-Sized Copper and Copper(II) Oxide Particles: A Cross-Disciplinary Study. Small 2009, 5, 389–399. [Google Scholar] [CrossRef] [PubMed]
  94. Morones, J.R.; Elechiguerra, J.L.; Camacho, A.; Holt, K.; Kouri, J.B.; Ramírez, J.T.; Yacaman, M.J. The Bactericidal Effect of Silver Nanoparticles. Nanotechnology 2005, 16, 2346–2353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Pal, S.; Tak, Y.K.; Song, J.M. Does the Antibacterial Activity of Silver Nanoparticles Depend on the Shape of the Nanoparticle? A Study of the Gram-Negative Bacterium Escherichia coli. Appl. Environ. Microbiol. 2007, 73, 1712–1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Chatterjee, A.K.; Chakraborty, R.; Basu, T. Mechanism of Antibacterial Activity of Copper Nanoparticles. Nanotechnology 2014, 25, 135101. [Google Scholar] [CrossRef] [PubMed]
  97. Raffi, M.; Mehrwan, S.; Bhatti, T.M.; Akhter, J.I.; Hameed, A.; Yawar, W.; ul Hasan, M.M. Investigations into the Antibacterial Behavior of Copper Nanoparticles against Escherichia coli. Ann. Microbiol. 2010, 60, 75–80. [Google Scholar] [CrossRef]
  98. Ruparelia, J.P.; Chatterjee, A.K.; Duttagupta, S.P.; Mukherji, S. Strain Specificity in Antimicrobial Activity of Silver and Copper Nanoparticles. Acta Biomater. 2008, 4, 707–716. [Google Scholar] [CrossRef]
  99. Wang, L.; Hu, C.; Shao, L. The Antimicrobial Activity of Nanoparticles: Present Situation and Prospects for the Future. Int. J. Nanomed. 2017, 12, 1227–1249. [Google Scholar] [CrossRef] [Green Version]
  100. Abbaszadegan, A.; Ghahramani, Y.; Gholami, A.; Hemmateenejad, B.; Dorostkar, S.; Nabavizadeh, M.; Sharghi, H. The Effect of Charge at the Surface of Silver Nanoparticles on Antimicrobial Activity against Gram-Positive and Gram-Negative Bacteria: A Preliminary Study. J. Nanomater. 2015, 2015, 53. [Google Scholar] [CrossRef] [Green Version]
  101. Gahlawat, G.; Choudhury, A.R. A Review on the Biosynthesis of Metal and Metal Salt Nanoparticles by Microbes. RSC Adv. 2019, 9, 12944–12967. [Google Scholar] [CrossRef] [Green Version]
  102. Gold, K.; Slay, B.; Knackstedt, M.; Gaharwar, A.K. Antimicrobial Activity of Metal and Metal-Oxide Based Nanoparticles. Adv. Therap. 2018, 1, 1700033. [Google Scholar] [CrossRef]
  103. Khezerlou, A.; Alizadeh-Sani, M.; Azizi-Lalabadi, M.; Ehsani, A. Nanoparticles and Their Antimicrobial Properties against Pathogens Including Bacteria, Fungi, Parasites and Viruses. Microb. Pathog. 2018, 123, 505–526. [Google Scholar] [CrossRef]
  104. Ninganagouda, S.; Rathod, V.; Singh, D.; Hiremath, J.; Singh, A.K.; Mathew, J.; ul-Haq, M. Growth Kinetics and Mechanistic Action of Reactive Oxygen Species Released by Silver Nanoparticles from Aspergillus niger on Escherichia coli. BioMed Res. Int. 2014, 2014, 753419. [Google Scholar] [CrossRef] [Green Version]
  105. Kaweeteerawat, C.; Chang, C.H.; Roy, K.R.; Liu, R.; Li, R.; Toso, D.; Fischer, H.; Ivask, A.; Ji, Z.; Zink, J.I.; et al. Cu Nanoparticles Have Different Impacts in Escherichia coli and Lactobacillus brevis than Their Microsized and Ionic Analogues. ACS Nano 2015, 9, 7215–7225. [Google Scholar] [CrossRef] [Green Version]
  106. El Badawy, A.M.; Silva, R.G.; Morris, B.; Scheckel, K.G.; Suidan, M.T.; Tolaymat, T.M. Surface Charge-Dependent Toxicity of Silver Nanoparticles. Environ. Sci. Technol. 2011, 45, 283–287. [Google Scholar] [CrossRef]
  107. Azam, A.; Ahmed; Oves; Khan; Habib; Memic, A. Antimicrobial Activity of Metal Oxide Nanoparticles against Gram-Positive and Gram-Negative Bacteria: A Comparative Study. IJN 2012, 7, 6003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chatterjee, A.K.; Sarkar, R.K.; Chattopadhyay, A.P.; Aich, P.; Chakraborty, R.; Basu, T. A Simple Robust Method for Synthesis of Metallic Copper Nanoparticles of High Antibacterial Potency against E. coli. Nanotechnology 2012, 23, 085103. [Google Scholar] [CrossRef]
  109. Jiang, S.; Lee, J.-H.; Kim, M.-G.; Myung, N.V.; Fredrickson, J.K.; Sadowsky, M.J.; Hur, H.-G. Biogenic Formation of As-S Nanotubes by Diverse Shewanella Strains. Appl. Environ. Microbiol. 2009, 75, 6896–6899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. 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, 177–182. [Google Scholar] [CrossRef]
  111. van de Lagemaat, M.; Grotenhuis, A.; van de Belt-Gritter, B.; Roest, S.; Loontjens, T.J.A.; Busscher, H.J.; van der Mei, H.C.; Ren, Y. Comparison of Methods to Evaluate Bacterial Contact-Killing Materials. Acta Biomater. 2017, 59, 139–147. [Google Scholar] [CrossRef] [PubMed]
  112. Molteni, C.; Abicht, H.K.; Solioz, M. Killing of Bacteria by Copper Surfaces Involves Dissolved Copper. Appl. Environ. Microbiol. 2010, 76, 4099–4101. [Google Scholar] [CrossRef] [Green Version]
  113. Warnes, S.L.; Green, S.M.; Michels, H.T.; Keevil, C.W. Biocidal Efficacy of Copper Alloys against Pathogenic Enterococci Involves Degradation of Genomic and Plasmid DNAs. Appl. Environ. Microbiol. 2010, 76, 5390–5401. [Google Scholar] [CrossRef] [Green Version]
  114. Warnes, S.L.; Caves, V.; Keevil, C.W. Mechanism of Copper Surface Toxicity in Escherichia coli O157:H7 and Salmonella Involves Immediate Membrane Depolarization Followed by Slower Rate of DNA Destruction Which Differs from That Observed for Gram-Positive Bacteria: Rapid Death of Gram-Negative Bacteria on Dry Copper Surfaces. Environ. Microbiol. 2012, 14, 1730–1743. [Google Scholar] [CrossRef] [PubMed]
  115. Ferris, F.G.; Beveridge, T.J.; Fyfe, W.S. Iron-Silica Crystallite Nucleation by Bacteria in a Geothermal Sediment. Nature 1986, 320, 609–611. [Google Scholar] [CrossRef]
  116. Dupont, C.L.; Grass, G.; Rensing, C. Copper Toxicity and the Origin of Bacterial Resistance—New Insights and Applications. Metallomics 2011, 3, 1109. [Google Scholar] [CrossRef]
  117. Santo, C.E.; Morais, P.V.; Grass, G. Isolation and Characterization of Bacteria Resistant to Metallic Copper Surfaces. Appl. Environ. Microbiol. 2010, 76, 1341–1348. [Google Scholar] [CrossRef] [Green Version]
  118. Quaranta, D.; Krans, T.; Santo, C.E.; Elowsky, C.G.; Domaille, D.W.; Chang, C.J.; Grass, G. Mechanisms of Contact-Mediated Killing of Yeast Cells on Dry Metallic Copper Surfaces. Appl. Environ. Microbiol. 2011, 77, 416–426. [Google Scholar] [CrossRef] [Green Version]
  119. Peters, K.; Pazos, M.; Edoo, Z.; Hugonnet, J.E.; Martorana, A.M.; Polissi, A.; VanNieuwenhze, M.S.; Arthur, M.; Vollmer, W. Copper inhibits peptidoglycan LD-transpeptidases suppressing β-lactam resistance due to bypass of penicillin-binding proteins. Proc. Natl. Acad. Sci. USA 2018, 115, 10786–10791. [Google Scholar] [CrossRef] [Green Version]
  120. Nandakumar, R.; Santo, C.E.; Madayiputhiya, N.; Grass, G. Quantitative Proteomic Profiling of the Escherichia coli Response to Metallic Copper Surfaces. Biometals 2011, 24, 429–444. [Google Scholar] [CrossRef]
  121. Luo, J.; Hein, C.; Mücklich, F.; Solioz, M. Killing of Bacteria by Copper, Cadmium, and Silver Surfaces Reveals Relevant Physicochemical Parameters. Biointerphases 2017, 12, 020301. [Google Scholar] [CrossRef] [Green Version]
  122. Mathews, S.; Hans, M.; Mücklich, F.; Solioz, M. Contact Killing of Bacteria on Copper Is Suppressed If Bacterial-Metal Contact Is Prevented and Is Induced on Iron by Copper Ions. Appl. Environ. Microbiol. 2013, 79, 2605–2611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Zeiger, M.; Solioz, M.; Edongué, H.; Arzt, E.; Schneider, A.S. Surface Structure Influences Contact Killing of Bacteria by Copper. MicrobiologyOpen 2014, 3, 327–332. [Google Scholar] [CrossRef]
  124. Tavakoli, A.; Hashemzadeh, M.S. Inhibition of Herpes Simplex Virus Type 1 by Copper Oxide Nanoparticles. J. Virol. Methods 2020, 275, 113688. [Google Scholar] [CrossRef]
  125. Hang, X.; Peng, H.; Song, H.; Qi, Z.; Miao, X.; Xu, W. Antiviral Activity of Cuprous Oxide Nanoparticles against Hepatitis C Virus in Vitro. J. Virol. Methods 2015, 222, 150–157. [Google Scholar] [CrossRef]
  126. Broglie, J.J.; Alston, B.; Yang, C.; Ma, L.; Adcock, A.F.; Chen, W.; Yang, L. Antiviral Activity of Gold/Copper Sulfide Core/Shell Nanoparticles against Human Norovirus Virus-Like Particles. PLoS ONE 2015, 10, e0141050. [Google Scholar] [CrossRef]
  127. Guo, X.; Sun, M.; Gao, R.; Qu, A.; Chen, C.; Xu, C.; Kuang, H.; Xu, L. Ultrasmall Copper (I) Sulfide Nanoparticles Prevent Hepatitis B Virus Infection. Angew. Chem. Int. Ed. 2021, 60, 13073–13080. [Google Scholar] [CrossRef]
  128. Fujimori, Y.; Sato, T.; Hayata, T.; Nagao, T.; Nakayama, M.; Nakayama, T.; Sugamata, R.; Suzuki, K. Novel Antiviral Characteristics of Nanosized Copper(I) Iodide Particles Showing Inactivation Activity against 2009 Pandemic H1N1 Influenza Virus. Appl. Environ. Microbiol. 2012, 78, 951–955. [Google Scholar] [CrossRef] [Green Version]
  129. Shionoiri, N.; Sato, T.; Fujimori, Y.; Nakayama, T.; Nemoto, M.; Matsunaga, T.; Tanaka, T. Investigation of the Antiviral Properties of Copper Iodide Nanoparticles against Feline Calicivirus. J. Biosci. Bioeng. 2012, 113, 580–586. [Google Scholar] [CrossRef]
  130. Thurman, R.B.; Gerba, C.P.; Bitton, G. The Molecular Mechanisms of Copper and Silver Ion Disinfection of Bacteria and Viruses. Crit. Rev. Environ. Control 1989, 18, 295–315. [Google Scholar] [CrossRef]
  131. Warnes, S.L.; Keevil, C.W. Inactivation of Norovirus on Dry Copper Alloy Surfaces. PLoS ONE 2013, 8, e75017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. van Doremalen, N.; Bushmaker, T.; Morris, D.H.; Holbrook, M.G.; Gamble, A.; Williamson, B.N.; Tamin, A.; Harcourt, J.L.; Thornburg, N.J.; Gerber, S.I.; et al. Aerosol and Surface Stability of SARS-CoV-2 as Compared with SARS-CoV-1. N. Engl. J. Med. 2020, 382, 1564–1567. [Google Scholar] [CrossRef] [PubMed]
  133. Warnes, S.L.; Little, Z.R.; Keevil, C.W. Human Coronavirus 229E Remains Infectious on Common Touch Surface Materials. mBio 2015, 6, e01697-15. [Google Scholar] [CrossRef] [Green Version]
  134. Karlström, A.R.; Levine, R.L. Copper Inhibits the Protease from Human Immunodeficiency Virus 1 by Both Cysteine-Dependent and Cysteine-Independent Mechanisms. Proc. Natl. Acad. Sci. USA 1991, 88, 5552–5556. [Google Scholar] [CrossRef] [Green Version]
  135. Sagripanti, J.L.; Routson, L.B.; Lytle, C.D. Virus Inactivation by Copper or Iron Ions Alone and in the Presence of Peroxide. Appl. Environ. Microbiol. 1993, 59, 4374–4376. [Google Scholar] [CrossRef] [Green Version]
  136. Jordan, F.T.; Nassar, T.J. The Influence of Copper on the Survival of Infectious Bronchitis Vaccine Virus in Water. Vet. Rec. 1971, 89, 609–610. [Google Scholar] [CrossRef] [PubMed]
  137. Yamamoto, N.; Hiatt, C.W.; Haller, W. Mechanism of Inactivation of Bacteriophages by Metals. Biochim. Biophys. Acta 1964, 91, 257–261. [Google Scholar] [CrossRef] [PubMed]
  138. Wong, K.; Morgan, A.R.; Paranchych, W. Controlled Cleavage of Phage R17 RNA within the Virion by Treatment with Ascorbate and Copper (II). Can. J. Biochem. 1974, 52, 950–958. [Google Scholar] [CrossRef] [PubMed]
  139. Sunada, K.; Minoshima, M.; Hashimoto, K. Highly Efficient Antiviral and Antibacterial Activities of Solid-State Cuprous Compounds. J. Hazard. Mater. 2012, 235–236, 265–270. [Google Scholar] [CrossRef]
  140. Yahya, M.T.; Straub, T.M.; Gerba, C.P. Inactivation of Coliphage MS-2 and Poliovirus by Copper, Silver, and Chlorine. Can. J. Microbiol. 1992, 38, 430–435. [Google Scholar] [CrossRef]
  141. Totsuka, A.; Ohtaki, K. The Effects of Amino Acids and Metals on the Infectivity of Poliovirus Ribonucleic Acid. Jpn. J. Microbiol. 1974, 18, 107–112. [Google Scholar] [CrossRef]
  142. Poggio, C.; Colombo, M.; Arciola, C.R.; Greggi, T.; Scribante, A.; Dagna, A. Copper-Alloy Surfaces and Cleaning Regimens against the Spread of SARS-CoV-2 in Dentistry and Orthopedics. From Fomites to Anti-Infective Nanocoatings. Materials 2020, 13, 3244. [Google Scholar] [CrossRef]
  143. Muller, M.P.; MacDougall, C.; Lim, M.; Ontario Agency for Health Protection and Promotion Public Health Ontario; Provincial Infectious Diseases Advisory Committee on Infection Prevention and Control. Provincial Infectious Diseases Advisory Committee on Infection Prevention and Control Antimicrobial Surfaces to Prevent Healthcare-Associated Infections: A Systematic Review. J. Hosp. Infect. 2016, 92, 7–13. [Google Scholar] [CrossRef] [PubMed]
  144. Behzadinasab, S.; Chin, A.; Hosseini, M.; Poon, L.; Ducker, W.A. A Surface Coating That Rapidly Inactivates SARS-CoV-2. ACS Appl. Mater. Interfaces 2020, 12, 34723–34727. [Google Scholar] [CrossRef]
  145. Mantlo, E.; Rhodes, T.; Boutros, J.; Patterson-Fortin, L.; Evans, A.; Paessler, S. In Vitro Efficacy of a Copper Iodine Complex PPE Disinfectant for SARS-CoV-2 Inactivation. F1000Res 2020, 9, 674. [Google Scholar] [CrossRef] [PubMed]
  146. Paital, B.; Das, K.; Parida, S.K. Inter Nation Social Lockdown versus Medical Care against COVID-19, a Mild Environmental Insight with Special Reference to India. Sci. Total Environ. 2020, 728, 138914. [Google Scholar] [CrossRef] [PubMed]
  147. Cortes, A.A.; Zuñiga, J.M. The Use of Copper to Help Prevent Transmission of SARS-Coronavirus and Influenza Viruses. A General Review. Diagn. Microbiol. Infect. Dis. 2020, 98, 115176. [Google Scholar] [CrossRef]
  148. Colorado, H.A.; Mendoza, D.E.; Lin, H.-T.; Gutierrez-Velasquez, E. Additive Manufacturing against the Covid-19 Pandemic: A Technological Model for the Adaptability and Networking. J. Mater. Res. Technol. 2022, 16, 1150–1164. [Google Scholar] [CrossRef]
  149. Sportelli, M.C.; Izzi, M.; Kukushkina, E.A.; Hossain, S.I.; Picca, R.A.; Ditaranto, N.; Cioffi, N. Can Nanotechnology and Materials Science Help the Fight against SARS-CoV-2? Nanomaterials 2020, 10, 802. [Google Scholar] [CrossRef] [Green Version]
  150. Aboubakr, H.A.; Sharafeldin, T.A.; Goyal, S.M. Stability of SARS-CoV-2 and Other Coronaviruses in the Environment and on Common Touch Surfaces and the Influence of Climatic Conditions: A Review. Transbound. Emerg. Dis. 2021, 68, 296–312. [Google Scholar] [CrossRef]
  151. Kampf, G.; Todt, D.; Pfaender, S.; Steinmann, E. Persistence of Coronaviruses on Inanimate Surfaces and Their Inactivation with Biocidal Agents. J. Hosp. Infect. 2020, 104, 246–251. [Google Scholar] [CrossRef] [Green Version]
  152. Bisht, N.; Dwivedi, N.; Kumar, P.; Venkatesh, M.; Yadav, A.K.; Mishra, D.; Solanki, P.; Verma, N.K.; Lakshminarayanan, R.; Ramakrishna, S.; et al. Recent Advances in Copper and Copper-Derived Materials for Antimicrobial Resistance and Infection Control. Curr. Opin. Biomed. Eng. 2022, 24, 100408. [Google Scholar] [CrossRef]
  153. Airey, P.; Verran, J. Potential Use of Copper as a Hygienic Surface; Problems Associated with Cumulative Soiling and Cleaning. J. Hosp. Infect. 2007, 67, 271–277. [Google Scholar] [CrossRef] [PubMed]
  154. Hewawaduge, C.; Senevirathne, A.; Jawalagatti, V.; Kim, J.W.; Lee, J.H. Copper-Impregnated Three-Layer Mask Efficiently Inactivates SARS-CoV2. Environ. Res. 2021, 196, 110947. [Google Scholar] [CrossRef] [PubMed]
  155. Malhotra, N.; Ger, T.-R.; Uapipatanakul, B.; Huang, J.-C.; Chen, K.H.-C.; Hsiao, C.-D. Review of Copper and Copper Nanoparticle Toxicity in Fish. Nanomaterials 2020, 10, 1126. [Google Scholar] [CrossRef] [PubMed]
  156. Reshma, V.G.; Mohanan, P.V. Quantum Dots: Applications and Safety Consequences. J. Lumin. 2019, 205, 287–298. [Google Scholar] [CrossRef]
  157. Tulinska, J.; Mikusova, M.L.; Liskova, A.; Busova, M.; Masanova, V.; Uhnakova, I.; Rollerova, E.; Alacova, R.; Krivosikova, Z.; Wsolova, L.; et al. Copper Oxide Nanoparticles Stimulate the Immune Response and Decrease Antioxidant Defense in Mice after Six-Week Inhalation. Front. Immunol. 2022, 13, 874253. [Google Scholar] [CrossRef]
  158. Wu, F.; Harper, B.J.; Crandon, L.E.; Harper, S.L. Assessment of Cu and CuO Nanoparticle Ecological Responses Using Laboratory Small-Scale Microcosms. Environ. Sci. Nano 2020, 7, 105–115. [Google Scholar] [CrossRef] [PubMed]
Figure 2. General mechanisms of contact killing on surfaces. (A) The “wet” method depends on copper dissolution: (1) Dissolution of copper ions from the metal surface [112]. (2) Cell membrane rupture due to the accumulation of ions in the membrane. (3) Copper entry and accumulation inside the cell. (4) ROS production due to the accumulation of ions within the cell. (5) Inhibition of essential enzymes by competition for divalent cation-binding sites used as cofactors [44]. (6) Damage and degradation of genomic and plasmid DNA because of the accumulation of ions and free radicals [3,5,6,32,45,112]. (B) “Dry” method independent of copper dissolution: (1) The toxic effect of the dry surface without ion release is unknown. Regarding this, (2) the effect of the surface due to the quick and concentrated dissolution of ions during the inoculation is postulated [39]. (3) ROS production by the accumulation of ions. (4) Membrane destabilization [42], (5) genotoxicity, and (6) displacement of Fe–S centers in proteins [44].
Figure 2. General mechanisms of contact killing on surfaces. (A) The “wet” method depends on copper dissolution: (1) Dissolution of copper ions from the metal surface [112]. (2) Cell membrane rupture due to the accumulation of ions in the membrane. (3) Copper entry and accumulation inside the cell. (4) ROS production due to the accumulation of ions within the cell. (5) Inhibition of essential enzymes by competition for divalent cation-binding sites used as cofactors [44]. (6) Damage and degradation of genomic and plasmid DNA because of the accumulation of ions and free radicals [3,5,6,32,45,112]. (B) “Dry” method independent of copper dissolution: (1) The toxic effect of the dry surface without ion release is unknown. Regarding this, (2) the effect of the surface due to the quick and concentrated dissolution of ions during the inoculation is postulated [39]. (3) ROS production by the accumulation of ions. (4) Membrane destabilization [42], (5) genotoxicity, and (6) displacement of Fe–S centers in proteins [44].
Ijms 24 10503 g002
Figure 3. General toxicity mechanisms of copper-based nanoparticles against viruses. (1) The virus is exposed to copper-based NPs. (2) In some cases, NPs could block the virus attachment to cells by blocking or inactivating proteins involved in this process [125]. (3) The virus interacts with copper-based NPs; this mechanism is unknown, but it is suggested that NPs could destabilize the capsid membrane (left). Conversely, copper-based NPs release copper ions (right). In both cases, the virus is exposed to a high concentration of localized copper, which causes (4) damage to the capsid, inactivation of proteins, and (5) degradation of the genetic material [126].
Figure 3. General toxicity mechanisms of copper-based nanoparticles against viruses. (1) The virus is exposed to copper-based NPs. (2) In some cases, NPs could block the virus attachment to cells by blocking or inactivating proteins involved in this process [125]. (3) The virus interacts with copper-based NPs; this mechanism is unknown, but it is suggested that NPs could destabilize the capsid membrane (left). Conversely, copper-based NPs release copper ions (right). In both cases, the virus is exposed to a high concentration of localized copper, which causes (4) damage to the capsid, inactivation of proteins, and (5) degradation of the genetic material [126].
Ijms 24 10503 g003
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

Ramos-Zúñiga, J.; Bruna, N.; Pérez-Donoso, J.M. Toxicity Mechanisms of Copper Nanoparticles and Copper Surfaces on Bacterial Cells and Viruses. Int. J. Mol. Sci. 2023, 24, 10503. https://doi.org/10.3390/ijms241310503

AMA Style

Ramos-Zúñiga J, Bruna N, Pérez-Donoso JM. Toxicity Mechanisms of Copper Nanoparticles and Copper Surfaces on Bacterial Cells and Viruses. International Journal of Molecular Sciences. 2023; 24(13):10503. https://doi.org/10.3390/ijms241310503

Chicago/Turabian Style

Ramos-Zúñiga, Javiera, Nicolás Bruna, and José M. Pérez-Donoso. 2023. "Toxicity Mechanisms of Copper Nanoparticles and Copper Surfaces on Bacterial Cells and Viruses" International Journal of Molecular Sciences 24, no. 13: 10503. https://doi.org/10.3390/ijms241310503

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