Review Open Access
Copyright ©2012 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Clin Infect Dis. Aug 25, 2012; 2(4): 77-90
Published online Aug 25, 2012. doi: 10.5495/wjcid.v2.i4.77
Fighting nosocomial infections with biocidal non-intrusive hard and soft surfaces
Gadi Borkow, Cupron Scientific, Hasadnaot 10, Herzelia 46733, Israel
Alastair Monk, Cupron Inc., 800 East Leigh Street, Richmond, VA 23219, United States
Author contributions: Borkow G wrote the main draft of the manuscript; Monk A assisted in finalizing the review.
Correspondence to: Dr. Gadi Borkow, Cupron Scientific, Hasadnaot 10, Herzelia 46733, Israel. gadi@cupron.com
Telephone: +972-546-611287 Fax: +972-9-8652115
Received: September 6, 2011
Revised: June 5, 2012
Accepted: July 4, 2012
Published online: August 25, 2012

Abstract

Approximately 7 million people worldwide acquire a healthcare associated infection each year. Despite aggressive monitoring, hand washing campaigns and other infection control measures, nosocomial infections (NI) rates, especially those caused by antibiotic resistant pathogens, are unacceptably high worldwide. Additional ways to fight these infections need to be developed. A potential overlooked and neglected source of nosocomial pathogens are those found in non-intrusive soft and hard surfaces located in clinical settings. Soft surfaces, such as patient pyjamas and beddings, can be an excellent substrate for bacterial and fungal growth under appropriate temperature and humidity conditions as those present between patients and the bed. Bed making in hospitals releases large quantities of microorganisms into the air, which contaminate the immediate and non-immediate surroundings. Microbes can survive on hard surfaces, such as metal trays, bed rails and door knobs, for very prolonged periods of time. Thus soft and hard surfaces that are in direct or indirect contact with the patients can serve as a source of nosocomial pathogens. Recently it has been demonstrated that copper surfaces and copper oxide containing textiles have potent intrinsic biocidal properties. This manuscript reviews the recent laboratory and clinical studies, which demonstrate that biocidal surfaces made of copper or containing copper can reduce the microbiological burden and the NI rates.

Key Words: Nosocomial infections, Health acquired infections, Copper, Copper oxide, Biocides, Surfaces, Microbiological burden



INTRODUCTION

A nosocomial, or hospital-acquired, infection is a new infection that develops in a patient during hospitalization. Nosocomial infections (NI) are a worldwide problem that occur both in developed and in developing countries. For example, in the United States approximately 2 million patients annually contract an infection while being hospitalized[1], and it is the fourth among the causes of death in the United States only behind heart disease, cancer and stroke[2]; in Europe in 2007 there were about 3 million healthcare associated infections (HAI), of which approximately 50 000 resulted in death[3]; in Germany alone around 500 000 to 600 000 NI occurred during 2006[4]; methicillin-resistant Staphylococcus aureus (MRSA) infections alone are estimated to affect more than 150 000 patients annually in the European Union[5]; in Australia, more than 177 000 NI occur per year[6]; in the province of Quebec, Canada, the rate of NI is estimated to be around 11%[7]; and the rates of NI in developing countries are even higher[8-11].

NI can be bacterial, viral, fungal, or even parasitic[12-15]. Some of the most common nosocomial pathogens are staphylococci (especially Staphylococcus aureus), Pseudomonas aeruginosa (P. aeruginosa), Escherichia coli (E. coli), Clostridium difficile (C. difficile), Streptococcus species, Enterobacter species, Acinetobacter species, Klebsiella species, influenza virus and noroviruses[16-21]. The prevalence rates of pathogens that cause NI and have a high level of resistance to antibiotic treatments, such as multidrug-resistant (MDR) P. aeruginosa, extended-spectrum β-lactamase producing Enterobacteriaceae, MDR Acinetobacter baumannii, MRSA, and vancomycin resistant enterococci (VRE), are constantly increasing around the globe[22-28], creating a serious threat to the spread and treatment of infectious diseases, because the resistant pathogens are significantly more difficult to treat (e.g.,[29]).

Many measures to reduce the risk of pathogens transmission are sought by health care officials, physicians and scientists. These include improvement of national surveillance of NI, use of aggressive antibiotic control programs to reduce the spread of antibiotic-resistant strains, healthcare staff education for improved hygiene, isolation of infected patients, ultraviolet light sterilization, use of disposable equipment, development of patient care techniques to reduce risks of infection, improved cleaning techniques, improvement of cleaning equipment and sanitary facilities, increase in nursing and janitorial resources and better nutrition (e.g.,[30-34]). It is estimated that by using several of the above strategies simultaneously about one third of NI may be eliminated[35,36]. These measures are not the scope of this review and are widely described elsewhere in the literature (e.g.,[31,37-40]). But it is clear that even in clinical settings where all or most of these measures are implemented, the rates of NI are still too high, and thus new approaches to further fight these infections need to be explored.

NI may occur via several manners. It is recognized by the infection control community that the most important and frequent modes of transmission of nosocomial pathogens are through direct-contact between an infected or colonized person (e.g., health worker, visitor or patient) and a susceptible host[41-44], and indirectly via contaminated intrusive medical devices[44-49], from the patient’s own flora from one part of the host’s body to another[50], and via airborne particles[21,51-55].

In addition to the above well described modes of transmission of nosocomial pathogens, others[56] and we[57] hypothesised that contaminated textiles in hospitals might be an important source of microbes contributing to endogenous, indirect-contact, and aerosol transmission of nosocomial-related pathogens. Textiles are an excellent substrate for bacterial and fungal growth under appropriate moisture and temperature conditions, and it was shown that bacteria and fungi can survive for prolonged periods in hospital fabrics[58,59]. Microbial shedding from the body occurs continuously[60]. Microbial shedding is greater in patients[54,61]. Thus a bacterium, when shed into a textile fabric between the patient and the bed, either on his pyjama, pillowcase, sheet, or mattress, would readily proliferate since the moisture and temperature in the textile microenvironment would promote its proliferation. Others and we presented data that substantiate this premise[62-65]. Importantly, it was found by others that bed making releases large quantities of microorganisms into the atmosphere and the bacteria levels in the air fall back to background levels only after approximately 30 min[52,66-68]. The released bacteria were shown to contaminate adjacent surfaces, such as bed sheets, over bed tables, and patients’ clothing, and even adjacent rooms via the air-conditioning systems. Similar results were reported following undressing and redressing of patients[69].

The contribution of contaminated hard surfaces, such as floors, bedrails, bedside tables and door knobs, to NI has been demonstrated too (e.g.,[70-80]). Similarly, contaminated textiles, such as contaminated sheets and pyjamas, in addition to being a source of aerosol transmission of microorganisms, can also directly contaminate the hospital personnel[56,76,81,82]. Hospital staff, even by using protective equipment such as gloves, can contaminate them by touching the contaminated textiles or contaminated surfaces and then transferring the microorganisms to other patients directly or indirectly by contaminating other surfaces, such as door knobs[76,83]. For example, it was found that 65% of the nurses who performed activities on patients with MRSA in wounds or urine, contaminated their nursing uniforms or gowns with MRSA. This in turn, can readily contaminate the clothing and hands of healthcare workers[54,76,83]. High similar contamination of gloves and gowns with MDR Acinetobacter baumannii by healthcare workers interacting with colonized patients has also been reported[84]. Furthermore, it was found that 42% of personnel with no direct contact with patients contaminated their gloves by touching contaminated surfaces[76].

Thus, we further hypothesized that use of antimicrobial textiles, especially in those that are in close contact with the patients, may significantly reduce bioburden in clinical settings and consequently reduce the risk of NI[57]. Being all surfaces biocidal in a hospital environment would further reduce the risk of pathogen transmission and NI since most common nosocomial pathogens can remain viable on surfaces for months[43,85]. Indeed, it has been shown that environmental disinfection interrupts the transmission of microbial pathogens[79,80,83,86,87]. However, there are increasing concerns that routine surface disinfection procedures in health care settings are frequently inadequate and possibly counterproductive[88,89]. Consequently, the notion that having potent safe biocidal non-intrusive hard and soft surfaces in medical settings, in direct or indirect contact with patients, capable of reducing the microbiological burden that would significantly contribute to reduction in transmission of nosocomial pathogens, is gaining recognition by the scientific community. This review focuses on the studies demonstrating that hard and soft surfaces containing copper reduce the microbiological burden in clinical settings and the NI rates.

COPPER HAS POTENT BIOCIDAL PROPERTIES

Copper and copper compounds have a wide spectrum of antibacterial, antifungal and antiviral properties (reviewed in[90,91]). The wide range of microorganisms, including gram negative and gram positive bacteria, yeast, fungi and enveloped and non-enveloped viruses, that have been shown to be killed by copper or copper compounds, are summarized in Table 1. Importantly, copper surfaces or copper compounds have also been shown to be efficacious against hard-to-kill spores[92-98].

Table 1 Demonstrated biocidal efficacy of copper.
Hard surfaceSoft surfaceOtherRef.
Bacteria
Acinetobacter baumannii1+++[130,164] UR2
Acinetobacter calcoaceticus/baumannii--+[93,94,165]
Acinetobacter johnsonii+--[105]
Acinetobacter lwoffii--+[166]
Bacillus cereus+-+[101,167-169]
Bacillus globigii--+[92]
Bacillus subtilis-++[165,169-175]
Bacillus macerans--+[176]
Brachybacterium conglomeratum+--[105]
Brevibacterium-+-UR
Campylobacter jejuni+--[129]
Citrobacter freundi--+[165,177]
Clostridium difficile+-+[93,97,98]
Clostridium tyrobutyricum--+[95]
Corynebacterium xerosis-+-UR
Deinococcus radiodurans+--[101]
Desulfovibrio desulfuricans--+[178]
Edwardsiella tarda--+[179]
Enterobacter aerogenes--+[168,180]
Enterobacter cloacae+++[127,128,168,175]
Enterococcus sp.1--+[93]
Enterococcus faecalis1+++[64,108,112,136,137,168,180]
Enterococcus faecium1+--[127,128,137,155,181]
Enterococcus gallinarum+--[137]
Enterococcus hirae+--[182]
Escherichia coli+++[64,100,101,105,108,109,112,127,128,133,139,147,155,165, 168-172,181,183-193]
Klebsiella pneumoniae+++[112,130,165,193-195]
Kocuria marina+--[105]
Kocuria palustris+--[105]
Legionella pneumophila+-+[93,140,159,196-198]
Listeria monocytogenes+++[64,140,180,199,200]
Mycobacterium tuberculosis1+--[130]
Micrococcus luteus++-[105,127,128] UR
Morganella morganii--+[177]
Pantoea stewartii+--[105]
Photobacterium leiognathi-+-[112]
Proteus mirabilis--+[194]
Proteus vulgaris--+[168]
Pseudomonas aeruginosa+++[112,127,128,130,144,164,167,168,171,172,175,201,202]
Pseudomonas fluorescens+--[199]
Pseudomonas nitroreducens--+[169]
Pseudomonas oleovorans+--[105]
Pseudomonas putita--+[203]
Pseudomonas striata+--[176]
Salmonella spp.+++[64,129,165,183]
Salmonella typhi+-+[141,174,177,190,194,203,204]
Salmonella typhimurium+--[141,142,199,201]
Sarcina lutea--+[167]
Serratia marcescens--+[171]
Shewanella putrefaciens+--[199]
Shigella dysenteriae--+[194]
Shigella flexnerii+-+[165,174,177,204]
Sphingomonas panni+--[105]
Staphylococcus aureus1+++[64,93,94,105,108,109,112,127,128,130,131,134,138,165,167-172, 175,181,184,199,200,205,206]
Staphylococcus epidermidis+++[105,168,191,195,207] UR
Staphylococcus haemolyticus+--[105]
Staphylococcus hominis+--[105]
Staphylococcus warnerii+--[105]
Stenotrophomonas maltophilia--+[164]
Streptococcus faecalis-+-[175]
Streptococcus pyogenes--+[168]
Streptococcus sp.--+[165,208]
Vibrio cholerae1+-+[141,190,209]
Yersinia pseudotuberculosis--+[180]
Xanthomonas compestris--+[202]
Fungi/Yeast
Alternaria brassicae--+[202]
Aspergillus bransilensis-+-UR
Aspergillus carbonarius--+[210]
Aspergillus flavus+-+[96,172,203,204]
Aspergillus fumigatus+-+[96,211]
Aspergillus niger+++[96,114,172,202,211-214]
Aspergillus oryzae--+[212]
Candida albicans+++[64,96,104,108,109,112-114,130,168,169,173,193,204,211,214,215]
Candida glabrata--+[168,180,194,204]
Candida krusei--+[168]
Candida parapsilosis--+[168]
Candida tropicalis--+[168,180]
Cronobacter sakazakii--+[216]
Cryptococcus neoformans--+[211]
Culvularia lunata--+[195]
Epidermophyton floccosum--+[211]
Fusarium culmonium+--[96]
Fusarium oxysporium+-+[96,202]
Fusarium solani+-+[96,195,204]
Microsporum canis--+[204,211]
Myrothecium verrucaria--+[212]
Penicillium chrysogenum+--[96]
Pleurotus ostreatus--+[185]
Pycnoporus cinnabarinus--+[185]
Rhizoctonia bataicola--+[195,203]
Rhizoctonia solani--+[213]
Rhizopus stolonifer--+[203]
Saccharomyces cerevisiae+-+[103,104,169,217]
Torulopsis pintolopesii--+[215]
Trichoderma viride--+[212]
Trichophyton longifusus--+[204]
Trichophyton mentagrophytes-++[113,114,194,212]
Tricophyton rubrum-++[113,211]
Tricophyton schoenleinii--+[194]
Virus
Avian influenza-++[111,205]
Adenovirus type 1++-[99,218]
Bacteriophages--+[219-223]
Coxsackie virus types B2 and B4+--[218]
Cytomegalovirus-+-[99]
Echovirus 4+--[218]
Herpes simplex virus--+[219,220]
Human immunodeficiency virus-++[99,108,110,224]
Infectious bronchitis virus--+[225]
Influenza A++-[99,111,135]
Junin virus--+[220]
Measles-+-[99]
Parainfluenza 3-+-[99]
Poliovirus+-+[222,226]
Pichinde-+-[99]
Punta Toro-+-[99]
Respiratory syncytial virus-+-[99]
Rhinovirus 2-+-[99]
Simian rotavirus SA11+--[218]
Vaccinia-+-[99]
West nile virus-+-[108]
Yellow fever-+-[99]

Copper exerts its toxicity to microorganisms through several parallel mechanisms, which eventually may lead to the microorganisms’ death even within minutes of their exposure to copper[94,99-106]. These include plasma membrane permeabilization, membrane lipid peroxidation, alteration of proteins and inhibition of their biological assembly and activity and denaturation of nucleic acids[90,91]. In general, the redox cycling between Cu2+ and Cu1+, which can catalyze the production of highly hydroxyl radicals, with subsequent damage to lipids, proteins, DNA and other biomolecules[90,107], makes copper further reactive and a particularly effective antimicrobial. Interestingly, two different “kill modes”, under dry and wet conditions, have been attributed to copper surfaces[101,102,104,105].

BIOCIDAL SOFT SURFACES IN THE HEALTHCARE ENVIRONMENT

Copper oxide is a non-soluble form of copper that, similarly to other copper compounds, has potent wide spectrum biocidal properties[90]. It has, therefore, been chosen as the active copper form to be introduced into textile fibres from which woven and non-woven fabrics can be produced[64,108,109]. These copper-impregnated products possess permanent broad-spectrum anti-bacterial, anti-fungal and antiviral properties that are not affected by washings[64,91,99,108-112] (Table 1). This technology, for example, enables the production of biocidal fabrics (which inter alia kill antibiotic resistant bacteria)[64,91,108,109], anti-fungal socks (which inter alia alleviate symptoms of athlete’s foot)[108,113], anti-viral masks and filters (which inter alia deactivate HIV-1, Influenza A and other viruses)[99,106,110,111], and anti-dust mite mattress-covers (which may reduce mite-related allergies)[108,114].

As explained in the previous chapter, we hypothesized that contaminated beddings may be an important overlooked source of nosocomial pathogens and therefore the use of potent biocidal beddings, especially pyjamas and sheets, that are in contact with the patients, may significantly reduce bioburden in clinical settings and consequently reduce the risk of NI[57]. Indeed, a pilot study with 30 patients, who slept overnight on regular sheets and then overnight on sheets containing copper-oxide demonstrated a statistically significant lower bacterial colonization on the copper-oxide containing sheets than on regular-sheets[64], clearly supporting our hypothesis.

Importantly, the development of biocidal textiles with the purpose of using them in clinical settings to reduce HAI is gaining momentum and other biocidal active ingredients have or are being explored. These include Cliniweave®[115], organofunctional silane[116], citric acid[117], silver[118,119], triclosan[120], quaternary ammonium compounds[121], chitosan and zeolite[122,123]. For biocidal textiles to be introduced into the hospital textiles they should have wide spectrum antimicrobial, antifungal and antiviral properties, be effective against the already existent antibiotic resistant microorganisms involved in NI, not allow for the development of microorganisms against the active component in them, be efficacious for the life of the material, not be affected by commercial washings, not cause skin irritation or sensitization and be safe to humans following continuous dermal exposure. Some of the above active ingredients have thus been found not to be appropriate for use in hospital related applications (e.g.,[120,124]).

Until recently, only a few trials in clinical settings have been performed with biocidal textiles. It was found that bioburden was significantly lower on garments worn by nurses when the garments were made from a silver and copper containing antibacterial fabric[125]. The antibacterial textiles were tested in two hospital units, an oncology surgery unit and an intensive care unit. Each garment was provided with a piece of test fabric sewed either on the right or left side of the garment, while the regular fabric of the garment on the other side was used as a control. Thirty garments were tested in each unit. They were all sterilized, so they would be free of bacteria at the beginning of the experiment. The nurses wore the same number of garments with the treated area on the left side 1 d and on the right side the following day. Both active and control sides of each garment were sampled simultaneously and the bioburden determined. The number of colony forming units (CFU) was significantly lower on the bioactive patches than on the control areas. The mean reduction rate was about 30% for the 60 garments tested. Reduction of about 50% of bioburden on sheets containing copper oxide compared to regular sheets, when used overnight by general ward patients, was demonstrated[64]. Similarly, reduction of bioburden on blankets containing a bound organofunctional silane was also reported[116]. Recently in a 16 wk, blinded cross-over clinical trial that compared levels of bacterial contamination, a significantly fewer MRSA colonies were detected on scrubs impregnated with nano-sized particles that increase the surface tension of the scrubs than on standard scrubs (http://www.vestexprotects.com/press/view/8-Vestagen-Announces-Completion-of-First-Clinical-Trial-of-Vestex). In contrast, a study that compared the contamination rates of silver containing jackets and pants and of standard textile clothing used by 10 emergency workers did not find any significant difference in the extent of microbial contamination between the textiles[119]. It may be that a larger sample size was required to prove the silver containing fabric efficacy. It should be taken into consideration that in contrary to in vitro conditions, a continual re-inoculation with pathogens occurs during real-life health care scenarios. In addition, the killing of the microorganisms is not on contact, as it takes time for the biocidal textiles to kill the exposed microorganisms. Thus, obtaining sterile hospital or health-care associated fabrics by biocidal textiles in a healthcare environment cannot be expected. Obviously, trials demonstrating that the use of biocidal textiles does not only reduce bioburden in clinical settings, but also reduces NI rates, still need to be conducted.

BIOCIDAL HARD SURFACES IN THE HEALTHCARE ENVIRONMENT

On February 2008 the USA Environmental Protection Agency (EPA) permitted the USA Copper Association to make public health claims and state that copper alloy products kill 99.9% of disease causing bacteria within 2 h and continue to do so when re-exposed[126]. This approval has now been given to 355 different copper alloys (including brass and bronze) following many years of independent laboratory testing based on rigorous EPA approved protocols. Copper is the only hard surface metal that has received approval by the EPA to make antimicrobial public health claims. In addition to the tests conducted by the USA Copper Association in order to obtain the approval for the registered health claims, the biocidal properties of copper surfaces was demonstrated by many others as well[96-98,100-102,104,105,127-142]. As can be seen in Table 1, copper surfaces can be regarded as a wide spectrum biocidal surface, as it has been found to be efficacious against a wide array of gram positive and negative bacteria, fungi and viruses. The biocidal efficacy of copper surfaces increases with the copper concentration[97,101,104,127,128,130,133,134,137,139], exposure periods[96-98,100-102,104,127-130,133-135,137,139,140,143], humidity[127,128,131,136] and temperature[98,129,131,133,144]. The higher the microorganism inoculum load is the longer it takes to reach complete elimination of the exposed microorganisms[133,134,137]. In contrast to stainless steel, which is the metal most widely used in hospital care environments, copper surfaces are highly reactive, and thus residual soil and build-up of microbial cells is more likely to occur in copper surfaces than on stainless steel[145]. Different cleaning solutions or products may have different effects on the continual efficacy of the copper surfaces[145] and thus the right cleaning and appropriate cleaning protocols of copper surfaces need to be developed[102].

Importantly, the significant contribution of copper surfaces to the reduction of bioburden in clinical settings has recently been demonstrated[132,146,147]. One trial was conducted in the United Kingdom[146], one in South Africa[147] and one in Germany[132].

In the United Kingdom study[146] the efficacy of copper surfaces to reduce bioburden was examined in a busy acute medical ward, which included gastroenterology patients, and a cross-over model was utilized. A toilet seat, tap handles and a ward entrance door push plate each containing copper (60%-70% copper content) were sampled for the presence of microorganisms and compared to equivalent standard, non-copper-containing items in the same ward. The items were installed at least 6 mo prior to the commencement of the study to allow both healthcare workers and staff to become accustomed to the copper containing items. The hospital staff followed their standard cleaning routines, which included disinfection of both the control and test fixtures approximately every 2 h. The items were sampled once weekly for 10 wk at 07:00 and at 17:00 to determine the number of microorganisms present following quite and busy time periods, respectively. The following specific indicator bacteria were quantified: methicillin-sensitive Staphylococcus aureus (MSSA), MRSA, VRE, C. difficile and E. coli. After 5 wk, the copper-containing and non-copper-containing items were interchanged to exclude any possibility of bias according to preferential use of any particular item based on location. Median numbers of microorganisms harbored by the copper-containing items were between 90% and 100% lower than their control equivalents at both sampling time-points, the microbial loads being highly statistically significantly different between the matched tested items (P values ranging from < 0.05 to < 0.0001). Three of the indicator microorganisms (MSSA, VRE and E. coli) were only isolated from control items. MRSA and C. difficile were not isolated during this study.

In the South Africa study[147], a comparative controlled study was conducted at a busy walk-in primary healthcare clinic in a rural region. Two similar adjacent consulting rooms were chosen. One was fitted with copper sheets (99.9% pure copper) on desk and trolleys that were in constant contact with staff and patients and on top of cupboards and windowsills where contact was less frequent. The other room remained with its original surfaces that did not include any copper surfaces. Cleaning procedures were the same for both rooms and no disinfectants were used. Samples for microbiological determinations were taken from 5 equivalent touch surfaces from each room. Sampling was undertaken for a period of 4 and a half days every 6 wk by the same person for a period of 6 mo. Samples were taken before cleaning (at 7 am), post cleaning but pre consultation (at 8 am) and post consultation (at 4 pm). The temperature and humidity in both sampling rooms were comparable during the study period covered - winter, spring and summer. The average number of consultations in each room during each sampling series during the 6 mo study was similar (65 study and 68 control room). Statistically significantly lower overall mean total CFU for all copper surfaces, including those in constant contact with staff and patients and those with less frequent contact, were found (P < 0.001), being the mean reduction 71%.

In the German study[132], an oncological/pneumological and a geriatric ward was used to test the efficacy of copper surfaces in reducing bioburden. All touch surfaces in patient bed rooms, rest rooms and staff rooms were replaced with new surfaces composed of metallic copper-containing alloys, while matched rooms, where no changes were made in the touch surfaces, served as controls rooms. All surfaces were routinely cleaned each morning with a disinfectant. The trial lasted 32 wk, 16 in the summer and 16 in the winter. During both test periods of 16 wk, the total number of CFU on metallic copper-containing surfaces was 63% of that on the control surfaces (P < 0.001). When analyzing per surface area, the differences were significant for door knobs, which had the highest overall microbial load. Bacterial loads in push plates and light switches were similar between the test and control samples. Interestingly, after disinfection of the copper and control surfaces, microbial repopulation of the surfaces was significantly delayed on copper alloys (P < 0.05).

In addition to the above studies, a clinical study was undertaken to compare the surface microbial contamination associated with pens constructed of either a copper alloy or stainless steel used by nurses on intensive care units. A significantly lower level of microbial contamination was found on the copper alloy pens[148].

Another study, conducted in the UK, investigated the efficacy of using biocidal hard surfaces impregnated with a silver based technology in reducing microbial contamination in a real-life hospital environment[149]. Two outpatient units were included in the 18 mo study. One unit was refurbished with the silver containing products, which included door knobs, blinds, tiles, sack holders and light switches. The other unit contained untreated items and served as a control. Both units were similar in terms of volume of people and layout and were subjected to similar standard cleaning practice. Both units were allowed to function for 12 mo before microbiological swabbing commenced. Swabs were collected over a 5-mo period from both units. The CFU counts in the unit containing the silver impregnated products were between 62% to 98% lower than the matched unit. CFU counts from the silver-treated materials were between 70% (fabrics) to 99% (laminates) lower than untreated equivalents. In addition, the bacterial contamination on untreated products in the ward containing the silver-impregnated products was on average 43.5% lower compared with untreated matched products in the control unit.

The above described trials clearly demonstrate that biocidal hard surfaces found in heath-care settings offer the potential to significantly reduce the number of microorganisms in the clinical environment and thus reduce the risk of HAI. However, the use of biocidal surfaces should not act as a replacement for cleaning in clinical areas, but as an adjunct in the fight against HAI.

IS MICROBIAL RESISTANCE TO COPPER A CONCERN?

Bacterial resistance is a major concern in infection control, as exemplified by the highly antibiotic resistant bacteria (with up to 2200-fold decreased sensitivity to the antibiotic (e.g.,[150]) that have evolved in less than 50 years of antibiotic usage, making infected patient treatment extremely difficult (e.g.,[29]). Thus, the possibility of development of resistance to biocides is a real concern[151,152]. Importantly, as opposed to antibiotics, in spite of copper being a part of the earth for millions of years, and being used by humans from the beginning of the civilization, no microorganisms that are highly resistant to copper have been found, but only microorganisms with reduced copper sensitivity (increased copper tolerance). For example, Enterococci bacteria isolated from the gut of pigs, which were fed for many months with high concentrations of copper in their diet, were 7 fold less susceptible to copper than Enterococci bacteria isolated from pigs not fed with copper[153,154]. The increased tolerance to copper is achieved by the induction of an efflux pump in the tolerant bacteria[154]. Outstandingly, the Enterococci and E. coli tolerant bacteria isolated from pig farms following the use of copper sulfate as feed supplement were rapidly killed when spread in a thin, moist layer on copper alloys with 85% or greater copper content or under dry conditions[155]. Tolerance, but not resistance, was found in nitrifying soil microorganisms exposed to Cu for nearly 80 years under field conditions[156]. Similarly, the spray of copper-containing compounds for years on vegetable and fruit crops to limit the spread of plant pathogenic bacteria and fungi, has favored the spread of copper tolerant genes among saprophytic and plant pathogenic bacteria[157]. The increased tolerance to copper was found to be associated with the amount of soluble copper and not with the total amount of copper[158]. Thus, even in soils where the concentration of copper was very high, but in a non-soluble form, no increase in tolerance to copper was observed[158]. The copper active ingredient used in the biocidal textiles is copper oxide, a non-soluble form of copper. Importantly, no resistant bacteria evolved in vitro when consecutively exposed to repeated fabrics containing 1% copper oxide[112]. Interestingly, bacteria were isolated from copper-containing surfaces and some exhibited prolonged (1 to 3 d) survival on dry but not on moist copper surfaces[105]. None of these isolates strains was copper resistant in culture[105]. Survival on copper-containing surfaces appeared to be the consequence of either endospore formation, survival on patches of dirt, or a special ability to endure a dry metallic copper surface.

The reason why no resistance to copper, but only tolerance, is found in microorganisms exposed to constant relatively high doses of copper, may be because copper exerts its biocidal/antimicrobial activity not through one mechanism (as most antibiotics), but through several parallel non-specific mechanisms[90,91]. As briefly mentioned previously, these mechanisms include: (1) denaturation of nucleic acids by binding to and/or disordering helical structures and/or by cross-linking between and within nucleic acid strands; (2) alteration of proteins and inhibition of their biological assembly and activity; (3) plasma membrane permeabilization; and (4) membrane lipid peroxidation. Furthermore, widespread appearance of bacteria tolerant or resistant to copper contact killing appears unlikely as plasmid DNA is completely degraded after cell death by contact killing, preventing the transfer of resistance determinants between organisms[137] and copper contact killing is very rapid precluding the acquisition of resistance during cell division[102].

Thus, even though some organisms have mechanisms of tolerance to excess copper as described above, in general, all microorganisms cannot cope when exposed to high concentrations of copper and are irreversibly damaged. As a result, despite having been present throughout human history, and despite repeated historic use of copper as an antimicrobial agent over the centuries, copper was and remains a broad-spectrum biocidal/antimicrobial compound and yet no bacteria fully resistant to copper have been discovered.

CONCLUSION

Similar to the efficient control of Legionella infections and the reduction of molds and yeasts that has been achieved in hospital systems by simply incorporating copper-silver ionization devices into the hospital water distribution systems[159-161], the use of soft and hard surfaces containing biocidal copper in products such as those described in Figure 1, may play an important role in reduction of NI in hospital care environments. Furthermore, as NI are now spreading out from the hospital environment into the community (e.g.,[162,163]), the use of textiles, such as those impregnated with copper oxide, and hard surfaces containing a high percentage of copper, may not only significantly contribute to the reduction of HAI, but may also confer protection when used in homes for the elderly and in other environments where immune compromised individuals are at high risk of contracting infections.

Figure 1
Figure 1 Use of copper in the detailed products, which are in direct or indirect contact with patients, may significantly contribute to the reduction of nosocomial pathogen loads and nosocomial infections. HEPA:
ACKNOWLEDGMENTS

I thank Ms. Myriam Edith Gargiulo for her technical help and good suggestions.

Footnotes

Peer reviewer: Guadalupe García-Elorriaga, PhD, Hospital de Infectología, Hospital de Infectología, CMN La Raza. Mexico City, Mexico City 02992, Mexico

S- Editor Wu X L- Editor A E- Editor Zheng XM

References
1.  Vallés J, Ferrer R. Bloodstream infection in the ICU. Infect Dis Clin North Am. 2009;23:557-569.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 27]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
2.  Klevens RM, Edwards JR, Richards CL, Horan TC, Gaynes RP, Pollock DA, Cardo DM. Estimating health care-associated infections and deaths in U.S. hospitals, 2002. Public Health Rep. 2007;122:160-166.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  European Centre for Disease Prevention and Control. Annual Epidemiological Report on Communicable Diseases in Europe 2010.  Available from: http://www.nric.org.uk/integratedcrd.nsf/d20f15baa90306438025755c0048057f/cb4a43418caa9428802577fb004c0c7d?OpenDocument.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Kerwat K, Graf J, Wulf H. [Nosocomial infections]. Anasthesiol Intensivmed Notfallmed Schmerzther. 2010;45:30-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 6]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
5.  Köck R, Becker K, Cookson B, van Gemert-Pijnen JE, Harbarth S, Kluytmans J, Mielke M, Peters G, Skov RL, Struelens MJ. Methicillin-resistant Staphylococcus aureus (MRSA): burden of disease and control challenges in Europe. Euro Surveill. 2010;15:19688.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Ferguson JK. Preventing healthcare-associated infection: risks, healthcare systems and behaviour. Intern Med J. 2009;39:574-581.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 29]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
7.   Available from: http://www2.publicationsduquebec.gouv.qc.ca/dynamicSearch/telecharge.php?type=5&file=2002C71F.PDF.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Joshi R, Reingold AL, Menzies D, Pai M. Tuberculosis among health-care workers in low- and middle-income countries: a systematic review. PLoS Med. 2006;3:e494.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 300]  [Cited by in F6Publishing: 330]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
9.  Hughes AJ, Ariffin N, Huat TL, Abdul Molok H, Hashim S, Sarijo J, Abd Latif NH, Abu Hanifah Y, Kamarulzaman A. Prevalence of nosocomial infection and antibiotic use at a university medical center in Malaysia. Infect Control Hosp Epidemiol. 2005;26:100-104.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 49]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
10.  Rosenthal VD, Maki DG, Salomao R, Moreno CA, Mehta Y, Higuera F, Cuellar LE, Arikan OA, Abouqal R, Leblebicioglu H. Device-associated nosocomial infections in 55 intensive care units of 8 developing countries. Ann Intern Med. 2006;145:582-591.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Rosenthal VD. Device-associated nosocomial infections in limited-resources countries: findings of the International Nosocomial Infection Control Consortium (INICC). Am J Infect Control. 2008;36:S171.e7-S171.12.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Alangaden GJ. Nosocomial fungal infections: epidemiology, infection control, and prevention. Infect Dis Clin North Am. 2011;25:201-225.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 128]  [Cited by in F6Publishing: 104]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
13.  Peleg AY, Hooper DC. Hospital-acquired infections due to gram-negative bacteria. N Engl J Med. 2010;362:1804-1813.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 937]  [Cited by in F6Publishing: 897]  [Article Influence: 64.1]  [Reference Citation Analysis (0)]
14.  Turnidge JD, Kotsanas D, Munckhof W, Roberts S, Bennett CM, Nimmo GR, Coombs GW, Murray RJ, Howden B, Johnson PD. Staphylococcus aureus bacteraemia: a major cause of mortality in Australia and New Zealand. Med J Aust. 2009;191:368-373.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Nevez G, Chabé M, Rabodonirina M, Virmaux M, Dei-Cas E, Hauser PM, Totet A. Nosocomial Pneumocystis jirovecii infections. Parasite. 2008;15:359-365.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Jones RN. Microbial etiologies of hospital-acquired bacterial pneumonia and ventilator-associated bacterial pneumonia. Clin Infect Dis. 2010;51 Suppl 1:S81-S87.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 444]  [Cited by in F6Publishing: 459]  [Article Influence: 32.8]  [Reference Citation Analysis (0)]
17.  Utsumi M, Makimoto K, Quroshi N, Ashida N. Types of infectious outbreaks and their impact in elderly care facilities: a review of the literature. Age Ageing. 2010;39:299-305.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 83]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
18.  Leclair MA, Allard C, Lesur O, Pépin J. Clostridium difficile infection in the intensive care unit. J Intensive Care Med. 2010;25:23-30.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 19]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
19.  Koopmans M. Noroviruses in healthcare settings: a challenging problem. J Hosp Infect. 2009;73:331-337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 38]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
20.  Zilberberg MD. Clostridium difficile-related hospitalizations among US adults, 2006. Emerg Infect Dis. 2009;15:122-124.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 42]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
21.  Lindsley WG, Blachere FM, Davis KA, Pearce TA, Fisher MA, Khakoo R, Davis SM, Rogers ME, Thewlis RE, Posada JA. Distribution of airborne influenza virus and respiratory syncytial virus in an urgent care medical clinic. Clin Infect Dis. 2010;50:693-698.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Mulvey MR, Simor AE. Antimicrobial resistance in hospitals: how concerned should we be? CMAJ. 2009;180:408-415.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 89]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
23.  Kunz AN, Brook I. Emerging resistant Gram-negative aerobic bacilli in hospital-acquired infections. Chemotherapy. 2010;56:492-500.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 82]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
24.  Coque TM, Baquero F, Canton R. Increasing prevalence of ESBL-producing Enterobacteriaceae in Europe. Euro Surveill. 2008;13:pii=19044.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Song JH, Chung DR. Respiratory infections due to drug-resistant bacteria. Infect Dis Clin North Am. 2010;24:639-653.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 19]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
26.  Dean N. Methicillin-resistant Staphylococcus aureus in community-acquired and health care-associated pneumonia: incidence, diagnosis, and treatment options. Hosp Pract (Minneap). 2010;38:7-15.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Heintz BH, Halilovic J, Christensen CL. Vancomycin-resistant enterococcal urinary tract infections. Pharmacotherapy. 2010;30:1136-1149.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 51]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
28.  Croft AC, D'Antoni AV, Terzulli SL. Update on the antibacterial resistance crisis. Med Sci Monit. 2007;13:RA103-RA118.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Durai R, Ng PC, Hoque H. Methicillin-resistant Staphylococcus aureus: an update. AORN J. 2010;91:599-606; quiz 607-609.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 35]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
30.  Pratt RJ, Pellowe CM, Wilson JA, Loveday HP, Harper PJ, Jones SR, McDougall C, Wilcox MH. epic2: National evidence-based guidelines for preventing healthcare-associated infections in NHS hospitals in England. J Hosp Infect. 2007;65 Suppl 1:S1-S64.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 514]  [Cited by in F6Publishing: 407]  [Article Influence: 23.9]  [Reference Citation Analysis (0)]
31.  Curtis LT. Prevention of hospital-acquired infections: review of non-pharmacological interventions. J Hosp Infect. 2008;69:204-219.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 104]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
32.  Hamilton D, Foster A, Ballantyne L, Kingsmore P, Bedwell D, Hall TJ, Hickok SS, Jeanes A, Coen PG, Gant VA. Performance of ultramicrofibre cleaning technology with or without addition of a novel copper-based biocide. J Hosp Infect. 2010;74:62-71.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
33.  Wren MW, Rollins MS, Jeanes A, Hall TJ, Coën PG, Gant VA. Removing bacteria from hospital surfaces: a laboratory comparison of ultramicrofibre and standard cloths. J Hosp Infect. 2008;70:265-271.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 48]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
34.  Dancer SJ, White LF, Lamb J, Girvan EK, Robertson C. Measuring the effect of enhanced cleaning in a UK hospital: a prospective cross-over study. BMC Med. 2009;7:28.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 185]  [Cited by in F6Publishing: 189]  [Article Influence: 12.6]  [Reference Citation Analysis (0)]
35.  Weinstein RA. Nosocomial infection update. Emerg Infect Dis. 1998;4:416-420.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 253]  [Cited by in F6Publishing: 260]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
36.  Scheckler WE, Brimhall D, Buck AS, Farr BM, Friedman C, Garibaldi RA, Gross PA, Harris JA, Hierholzer WJ, Martone WJ. Requirements for infrastructure and essential activities of infection control and epidemiology in hospitals: a consensus panel report. Society for Healthcare Epidemiology of America. Infect Control Hosp Epidemiol. 1998;19:114-124.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 51]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
37.  Bearman GM, Munro C, Sessler CN, Wenzel RP. Infection control and the prevention of nosocomial infections in the intensive care unit. Semin Respir Crit Care Med. 2006;27:310-324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 22]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
38.  Boyce JM, Pittet D. Guideline for Hand Hygiene in Health-Care Settings. Recommendations of the Healthcare Infection Control Practices Advisory Committee and the HIPAC/SHEA/APIC/IDSA Hand Hygiene Task Force. Am J Infect Control. 2002;30:S1-46.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 430]  [Cited by in F6Publishing: 453]  [Article Influence: 20.6]  [Reference Citation Analysis (0)]
39.  Dancer SJ. Hospital cleaning in the 21st century. Eur J Clin Microbiol Infect Dis. 2011;30:1473-1481.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Muto CA, Jernigan JA, Ostrowsky BE, Richet HM, Jarvis WR, Boyce JM, Farr BM. SHEA guideline for preventing nosocomial transmission of multidrug-resistant strains of Staphylococcus aureus and enterococcus. Infect Control Hosp Epidemiol. 2003;24:362-386.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 903]  [Cited by in F6Publishing: 1110]  [Article Influence: 52.9]  [Reference Citation Analysis (0)]
41.  Kampf G, Kramer A. Epidemiologic background of hand hygiene and evaluation of the most important agents for scrubs and rubs. Clin Microbiol Rev. 2004;17:863-893, table of contents.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 460]  [Cited by in F6Publishing: 415]  [Article Influence: 20.8]  [Reference Citation Analysis (0)]
42.  Ayliffe GA, Babb JR, Davies JG, Lilly HA. Hand disinfection: a comparison of various agents in laboratory and ward studies. J Hosp Infect. 1988;11:226-243.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 163]  [Cited by in F6Publishing: 169]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
43.  Dancer SJ. Importance of the environment in meticillin-resistant Staphylococcus aureus acquisition: the case for hospital cleaning. Lancet Infect Dis. 2008;8:101-113.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 348]  [Cited by in F6Publishing: 299]  [Article Influence: 18.7]  [Reference Citation Analysis (0)]
44.  Gastmeier P, Stamm-Balderjahn S, Hansen S, Zuschneid I, Sohr D, Behnke M, Vonberg RP, Rüden H. Where should one search when confronted with outbreaks of nosocomial infection? Am J Infect Control. 2006;34:603-605.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 31]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
45.  Hellmann M, Mehta SD, Bishai DM, Mears SC, Zenilman JM. The estimated magnitude and direct hospital costs of prosthetic joint infections in the United States, 1997 to 2004. J Arthroplasty. 2010;25:766-771.e1.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 33]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
46.  Noimark S, Dunnill CW, Wilson M, Parkin IP. The role of surfaces in catheter-associated infections. Chem Soc Rev. 2009;38:3435-3448.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 171]  [Cited by in F6Publishing: 174]  [Article Influence: 11.6]  [Reference Citation Analysis (0)]
47.  Rosenthal VD, Maki DG, Jamulitrat S, Medeiros EA, Todi SK, Gomez DY, Leblebicioglu H, Abu Khader I, Miranda Novales MG, Berba R. International Nosocomial Infection Control Consortium (INICC) report, data summary for 2003-2008, issued June 2009. Am J Infect Control. 2010;38:95-104.e2.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 255]  [Cited by in F6Publishing: 262]  [Article Influence: 18.7]  [Reference Citation Analysis (0)]
48.  Trautner BW, Darouiche RO. Catheter-associated infections: pathogenesis affects prevention. Arch Intern Med. 2004;164:842-850.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 220]  [Cited by in F6Publishing: 190]  [Article Influence: 9.5]  [Reference Citation Analysis (0)]
49.  von Eiff C, Jansen B, Kohnen W, Becker K. Infections associated with medical devices: pathogenesis, management and prophylaxis. Drugs. 2005;65:179-214.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 315]  [Cited by in F6Publishing: 260]  [Article Influence: 13.7]  [Reference Citation Analysis (0)]
50.  Kropec A, Huebner J, Riffel M, Bayer U, Benzing A, Geiger K, Daschner FD. Exogenous or endogenous reservoirs of nosocomial Pseudomonas aeruginosa and Staphylococcus aureus infections in a surgical intensive care unit. Intensive Care Med. 1993;19:161-165.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 45]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
51.  Allen KD, Green HT. Hospital outbreak of multi-resistant Acinetobacter anitratus: an airborne mode of spread? J Hosp Infect. 1987;9:110-119.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 131]  [Cited by in F6Publishing: 135]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
52.  Shiomori T, Miyamoto H, Makishima K, Yoshida M, Fujiyoshi T, Udaka T, Inaba T, Hiraki N. Evaluation of bedmaking-related airborne and surface methicillin-resistant Staphylococcus aureus contamination. J Hosp Infect. 2002;50:30-35.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 142]  [Cited by in F6Publishing: 147]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
53.  Wilson RD, Huang SJ, McLean AS. The correlation between airborne methicillin-resistant Staphylococcus aureus with the presence of MRSA colonized patients in a general intensive care unit. Anaesth Intensive Care. 2004;32:202-209.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Beggs CB. The airborne transmission of infection in hospital buildings: fact or fiction? Indoor Built Environ. 2003;12:9-18.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Srikanth P, Sudharsanam S, Steinberg R. Bio-aerosols in indoor environment: composition, health effects and analysis. Indian J Med Microbiol. 2008;26:302-312.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 101]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
56.  Nicas M, Sun G. An integrated model of infection risk in a health-care environment. Risk Anal. 2006;26:1085-1096.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 98]  [Cited by in F6Publishing: 44]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
57.  Borkow G, Gabbay J. Biocidal textiles can help fight nosocomial infections. Med Hypotheses. 2008;70:990-994.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 112]  [Cited by in F6Publishing: 117]  [Article Influence: 6.9]  [Reference Citation Analysis (0)]
58.  Neely AN, Maley MP. Survival of enterococci and staphylococci on hospital fabrics and plastic. J Clin Microbiol. 2000;38:724-726.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Neely AN, Orloff MM. Survival of some medically important fungi on hospital fabrics and plastics. J Clin Microbiol. 2001;39:3360-3361.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 88]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
60.  Noble WA. Dispersal of microorganisms from skin. Microbiology of human skin. 2nd ed. London: Lloyd-Luke Ltd 1981; 77-85.  [PubMed]  [DOI]  [Cited in This Article: ]
61.  Coronel D, Escarment J, Boiron A, Dusseau JY, Renaud F, Bret M, Freney J. Infection et contamination bacterienne de l'environnement des patients: les draps. Reanimation. 2001;10S:43-44.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Reiss-Levy E, McAllister E. Pillows spread methicillin-resistant staphylococci. Med J Aust. 1979;1:92.  [PubMed]  [DOI]  [Cited in This Article: ]
63.  Ndawula EM, Brown L. Mattresses as reservoirs of epidemic methicillin-resistant Staphylococcus aureus. Lancet. 1991;337:488.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 44]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
64.  Gabbay J, Borkow G, Mishal J, Magen E, Zatcoff R, Shemer-Avni Y. Copper oxide impregnated textiles with potent biocidal activities. J Ind Text. 2006;35:323-335.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Malnick S, Bardenstein R, Huszar M, Gabbay J, Borkow G. Pyjamas and sheets as a potential source of nosocomial pathogens. J Hosp Infect. 2008;70:89-92.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 30]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
66.  GREENE VW, BOND RG, MICHAELSEN GS. Air handling systems must be planned to reduce the spread of infection. Mod Hosp. 1960;95:136-144.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  Solberg CO. A study of carriers of Staphylococcus aureus with special regard to quantitative bacterial estimations. Acta Med Scand Suppl. 1965;436:1-96.  [PubMed]  [DOI]  [Cited in This Article: ]
68.  Rutala WA, Katz EB, Sherertz RJ, Sarubbi FA. Environmental study of a methicillin-resistant Staphylococcus aureus epidemic in a burn unit. J Clin Microbiol. 1983;18:683-688.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Noble WC, Davies RR. Studies on the dispersal of staphylococci. J Clin Pathol. 1965;18:16-19.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 61]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
70.  Weber DJ, Rutala WA, Miller MB, Huslage K, Sickbert-Bennett E. Role of hospital surfaces in the transmission of emerging health care-associated pathogens: norovirus, Clostridium difficile, and Acinetobacter species. Am J Infect Control. 2010;38:S25-S33.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 488]  [Cited by in F6Publishing: 473]  [Article Influence: 33.8]  [Reference Citation Analysis (0)]
71.  Sabino R, Sampaio P, Carneiro C, Rosado L, Pais C. Isolates from hospital environments are the most virulent of the Candida parapsilosis complex. BMC Microbiol. 2011;11:180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 29]  [Article Influence: 2.2]  [Reference Citation Analysis (0)]
72.  Young JM, Naqvi M, Richards L. Microbial contamination of hospital bed handsets. Am J Infect Control. 2005;33:170-174.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 21]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
73.  Hota B. Contamination, disinfection, and cross-colonization: are hospital surfaces reservoirs for nosocomial infection? Clin Infect Dis. 2004;39:1182-1189.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 431]  [Cited by in F6Publishing: 425]  [Article Influence: 21.3]  [Reference Citation Analysis (0)]
74.  Lu PL, Siu LK, Chen TC, Ma L, Chiang WG, Chen YH, Lin SF, Chen TP. Methicillin-resistant Staphylococcus aureus and Acinetobacter baumannii on computer interface surfaces of hospital wards and association with clinical isolates. BMC Infect Dis. 2009;9:164.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 37]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
75.  Mutters R, Nonnenmacher C, Susin C, Albrecht U, Kropatsch R, Schumacher S. Quantitative detection of Clostridium difficile in hospital environmental samples by real-time polymerase chain reaction. J Hosp Infect. 2009;71:43-48.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 50]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
76.  Boyce JM, Potter-Bynoe G, Chenevert C, King T. Environmental contamination due to methicillin-resistant Staphylococcus aureus: possible infection control implications. Infect Control Hosp Epidemiol. 1997;18:622-627.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 166]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
77.  Goodman ER, Platt R, Bass R, Onderdonk AB, Yokoe DS, Huang SS. Impact of an environmental cleaning intervention on the presence of methicillin-resistant Staphylococcus aureus and vancomycin-resistant enterococci on surfaces in intensive care unit rooms. Infect Control Hosp Epidemiol. 2008;29:593-599.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 227]  [Cited by in F6Publishing: 237]  [Article Influence: 14.8]  [Reference Citation Analysis (0)]
78.  Bhalla A, Pultz NJ, Gries DM, Ray AJ, Eckstein EC, Aron DC, Donskey CJ. Acquisition of nosocomial pathogens on hands after contact with environmental surfaces near hospitalized patients. Infect Control Hosp Epidemiol. 2004;25:164-167.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 307]  [Cited by in F6Publishing: 314]  [Article Influence: 34.9]  [Reference Citation Analysis (0)]
79.  Eckstein BC, Adams DA, Eckstein EC, Rao A, Sethi AK, Yadavalli GK, Donskey CJ. Reduction of Clostridium Difficile and vancomycin-resistant Enterococcus contamination of environmental surfaces after an intervention to improve cleaning methods. BMC Infect Dis. 2007;7:61.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 181]  [Article Influence: 10.6]  [Reference Citation Analysis (0)]
80.  Hayden MK, Bonten MJ, Blom DW, Lyle EA, van de Vijver DA, Weinstein RA. Reduction in acquisition of vancomycin-resistant enterococcus after enforcement of routine environmental cleaning measures. Clin Infect Dis. 2006;42:1552-1560.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 301]  [Cited by in F6Publishing: 307]  [Article Influence: 17.1]  [Reference Citation Analysis (0)]
81.  Gustafson TL, Kobylik B, Hutcheson RH, Schaffner W. Protective effect of anticholinergic drugs and psyllium in a nosocomial outbreak of Norwalk gastroenteritis. J Hosp Infect. 1983;4:367-374.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 13]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
82.  Takahashi A, Yomoda S, Tanimoto K, Kanda T, Kobayashi I, Ike Y. Streptococcus pyogenes hospital-acquired infection within a dermatological ward. J Hosp Infect. 1998;40:135-140.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
83.  Cozad A, Jones RD. Disinfection and the prevention of infectious disease. Am J Infect Control. 2003;31:243-254.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 87]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
84.  Morgan DJ, Liang SY, Smith CL, Johnson JK, Harris AD, Furuno JP, Thom KA, Snyder GM, Day HR, Perencevich EN. Frequent multidrug-resistant Acinetobacter baumannii contamination of gloves, gowns, and hands of healthcare workers. Infect Control Hosp Epidemiol. 2010;31:716-721.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 134]  [Article Influence: 9.6]  [Reference Citation Analysis (0)]
85.  Kramer A, Schwebke I, Kampf G. How long do nosocomial pathogens persist on inanimate surfaces? A systematic review. BMC Infect Dis. 2006;6:130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1617]  [Cited by in F6Publishing: 1376]  [Article Influence: 76.4]  [Reference Citation Analysis (0)]
86.  Hsu J, Abad C, Dinh M, Safdar N. Prevention of endemic healthcare-associated Clostridium difficile infection: reviewing the evidence. Am J Gastroenterol. 2010;105:2327-2329; quiz 2340.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 48]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
87.  Wilson AP, Smyth D, Moore G, Singleton J, Jackson R, Gant V, Jeanes A, Shaw S, James E, Cooper B. The impact of enhanced cleaning within the intensive care unit on contamination of the near-patient environment with hospital pathogens: a randomized crossover study in critical care units in two hospitals. Crit Care Med. 2011;39:651-658.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 77]  [Cited by in F6Publishing: 76]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
88.  Sattar SA. Promises and pitfalls of recent advances in chemical means of preventing the spread of nosocomial infections by environmental surfaces. Am J Infect Control. 2010;38:S34-S40.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 48]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
89.  Carling PC, Bartley JM. Evaluating hygienic cleaning in health care settings: what you do not know can harm your patients. Am J Infect Control. 2010;38:S41-S50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 140]  [Cited by in F6Publishing: 130]  [Article Influence: 9.3]  [Reference Citation Analysis (0)]
90.  Borkow G, Gabbay J. Copper as a biocidal tool. Curr Med Chem. 2005;12:2163-2175.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 454]  [Cited by in F6Publishing: 326]  [Article Influence: 17.2]  [Reference Citation Analysis (0)]
91.  Borkow G, Gabbay J. Copper, an ancient remedy returning to fight microbial, fungal and viral infections. Curr Chem Biol. 2009;3:272-278.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 43]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
92.  Cross JB, Currier RP, Torraco DJ, Vanderberg LA, Wagner GL, Gladen PD. Killing of bacillus spores by aqueous dissolved oxygen, ascorbic acid, and copper ions. Appl Environ Microbiol. 2003;69:2245-2252.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 72]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
93.  Gant VA, Wren MW, Rollins MS, Jeanes A, Hickok SS, Hall TJ. Three novel highly charged copper-based biocides: safety and efficacy against healthcare-associated organisms. J Antimicrob Chemother. 2007;60:294-299.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 53]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
94.  Hall TJ, Wren MW, Jeanes A, Gant VA. A comparison of the antibacterial efficacy and cytotoxicity to cultured human skin cells of 7 commercial hand rubs and Xgel, a new copper-based biocidal hand rub. Am J Infect Control. 2009;37:322-326.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 18]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
95.  Mato Rodriguez L, Alatossava T. Effects of copper on germination, growth and sporulation of Clostridium tyrobutyricum. Food Microbiol. 2010;27:434-437.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 19]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
96.  Weaver L, Michels HT, Keevil CW. Potential for preventing spread of fungi in air-conditioning systems constructed using copper instead of aluminium. Lett Appl Microbiol. 2010;50:18-23.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 67]  [Cited by in F6Publishing: 72]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
97.  Weaver L, Michels HT, Keevil CW. Survival of Clostridium difficile on copper and steel: futuristic options for hospital hygiene. J Hosp Infect. 2008;68:145-151.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 121]  [Cited by in F6Publishing: 108]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
98.  Wheeldon LJ, Worthington T, Lambert PA, Hilton AC, Lowden CJ, Elliott TS. Antimicrobial efficacy of copper surfaces against spores and vegetative cells of Clostridium difficile: the germination theory. J Antimicrob Chemother. 2008;62:522-525.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 90]  [Cited by in F6Publishing: 78]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
99.  Borkow G, Sidwell RW, Smee DF, Barnard DL, Morrey JD, Lara-Villegas HH, Shemer-Avni Y, Gabbay J. Neutralizing viruses in suspensions by copper oxide-based filters. Antimicrob Agents Chemother. 2007;51:2605-2607.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 50]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
100.  Espírito Santo C, Taudte N, Nies DH, Grass G. Contribution of copper ion resistance to survival of Escherichia coli on metallic copper surfaces. Appl Environ Microbiol. 2008;74:977-986.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 204]  [Cited by in F6Publishing: 155]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
101.  Espírito Santo C, Lam EW, Elowsky CG, Quaranta D, Domaille DW, Chang CJ, Grass G. Bacterial killing by dry metallic copper surfaces. Appl Environ Microbiol. 2011;77:794-802.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 334]  [Cited by in F6Publishing: 269]  [Article Influence: 19.2]  [Reference Citation Analysis (0)]
102.  Grass G, Rensing C, Solioz M. Metallic copper as an antimicrobial surface. Appl Environ Microbiol. 2011;77:1541-1547.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 895]  [Cited by in F6Publishing: 721]  [Article Influence: 51.5]  [Reference Citation Analysis (0)]
103.  Ohsumi Y, Kitamoto K, Anraku Y. Changes induced in the permeability barrier of the yeast plasma membrane by cupric ion. J Bacteriol. 1988;170:2676-2682.  [PubMed]  [DOI]  [Cited in This Article: ]
104.  Quaranta D, Krans T, Espírito Santo C, Elowsky CG, Domaille DW, Chang CJ, Grass G. Mechanisms of contact-mediated killing of yeast cells on dry metallic copper surfaces. Appl Environ Microbiol. 2011;77:416-426.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 120]  [Article Influence: 8.6]  [Reference Citation Analysis (0)]
105.  Santo CE, Morais PV, Grass G. Isolation and characterization of bacteria resistant to metallic copper surfaces. Appl Environ Microbiol. 2010;76:1341-1348.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 95]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
106.  Borkow G, Covington CY, Gautam B, Anzala O, Oyugi J, Juma M, Abdullah MS. Prevention of human immunodeficiency virus breastmilk transmission with copper oxide: proof-of-concept study. Breastfeed Med. 2011;6:165-170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 12]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
107.  Valko M, Morris H, Cronin MT. Metals, toxicity and oxidative stress. Curr Med Chem. 2005;12:1161-1208.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3118]  [Cited by in F6Publishing: 3109]  [Article Influence: 163.6]  [Reference Citation Analysis (2)]
108.  Borkow G, Gabbay J. Putting copper into action: copper-impregnated products with potent biocidal activities. FASEB J. 2004;18:1728-1730.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Borkow G, Gabbay J. Endowing textiles with permanent potent biocidal properties by impregnating them with copper oxide. JTATM. 2006;5:1-3.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  Borkow G, Lara HH, Covington CY, 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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 49]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
111.  Borkow G, Zhou SS, Page T, Gabbay J. A novel anti-influenza copper oxide containing respiratory face mask. PLoS One. 2010;5:e11295.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 223]  [Cited by in F6Publishing: 240]  [Article Influence: 17.1]  [Reference Citation Analysis (0)]
112.  Borkow G, Okon-Levy N, Gabbay J. Copper oxide impregnated wound dressing: biocidal and safety studies. Wounds. 2010;22:301–310.  [PubMed]  [DOI]  [Cited in This Article: ]
113.  Zatcoff RC, Smith MS, Borkow G. Treatment of tinea pedis with socks containing copper-oxide impregnated fibers. Foot (Edinb). 2008;18:136-141.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 40]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
114.  Mumcuoglu KY, Gabbay J, Borkow G. Copper oxide-impregnated fabrics for the control of house dust mites. Int J Pest Manage. 2008;54:235-240.  [PubMed]  [DOI]  [Cited in This Article: ]
115.  O'Hanlon SJ, Enright MC. A novel bactericidal fabric coating with potent in vitro activity against meticillin-resistant Staphylococcus aureus (MRSA). Int J Antimicrob Agents. 2009;33:427-431.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 22]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
116.  White WC, Bellfield R, Ellis J, Vandendaele IP. Controlling the spread of nosocomial infections in hospital wards by the use of antimicrobials on textiles, facilities and devices.  MEDTEX 07 - Fourth International Conference and Exhibition on Healthcare and Medical Textiles; 2007 Jul 16-18; Bolton, UK.  [PubMed]  [DOI]  [Cited in This Article: ]
117.  Bischof Vukušić S, Flinčec Grgac S, Budimir A, Kalenić S. Cotton textiles modified with citric acid as efficient anti-bacterial agent for prevention of nosocomial infections. Croat Med J. 2011;52:68-75.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 30]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
118.  Lansdown AB. Silver in health care: antimicrobial effects and safety in use. Curr Probl Dermatol. 2006;33:17-34.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 464]  [Cited by in F6Publishing: 398]  [Article Influence: 22.1]  [Reference Citation Analysis (0)]
119.  Groß R, Hübner N, Assadian O, Jibson B, Kramer A; Working Section for Clinical Antiseptic of the German Society for Hospital Hygiene. Pilot study on the microbial contamination of conventional vs. silver-impregnated uniforms worn by ambulance personnel during one week of emergency medical service. GMS Krankenhhyg Interdiszip. 2010;5:Doc09.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 5]  [Reference Citation Analysis (0)]
120.  Yazdankhah SP, Scheie AA, Høiby EA, Lunestad BT, Heir E, Fotland TØ, Naterstad K, Kruse H. Triclosan and antimicrobial resistance in bacteria: an overview. Microb Drug Resist. 2006;12:83-90.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 258]  [Cited by in F6Publishing: 219]  [Article Influence: 12.2]  [Reference Citation Analysis (0)]
121.  Huang J, Murata H, Koepsel RR, Russell AJ, Matyjaszewski K. Antibacterial polypropylene via surface-initiated atom transfer radical polymerization. Biomacromolecules. 2007;8:1396-1399.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 269]  [Cited by in F6Publishing: 214]  [Article Influence: 12.6]  [Reference Citation Analysis (0)]
122.  Takai K, Ohtsuka T, Senda Y, Nakao M, Yamamoto K, Matsuoka J, Hirai Y. Antibacterial properties of antimicrobial-finished textile products. Microbiol Immunol. 2002;46:75-81.  [PubMed]  [DOI]  [Cited in This Article: ]
123.  Ranjbar-Mohammadi M, Arami M, Bahrami H, Mazaheri F, Mahmoodi NM. Grafting of chitosan as a biopolymer onto wool fabric using anhydride bridge and its antibacterial property. Colloids Surf B Biointerfaces. 2010;76:397-403.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 61]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
124.  Kalyon BD, Olgun U. Antibacterial efficacy of triclosan-incorporated polymers. Am J Infect Control. 2001;29:124-125.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 104]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
125.  Renaud FNR, Doré J, Freney HJ, Coronel B, Dusseau JY. Evaluation of antibacterial properties of a textile product with antimicrobial finish in a hospital environment. J Ind Text. 2006;36:89-94.  [PubMed]  [DOI]  [Cited in This Article: ]
126.  United States Environmental Protection Agency. Reregistration Eligibility Decision (RED) for Coppers.  2009; p9. Available from: http://www.epa.gov/oppsrrd1/REDs/copper_red_amend.pdf.  [PubMed]  [DOI]  [Cited in This Article: ]
127.  de Veer I, Wilke K, Rüden H. [Bacterial reducing qualities of copper-containing and non-copper-containing materials. I. Contamination and sedimentation in humid and dry conditions]. Zentralbl Hyg Umweltmed. 1993;195:66-87.  [PubMed]  [DOI]  [Cited in This Article: ]
128.  De Veer I, Wilke K, Rüden H. [Bacteria-reducing properties of copper-containing and non-copper-containing materials. II. Relationship between microbiocide effect of copper-containing materials and copper ion concentration after contamination with moist and dry hands]. Zentralbl Hyg Umweltmed. 1994;195:516-528.  [PubMed]  [DOI]  [Cited in This Article: ]
129.  Faúndez G, Troncoso M, Navarrete P, Figueroa G. Antimicrobial activity of copper surfaces against suspensions of Salmonella enterica and Campylobacter jejuni. BMC Microbiol. 2004;4:19.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 187]  [Cited by in F6Publishing: 155]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
130.  Mehtar S, Wiid I, Todorov SD. The antimicrobial activity of copper and copper alloys against nosocomial pathogens and Mycobacterium tuberculosis isolated from healthcare facilities in the Western Cape: an in-vitro study. J Hosp Infect. 2008;68:45-51.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 140]  [Cited by in F6Publishing: 142]  [Article Influence: 8.4]  [Reference Citation Analysis (0)]
131.  Michels HT, Noyce JO, Keevil CW. Effects of temperature and humidity on the efficacy of methicillin-resistant Staphylococcus aureus challenged antimicrobial materials containing silver and copper. Lett Appl Microbiol. 2009;49:191-195.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 116]  [Cited by in F6Publishing: 95]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
132.  Mikolay A, Huggett S, Tikana L, Grass G, Braun J, Nies DH. Survival of bacteria on metallic copper surfaces in a hospital trial. Appl Microbiol Biotechnol. 2010;87:1875-1879.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 107]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
133.  Noyce JO, Michels H, Keevil CW. Use of copper cast alloys to control Escherichia coli O157 cross-contamination during food processing. Appl Environ Microbiol. 2006;72:4239-4244.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 151]  [Cited by in F6Publishing: 156]  [Article Influence: 8.7]  [Reference Citation Analysis (0)]
134.  Noyce JO, Michels H, Keevil CW. 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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 265]  [Cited by in F6Publishing: 275]  [Article Influence: 15.3]  [Reference Citation Analysis (0)]
135.  Noyce JO, Michels H, Keevil CW. Inactivation of influenza A virus on copper versus stainless steel surfaces. Appl Environ Microbiol. 2007;73:2748-2750.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 200]  [Cited by in F6Publishing: 204]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
136.  Robine E, Boulangé-Petermann L, Derangère D. Assessing bactericidal properties of materials: the case of metallic surfaces in contact with air. J Microbiol Methods. 2002;49:225-234.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 33]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
137.  Warnes SL, Green SM, Michels HT, Keevil CW. Biocidal efficacy of copper alloys against pathogenic enterococci involves degradation of genomic and plasmid DNAs. Appl Environ Microbiol. 2010;76:5390-5401.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 89]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
138.  Weaver L, Noyce JO, Michels HT, Keevil CW. Potential action of copper surfaces on meticillin-resistant Staphylococcus aureus. J Appl Microbiol. 2010;109:2200-2205.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 91]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
139.  Wilks SA, Michels H, Keevil CW. The survival of Escherichia coli O157 on a range of metal surfaces. Int J Food Microbiol. 2005;105:445-454.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 231]  [Cited by in F6Publishing: 190]  [Article Influence: 10.0]  [Reference Citation Analysis (0)]
140.  Wilks SA, Michels HT, Keevil CW. Survival of Listeria monocytogenes Scott A on metal surfaces: implications for cross-contamination. Int J Food Microbiol. 2006;111:93-98.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 137]  [Cited by in F6Publishing: 140]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
141.  Sharan R, Chhibber S, Reed RH. Inactivation and sub-lethal injury of salmonella typhi, salmonella typhimurium and vibrio cholerae in copper water storage vessels. BMC Infect Dis. 2011;11:204.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
142.  Sharan R, Chhibber S, Reed RH. A murine model to study the antibacterial effect of copper on infectivity of Salmonella enterica serovar Typhimurium. Int J Environ Res Public Health. 2011;8:21-36.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
143.  Sharan R, Chhibber S, Attri S, Reed RH. Inactivation and sub-lethal injury of Escherichia coli in a copper water storage vessel: effect of inorganic and organic constituents. Antonie Van Leeuwenhoek. 2010;98:103-115.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 8]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
144.  Elguindi J, Wagner J, Rensing C. Genes involved in copper resistance influence survival of Pseudomonas aeruginosa on copper surfaces. J Appl Microbiol. 2009;106:1448-1455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 77]  [Cited by in F6Publishing: 81]  [Article Influence: 5.4]  [Reference Citation Analysis (0)]
145.  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.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 72]  [Cited by in F6Publishing: 74]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
146.  Casey AL, Adams D, Karpanen TJ, Lambert PA, Cookson BD, Nightingale P, Miruszenko L, Shillam R, Christian P, Elliott TS. Role of copper in reducing hospital environment contamination. J Hosp Infect. 2010;74:72-77.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 226]  [Cited by in F6Publishing: 231]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
147.  Marais F, Mehtar S, Chalkley L. Antimicrobial efficacy of copper touch surfaces in reducing environmental bioburden in a South African community healthcare facility. J Hosp Infect. 2010;74:80-82.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 56]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
148.  Casey AL, Karpanen TJ, Adams D, Lambert PA, Nightingale P, Miruszenko L, Elliott TS. A comparative study to evaluate surface microbial contamination associated with copper-containing and stainless steel pens used by nurses in the critical care unit. Am J Infect Control. 2011;39:e52-e54.  [PubMed]  [DOI]  [Cited in This Article: ]
149.  Taylor L, Phillips P, Hastings R. Reduction of bacterial contamination in a healthcare environment by silver antimicrobial technology. J Infect Prev. 2009;10:6-12.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 26]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
150.  Goto H, Shimada K, Ikemoto H, Oguri T. Antimicrobial susceptibility of pathogens isolated from more than 10,000 patients with infectious respiratory diseases: a 25-year longitudinal study. J Infect Chemother. 2009;15:347-360.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 26]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
151.  Maillard JY. Bacterial resistance to biocides in the healthcare environment: should it be of genuine concern? J Hosp Infect. 2007;65 Suppl 2:60-72.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 84]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
152.  Maillard JY. Antimicrobial biocides in the healthcare environment: efficacy, usage, policies, and perceived problems. Ther Clin Risk Manag. 2005;1:307-320.  [PubMed]  [DOI]  [Cited in This Article: ]
153.  Aarestrup FM, Hasman H. Susceptibility of different bacterial species isolated from food animals to copper sulphate, zinc chloride and antimicrobial substances used for disinfection. Vet Microbiol. 2004;100:83-89.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 139]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
154.  Hasman H, Kempf I, Chidaine B, Cariolet R, Ersbøll AK, Houe H, Bruun Hansen HC, Aarestrup FM. Copper resistance in Enterococcus faecium, mediated by the tcrB gene, is selected by supplementation of pig feed with copper sulfate. Appl Environ Microbiol. 2006;72:5784-5789.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 91]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
155.  Elguindi J, Moffitt S, Hasman H, Andrade C, Raghavan S, Rensing C. Metallic copper corrosion rates, moisture content, and growth medium influence survival of copper ion-resistant bacteria. Appl Microbiol Biotechnol. 2011;89:1963-1970.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 57]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
156.  Fait G, Broos K, Zrna S, Lombi E, Hamon R. Tolerance of nitrifying bacteria to copper and nickel. Environ Toxicol Chem. 2006;25:2000-2005.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 7]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
157.  Andersen GL, Menkissoglou O, Lindow SE. Occurrence and properties of copper-tolerant strains of pseudomonas syringae isolated from fruit trees in California. Phytopathology. 1991;81:648-656.  [PubMed]  [DOI]  [Cited in This Article: ]
158.  Saeki K, Kunito T, Oyaizu H, Matsumoto S. Relationships between bacterial tolerance levels and forms of copper and zinc in soils. J Environ Qual. 2002;31:1570-1575.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 5]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
159.  Stout JE, Yu VL. Experiences of the first 16 hospitals using copper-silver ionization for Legionella control: implications for the evaluation of other disinfection modalities. Infect Control Hosp Epidemiol. 2003;24:563-568.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 103]  [Cited by in F6Publishing: 105]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
160.  Lin YE, Stout JE, Yu VL. Controlling Legionella in hospital drinking water: an evidence-based review of disinfection methods. Infect Control Hosp Epidemiol. 2011;32:166-173.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 123]  [Cited by in F6Publishing: 100]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
161.  Pedro-Botet ML, Sanchez I, Sabria M, Sopena N, Mateu L, García-Núñez M, Rey-Joly C. Impact of copper and silver ionization on fungal colonization of the water supply in health care centers: implications for immunocompromised patients. Clin Infect Dis. 2007;45:84-86.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 17]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
162.  Olesevich M, Kennedy A. Emergence of community-acquired methicillin-resistant Staphylococcus aureus soft tissue infections. J Pediatr Surg. 2007;42:765-768.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 19]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
163.  Patyi M, Varga É, Kristóf K. Curiosities of the methicillin-resistant Staphyslococcus aureus survey - possibility of pseudo-outbreak and transmission to household contacts. Acta Microbiol Immunol Hung. 2011;58:135-144.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 2]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
164.  Huang HI, Shih HY, Lee CM, Yang TC, Lay JJ, Lin YE. In vitro efficacy of copper and silver ions in eradicating Pseudomonas aeruginosa, Stenotrophomonas maltophilia and Acinetobacter baumannii: implications for on-site disinfection for hospital infection control. Water Res. 2008;42:73-80.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 48]  [Cited by in F6Publishing: 32]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
165.  Cortes P, Atria AM, Contreras M, Garland MT, Peña O, Corsini G. Magnetic properties and antibacterial activity of tetranuclear copper complexes bridged by oxo group. J Chil Chem Soc. 2006;51:957-960.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 2]  [Article Influence: 0.1]  [Reference Citation Analysis (0)]
166.  Moore G, Hall TJ, Wilson AP, Gant VA. The efficacy of the inorganic copper-based biocide CuWB50 is compromised by hard water. Lett Appl Microbiol. 2008;46:655-660.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 10]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
167.  Podunavac-Kuzmanović S, Cvetković D. Antimicrobial investigations of copper(II) complexes with some 1-benzylbenzimidazole derivatives. Rev Roum Chim. 2010;55:363-367.  [PubMed]  [DOI]  [Cited in This Article: ]
168.  Arslan H, Duran N, Borekci G, Koray Ozer C, Akbay C. Antimicrobial activity of some thiourea derivatives and their nickel and copper complexes. Molecules. 2009;14:519-527.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 107]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
169.  Sun LM, Zhang CL, Li P. Characterization, antimicrobial activity, and mechanism of a high-performance (-)-epigallocatechin-3-gallate (EGCG)-CuII/polyvinyl alcohol (PVA) nanofibrous membrane. J Agric Food Chem. 2011;59:5087-5092.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 24]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
170.  Ruparelia JP, Chatterjee AK, Duttagupta SP, Mukherji S. Strain specificity in antimicrobial activity of silver and copper nanoparticles. Acta Biomater. 2008;4:707-716.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1302]  [Cited by in F6Publishing: 909]  [Article Influence: 56.8]  [Reference Citation Analysis (0)]
171.  Patel MN, Dosi PA, Bhatt BS, Thakkar VR. Synthesis, characterization, antibacterial activity, SOD mimic and interaction with DNA of drug based copper(II) complexes. Spectrochim Acta A Mol Biomol Spectrosc. 2011;78:763-770.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 26]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
172.  Revanasiddappa HD, Vijaya B, Shiva Kumar L, Shiva Prasad K. Synthesis, characterization and antimicrobial activity of Cu(ii), Co(ii), Ni(ii) and Mn(ii) complexes with desipramine. World J Chem. 2010;5:18-25.  [PubMed]  [DOI]  [Cited in This Article: ]
173.  Suksrichavalit T, Prachayasittikul S, Nantasenamat C, Isarankura-Na-Ayudhya C, Prachayasittikul V. Copper complexes of pyridine derivatives with superoxide scavenging and antimicrobial activities. Eur J Med Chem. 2009;44:3259-3265.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 122]  [Cited by in F6Publishing: 129]  [Article Influence: 8.6]  [Reference Citation Analysis (0)]
174.  Chohan ZH, Iqbal MS, Aftab SK, Rauf A. Antibacterial dimeric copper(II) complexes with chromone-derived compounds. J Enzyme Inhib Med Chem. 2012;27:223-231.  [PubMed]  [DOI]  [Cited in This Article: ]
175.  Xie LJ, Chu W, Sun JH, Wu P, Tong DG. Synthesis of copper oxide vegetable sponges and their antibacterial, electrochemical and photocatalytic performance. J Mater Sci. 2011;46:2179-2184.  [PubMed]  [DOI]  [Cited in This Article: ]
176.  Chandra S, Raizada S, Tyagi M, Sharma PK. Spectroscopic and biological approach of Ni(II) and Cu(II) complexes of 2-pyridinecarboxaldehyde thiosemicarbazone. Spectrochim Acta A Mol Biomol Spectrosc. 2008;69:816-821.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 32]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
177.  Agwara MO, Ndifon PT, Ndosiri NB, Paboudam AG, Yufanyi DM, Mohamadou A. Synthesis, characterisation and antimicrobial activities of cobalt(II), copper(II) and zinc(II) mixed-ligand complexes containing 1,10-phenanthroline and 2,2’-bipyridine. Bull Chem Soc Ethiop. 2010;24:383-389.  [PubMed]  [DOI]  [Cited in This Article: ]
178.  Sani RK, Peyton BM, Brown LT. Copper-induced inhibition of growth of Desulfovibrio desulfuricans G20: assessment of its toxicity and correlation with those of zinc and lead. Appl Environ Microbiol. 2001;67:4765-4772.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 134]  [Cited by in F6Publishing: 96]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
179.  Hu YH, Dang W, Liu CS, Sun L. Analysis of the effect of copper on the virulence of a pathogenic Edwardsiella tarda strain. Lett Appl Microbiol. 2010;50:97-103.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 14]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
180.  Colak A, Terzi Ü, Col M, Karaoglu ŞA, Karaböcek S, Küçükdumlu A, Ayaz FA. DNA binding, antioxidant and antimicrobial activities of homo- and heteronuclear copper(II) and nickel(II) complexes with new oxime-type ligands. Eur J Med Chem. 2010;45:5169-5175.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 53]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
181.  Nie Y, Kalapos C, Nie X, Murphy M, Hussein R, Zhang J. Superhydrophilicity and antibacterial property of a Cu-dotted oxide coating surface. Ann Clin Microbiol Antimicrob. 2010;9:25.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 29]  [Cited by in F6Publishing: 19]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
182.  Molteni C, Abicht HK, Solioz M. Killing of bacteria by copper surfaces involves dissolved copper. Appl Environ Microbiol. 2010;76:4099-4101.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 116]  [Cited by in F6Publishing: 93]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
183.  Ibrahim SA, Yang H, Seo CW. Antimicrobial activity of lactic acid and copper on growth of Salmonella and Escherichia coli O157:H7 in laboratory medium and carrot juice. Food Chem. 2008;109:137-143.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 75]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
184.  Ren G, Hu D, Cheng EW, Vargas-Reus MA, Reip P, Allaker RP. Characterisation of copper oxide nanoparticles for antimicrobial applications. Int J Antimicrob Agents. 2009;33:587-590.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 991]  [Cited by in F6Publishing: 663]  [Article Influence: 44.2]  [Reference Citation Analysis (0)]
185.  Malachová K, Praus P, Rybková Z, Kozák O. Antibacterial and antifungal activities of silver, copper and zinc montmorillonites. Appl Clay Sci. 2011;53:642-645.  [PubMed]  [DOI]  [Cited in This Article: ]
186.  Qiu JH, Zhang YW, Zhang YT, Zhang HQ, Liu JD. Synthesis and antibacterial activity of copper-immobilized membrane comprising grafted poly(4-vinylpyridine) chains. J Colloid Interface Sci. 2011;354:152-159.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 60]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
187.  Zhang W, Zhang Y, Ji J, Yan Q, Huang A, Chu PK. Antimicrobial polyethylene with controlled copper release. J Biomed Mater Res A. 2007;83:838-844.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 47]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
188.  Grey B, Steck TR. Concentrations of copper thought to be toxic to Escherichia coli can induce the viable but nonculturable condition. Appl Environ Microbiol. 2001;67:5325-5327.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 48]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
189.  Torres A, Ruales C, Pulgarin C, Aimable A, Bowen P, Sarria V, Kiwi J. Innovative high-surface-area CuO pretreated cotton effective in bacterial inactivation under visible light. ACS Appl Mater Interfaces. 2010;2:2547-2552.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 21]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
190.  Sudha VB, Singh KO, Prasad SR, Venkatasubramanian P. Killing of enteric bacteria in drinking water by a copper device for use in the home: laboratory evidence. Trans R Soc Trop Med Hyg. 2009;103:819-822.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 33]  [Cited by in F6Publishing: 37]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
191.  Yates HM, Brook LA, Ditta IB, Evans P, Foster HA, Sheel DW, Steele A. Photo-induced self-cleaning and biocidal behaviour of titania and copper oxide multilayers. J Photochem Photobiol A. 2008;197:197-205.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 70]  [Cited by in F6Publishing: 70]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
192.  Nan L, Liu Y, Lü M, Yang K. Study on antibacterial mechanism of copper-bearing austenitic antibacterial stainless steel by atomic force microscopy. J Mater Sci Mater Med. 2008;19:3057-3062.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 45]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
193.  Rosu T, Pahontu E, Pasculescu S, Georgescu R, Stanica N, Curaj A, Popescu A, Leabu M. Synthesis, characterization antibacterial and antiproliferative activity of novel Cu(II) and Pd(II) complexes with 2-hydroxy-8-R-tricyclo[7.3.1.0.(2,7)]tridecane-13-one thiosemicarbazone. Eur J Med Chem. 2010;45:1627-1634.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 69]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
194.  Chohan ZH, Arif M, Rashid A. Copper (II) and zinc (ii) metal-based salicyl-, furanyl-, thienyl- and pyrrolyl-derived ONNO, NNNO, ONNS & amp; NNNS donor asymmetrically mixed schiff-bases with antibacterial and antifungal potentials. J Enzyme Inhib Med Chem. 2008;23:785-796.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
195.  Raman N, Sobha S. Synthesis, characterization, DNA interaction and antimicrobial screening of isatin-based polypyridyl mixed ligand Cu(II) and Zn(II) complexes. J Serb Chem Soc. 2010;75:773-788.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 37]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
196.  Casari E, Ferrario A, Montanelli A. Prolonged effect of two combined methods for Legionella disinfection in a hospital water system. Ann Ig. 2007;19:525-532.  [PubMed]  [DOI]  [Cited in This Article: ]
197.  Stout JE, Lin YS, Goetz AM, Muder RR. Controlling Legionella in hospital water systems: experience with the superheat-and-flush method and copper-silver ionization. Infect Control Hosp Epidemiol. 1998;19:911-914.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 33]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
198.  Landeen LK, Yahya MT, Gerba CP. Efficacy of copper and silver ions and reduced levels of free chlorine in inactivation of Legionella pneumophila. Appl Environ Microbiol. 1989;55:3045-3050.  [PubMed]  [DOI]  [Cited in This Article: ]
199.  Russell SM. The effect of an acidic, copper sulfate-based commercial sanitizer on indicator, pathogenic, and spoilage bacteria associated with broiler chicken carcasses when applied at various intervention points during poultry processing. Poult Sci. 2008;87:1435-1440.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 20]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
200.  López-Carballo G, Hernández-Muñoz P, Gavara R, Ocio MJ. Photoactivated chlorophyllin-based gelatin films and coatings to prevent microbial contamination of food products. Int J Food Microbiol. 2008;126:65-70.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 59]  [Cited by in F6Publishing: 46]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
201.  Keyhani E, Abdi-Oskouei F, Attar F, Keyhani J. DNA strand breaks by metal-induced oxygen radicals in purified Salmonella typhimurium DNA. Ann N Y Acad Sci. 2006;1091:52-64.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 11]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
202.  Chandra S, Jain D, Sharma AK, Sharma P. Coordination modes of a schiff base pentadentate derivative of 4-aminoantipyrine with cobalt(II), nickel(II) and copper(II) metal ions: synthesis, spectroscopic and antimicrobial studies. Molecules. 2009;14:174-190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 121]  [Cited by in F6Publishing: 127]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
203.  Raman N, Baskaran T, Selvan A, Jeyamurugan R. DNA interaction and antimicrobial studies of novel copper (II) complex having ternary Schiff base. J Iran Chem Res. 2008;1:129-139.  [PubMed]  [DOI]  [Cited in This Article: ]
204.  Chohan ZH, Shaikh AU, Supuran CT. In-vitro antibacterial, antifungal and cytotoxic activity of cobalt (II), copper (II), nickel (II) and zinc (II) complexes with furanylmethyl- and thienylmethyl-dithiolenes: [1, 3-dithiole- 2-one and 1,3-dithiole-2-thione]. J Enzyme Inhib Med Chem. 2006;21:733-740.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 26]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
205.  Carson KC, Bartlett JG, Tan TJ, Riley TV. In Vitro susceptibility of methicillin-resistant Staphylococcus aureus and methicillin-susceptible Staphylococcus aureus to a new antimicrobial, copper silicate. Antimicrob Agents Chemother. 2007;51:4505-4507.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 12]  [Cited by in F6Publishing: 13]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
206.  Luna VA, Hall TJ, King DS, Cannons AC. Susceptibility of 169 USA300 methicillin-resistant Staphylococcus aureus isolates to two copper-based biocides, CuAL42 and CuWB50. J Antimicrob Chemother. 2010;65:939-941.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 9]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
207.  Neel EA, Ahmed I, Pratten J, Nazhat SN, Knowles JC. Characterisation of antibacterial copper releasing degradable phosphate glass fibres. Biomaterials. 2005;26:2247-2254.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 216]  [Cited by in F6Publishing: 142]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
208.  Mulligan AM, Wilson M, Knowles JC. The effect of increasing copper content in phosphate-based glasses on biofilms of Streptococcus sanguis. Biomaterials. 2003;24:1797-1807.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 87]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
209.  Bhattacharya K, Niyogi SK, Choudhuri SK. Role of a novel copper chelate in modulation of resistance by time and dose-dependent potential on the growth of tetracycline-resistant Vibrio cholerae O1. Int J Antimicrob Agents. 2011;38:182-183.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 4]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
210.  Bellí N, Marín S, Sanchis V, Ramos AJ. Impact of fungicides on Aspergillus carbonarius growth and ochratoxin A production on synthetic grape-like medium and on grapes. Food Addit Contam. 2006;23:1021-1029.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 35]  [Article Influence: 2.1]  [Reference Citation Analysis (0)]
211.  Kumbhar AS, Padhye SB, Saraf AP, Mahajan HB, Chopade BA, West DX. Novel broad-spectrum metal-based antifungal agents. Correlations amongst the structural and biological properties of copper (II) 2-acetylpyridine N4-dialkylthiosemicarbazones. Biol Met. 1991;4:141-143.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 20]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
212.  Gershon H, Clarke DD, Gershon M. Synergistic antifungal action of 8-quinolinol and its bischelate with copper(II) and with mixed ligand chelates composed of copper(II), 8-quinolinol, and aromatic hydroxy acids. J Pharm Sci. 1989;78:975-978.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 21]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
213.  Shakru R, Subhashini NJP, Shivaraj SKK. Synthesis, characterization and antimicrobial studies on Cobalt (II), Nickel (II), Copper (II) and Zinc (II) complexes of N, O, S donor Schiff bases. J Chem Pharm Res. 2010;2:38-46.  [PubMed]  [DOI]  [Cited in This Article: ]
214.  Syed Tajudeen S, Santhalakshmi S, Geetha K. Studies on antimicrobial activity of some hydrazones and its copper complexes. J Pharm Res. 2010;3:2759-2760.  [PubMed]  [DOI]  [Cited in This Article: ]
215.  Lin MY, Huang KJ, Kleven SH. In vitro comparison of the activity of various antifungal drugs against new yeast isolates causing thrush in poultry. Avian Dis. 1989;33:416-421.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 5]  [Article Influence: 0.1]  [Reference Citation Analysis (0)]
216.  Al-Holy MA, Castro LF, Al-Qadiri HM. Inactivation of Cronobacter spp. (Enterobacter sakazakii) in infant formula using lactic acid, copper sulfate and monolaurin. Lett Appl Microbiol. 2010;50:246-251.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 35]  [Cited by in F6Publishing: 37]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
217.  Vagabov VM, Ivanov AY, Kulakovskaya TV, Kulakovskaya EV, Petrov VV, Kulaev IS. Efflux of potassium ions from cells and spheroplasts of Saccharomyces cerevisiae yeast treated with silver and copper ions. Biochemistry (Mosc). 2008;73:1224-1227.  [PubMed]  [DOI]  [Cited in This Article: ]
218.  The International Copper Association. Effects of copper and other domestic plumbing materials on the survival of waterborne viruses. 2004; Available from: http://www.copperinfo.com.  [PubMed]  [DOI]  [Cited in This Article: ]
219.  Sagripanti JL. Metal-based formulations with high microbicidal activity. Appl Environ Microbiol. 1992;58:3157-3162.  [PubMed]  [DOI]  [Cited in This Article: ]
220.  Sagripanti JL, Routson LB, Lytle CD. Virus inactivation by copper or iron ions alone and in the presence of peroxide. Appl Environ Microbiol. 1993;59:4374-4376.  [PubMed]  [DOI]  [Cited in This Article: ]
221.  Wong K, Morgan AR, 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.  [PubMed]  [DOI]  [Cited in This Article: ]
222.  Yahaya MT, Straub TM, Yahaya MT. Inactivation of poliovirus and bacteriophage MS-2 in copper, galvanised and plastic domestic water pipes. Int Copper Res Assoc. 2001;Project 48.  [PubMed]  [DOI]  [Cited in This Article: ]
223.  Yamamoto N, Hiatt CW, Haller W. Mechanism of inactivation of bacteriophages by metals. Biochim Biophys Acta. 1964;91:257-261.  [PubMed]  [DOI]  [Cited in This Article: ]
224.  Sagripanti JL, Lightfoote MM. Cupric and ferric ions inactivate HIV. AIDS Res Hum Retroviruses. 1996;12:333-337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 37]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
225.  Jordan FT, Nassar TJ. The influence of copper on the survival of infectious bronchitis vaccine virus in water. Vet Rec. 1971;89:609-610.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 12]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
226.  Totsuka A, Otaki K. The effects of amino acids and metals on the infectivity of poliovirus ribonucleic acid. Jpn J Microbiol. 1974;18:107-112.  [PubMed]  [DOI]  [Cited in This Article: ]