Sei sulla pagina 1di 6

International Journal of Food Microbiology 113 (2007) 315 – 320

www.elsevier.com/locate/ijfoodmicro

Efflux pump activity in fluoroquinolone and tetracycline resistant


Salmonella and E. coli implicated in reduced susceptibility
to household antimicrobial cleaning agents
C.A. Thorrold a , M.E. Letsoalo b , A.G. Dusé a , E. Marais a,⁎
a
Department of Clinical Microbiology and Infectious Diseases, University of the Witwatersrand Medical School and National Health Laboratory Services,
P.O. Box 2115, Houghton 2041, Johannesburg, South Africa
b
Biostatistics and Epidemiology Unit, National Institute for Communicable Diseases, Private Bag X4, Sandringham, 2131, South Africa
Received 6 August 2005; received in revised form 12 November 2005; accepted 24 August 2006

Abstract

It has been shown that the inappropriate use of antimicrobial household agents selects for organisms with resistance mechanisms (e.g. efflux
pumps), which could lead to the development of antibiotic resistance. The reverse hypothesis, that antibiotic-resistant organisms become tolerant
to other antibacterial agents (e.g. disinfectants) due to the action of efflux pumps, has however not been extensively examined. The objective of
this study was to establish whether there is a link between antibiotic resistance in potential gastrointestinal pathogens and reduced sensitivity of
these organisms to commonly used household antimicrobial agents. In this study, tetracycline and ofloxacin sensitive and resistant Escherichia
coli (9 strains) and Salmonella spp. (8 strains) were isolated from poultry and clinical samples. In order to assess whether these bacteria had active
efflux pumps, ethidium bromide accumulation assays were performed. Extrusion of the active components of three commercial household agents
(triclosan, sodium salicylate, and ortho-phenylphenol) by efflux pumps was tested using spectrophotometric accumulation assays. In order to
simulate the kitchen environment, in-use disinfectant testing using the commercial household agents was performed to determine changes in their
efficacy due to antibiotic resistance. Active efflux pump activity and extrusion of all three active ingredients was observed only in the antibiotic
resistant organisms. The antibiotic sensitive bacteria were also more susceptible than the resistant isolates to the household antimicrobial agents at
concentrations below that recommended by the manufacturer. These resistant bacteria could potentially be selected for and result in hard to treat
infections.
© 2006 Elsevier B.V. All rights reserved.

Keywords: Household antimicrobials; Eff lux pump; Food pathogens

1. Introduction already been described in a number of clinically important bac-


teria, including Salmonella typhimurium (Webber and Piddock,
The occurrence of bacteria resistant to multiple antibiotics 2003), and Escherichia coli (Poole, 2000). Efflux pumps can
presents a serious problem in the treatment of bacterial infec- transport a variety of structurally dissimilar compounds or may
tions (Bolhuis et al., 1995). One of the underlying mechanisms be specific for a single substrate (Poole, 2000; Webber and
for this multidrug resistance phenotype is the expression and Piddock, 2003).
increased activity of efflux pumps, which facilitate the transfer There is little doubt that the inappropriate clinical application
of toxic substances from within cells to the external environ- of antimicrobial agents has led to the increased rates of multi-
ment (Bohn and Bouloc, 1998; Webber and Piddock, 2003). drug resistant bacteria (Cruchaga et al., 2001), and the use of
Efflux systems that contribute to antibiotic resistance have antibiotics in the veterinary and farming arena has recently
come under scrutiny (Singer et al., 2003). The introduction of
⁎ Corresponding author. Tel.: +27 11 489 8587; fax: +27 11 489 8530. antibiotics in livestock rearing has resulted in a variety of
E-mail address: maraise@pathology.wits.ac.za (E. Marais). developments in animal husbandry practices (Manie et al.,
0168-1605/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.ijfoodmicro.2006.08.008
316 C.A. Thorrold et al. / International Journal of Food Microbiology 113 (2007) 315–320

1998; Butaye et al., 2003). These include the breeding of ani- Table 2
mals in batteries/pens that are filled to capacity at optimal Zone diameters (including disk) of the eight Salmonella strains and nine E. coli
strains at time of study and after serial passage (brackets)
temperature and low light intensity to enhance growth rates and
increase body mass (Manie et al., 1998). Many antibiotics are Bacterial Zone diameter (mm) Result (R /S)
species
administered therapeutically for disease prevention, but sub- Tetracycline Of loxacin Tetracycline Of loxacin
therapeutic doses are also administered to improve feed effi- (30 μg) (5 μg) (30 μg) (5 μg)
ciency and accelerate weight gain in poultry, swine, sheep and Salmonella # 1 19 (20) 29 (29) S (S) S (S)
cattle (Manie et al., 1998; Chopra and Roberts, 2001; Geornaras Salmonella # 2 20 (19) 28 (32) S (S) S (S)
Salmonella # 3 21 (20) 28 (26) S (S) S (S)
et al., 2001; White et al., 2001; McEwen and Fedorka-Cray,
Salmonella # 4 21 (19) 30 (31) S (S) S (S)
2002; Butaye et al., 2003). Salmonella # 5 6a 30 a R S
Although not everyone is in agreement as to whether sub- Salmonella # 6 8 (9) 29 (33) R (R) S (S)
therapeutic doses lead to the development of antibiotic resis- Salmonella # 7 13 (14) 39 (36) R (R) S (S)
tance in bacteria, there is extensive evidence that they result in Salmonella # 8 8 (9) 46 (37) R (R) S (S)
E. coli # 1 28 (26) 36 (37) S (S) S (S)
significant selection pressures on animal pathogens and com-
E. coli # 2 26 (25) 33 (32) S (S) S (S)
mensals (Quednau et al., 1998; McEwen and Fedorka-Cray, E. coli # 3 29 a 42 a S S
2002). The antibiotic resistance that has arisen in zoonotic E. coli # 4 28 (27) 38 (37) S (S) S (S)
organisms such as Salmonella, Listeria and E. coli poses a E. coli # 5 6a 11 a R R
threat to both animal and human health (Chopra and Roberts, E. coli # 6 6 (7) 6 (7) R (R) R (R)
E. coli # 7 6 (6) 5 (8) R (R) R (R)
2001). Since potential pathogens can be transferred from
E. coli # 8 6 (6) 8 (9) R (R) R (R)
animals to humans via the food chain (Parsonnet and Kass, E. coli # 9 14 (18) 6 (6) R (R) R (R)
1987), they could result in diseases that are difficult to treat or
A zone size of ≥14 and ≥12 is considered as resistant for tetracycline and
transfer their resistance genes to other bacteria (Chopra and of loxacin, respectively (NCCLS, 2001).
Roberts, 2001). a
Serial passage not performed as strains no longer viable.
Several brands of household disinfectant products are mar-
keted as being superior to others due to the addition of one or
more antibacterial compounds, since it is claimed that these shown in Table 2. Ethics clearance was obtained from the
products kill or inactivate bacteria more efficiently than the University of the Witwatersrand for the use of patient isolates.
regular items (Kusumaningrum et al., 2002). Researchers have
recently started to express concern regarding the use of anti- 2.2. Media
microbial chemicals in the domestic setting due to possible
selection of resistant organisms (McDonnell and Russell, All media components were purchased from Diagnostic
1999). Media Products (Rietfontein, Johannesburg, South Africa), un-
This study was undertaken to establish whether there is a less otherwise indicated, and all chemicals used were of
relationship between antibiotic resistant organisms and reduced analytical grade.
susceptibility to the antimicrobial agents found in commonly
used household disinfectant products, and whether this could be 2.3. Isolation of bacteria from poultry and clinical samples
due to the action of efflux pumps.
Twenty five grams of the poultry sample was processed by
2. Materials and methods stomaching in 225 ml sterile buffered peptone water. E. coli and
salmonellae were isolated using standard microbiological cul-
2.1. Bacterial strains ture techniques. In order to isolate tetracycline or ofloxacin
resistant bacteria, 160 μg of tetracycline (Oxoid, Basingstoke,
E. coli and Salmonella species were isolated from fresh and England), or 80 μg of ofloxacin (Oxoid) was added to 10 ml of
frozen chicken products, clinical samples, and from the the enrichment broths. Salmonella isolates were typed using
Onderstepoort Veterinary Institute, University of Pretoria, API 10S (bioMerieux, Durham, USA), and 3M Petrifilms for
South Africa. The source of the bacterial isolates is shown in coliforms and E. coli (3M Microbiology Products, USA) were
Table 1 and the antibiotics that each strain was resistant to are used to confirm the presence of E. coli. Standard microbiolog-
ical procedures were used to isolate E. coli and Salmonella
from the clinical samples and veterinary samples.
Table 1
The source of gastrointestinal pathogens examined in this study 2.4. Identification and susceptibility testing
Source Bacterial isolates Number
Poultry samples Salmonella #s 1–6 and E. coli #s 1–5 11 The tetracycline and ofloxacin resistance profiles of the
Humans with infectious Salmonella #s 7 and 8 and E. coli #s 8 4 bacteria were determined using disk diffusion susceptibility
intestinal disease and 9 testing on Mueller–Hinton agar plates with reference to the
Onderstepoort Veterinary E. coli #s 6 and 7 2 National Committee for Clinical Laboratory Standards
Institute
(NCCLS, 2001) criteria (disk content of each antibiotic shown
C.A. Thorrold et al. / International Journal of Food Microbiology 113 (2007) 315–320 317

in Table 2). Only isolates with high level resistance were ana- independent occasions for each isolate and the results plotted as
lyzed further. Susceptibility testing was repeated on all isolates graphs of time vs. absorbance.
after serial passage (Suller and Russell, 2000) to confirm long
term retention of antibiotic resistance. Macro-restriction analysis 2.7. In-use disinfection procedure
was performed as described (Botteldoorn et al., 2004; Vali et al.,
2004) to ensure that the bacteria used in this study were not Bacteria were grown in 10 ml BHI overnight and added to
clonally related. 50 ml of a 10% sterilized milk solution (Oxoid). One perforated
100% viscose fibre dishcloth (purchased from a supermarket and
2.5. Ethidium bromide accumulation assays autoclaved prior to use) was added to each isolate and squeezed
with gloved hands to distribute the soiling suspension throughout
Ethidium bromide accumulation was assayed as previously the cloth. The cloth was transferred to a sterile plastic bag and left
described (Baranova and Neyfakh, 1997) with slight modifica- to incubate at room temperature for 2 h.
tions. Briefly, bacteria were grown in brain heart infusion (BHI) The dishcloths were aseptically divided into 4 equal portions of
broth, pelleted by centrifugation, and resuspended to an optical 25 cm2. One portion was transferred to quarter-strength Ringer
density of 0.2 at A600. Ethidium bromide was added to the cell solution for determination of the bacterial count without treat-
suspension at a final concentration of 2 μg/ml. Sample fluo- ment. The remaining cloth portions were immersed in the house-
rescence was used as a measure of the amount of ethidium hold disinfectant products (containing sodium salicylate, triclosan,
bromide incorporated by the cells and was recorded using a or 2-phenylphenol) at the recommended concentration, as well as
standard visible range spectrofluorimeter at excitation and emis- 25% and 50% of this concentration. After 2 min, the cloth portions
sion wavelengths of 530 and 600 nm, respectively. Five minutes were rinsed in running tap water for 30 s and a cloth portion from
after ethidium bromide addition, carbonyl cyanide m-chloro- each disinfectant treatment was transferred to flasks containing
phenylhydrazone (CCCP) (Sigma-Aldrich, MO, USA), an ef- 10 ml of quarter-strength Ringer solution for determination of total
flux pump inhibitor, was added to the cell suspension at a final bacterial counts. Serial dilutions of the rinse fluid were plated onto
concentration of 5 μg/ml. The natural fluorescence of the bac- nutrient agar plates. The colony forming units (cfus) were deter-
terial cells was subtracted by adjusting the zero point prior to mined and expressed as % reduction in bacterial numbers relative
ethidium bromide addition. to the cfus from the untreated aliquot. This procedure was repeated
in duplicate on two independent occasions for each isolate.
2.6. Measurement of chemical accumulation
2.8. Statistical analysis
Chemical accumulation of the antimicrobial components in
several household detergents was assayed by the method of Multilevel linear models, also known as Hierarchical linear
Chapman and Georgopapadakou (1988) with slight modifica- models (HLM) were employed with strains at level-1 or lower
tions. Bacteria were grown overnight at 37 °C in Luria Bertani level and groups (Salmonella and E. coli) at level-2 or higher
broth, harvested by centrifugation, washed in phosphate buf- level. STATA 8.2 (Stata Corporation, College Station, Texas,
fered saline (PBS), and adjusted to a calculated A600 of 10.0 in USA) was used in the analysis of results. Results were considered
PBS. The chemicals, sodium salicylate (Sigma Aldrich, MO, significant at the 5% level.
USA), or triclosan (kind gift from Ciba Geigy, Manchester, UK),
or ortho-phenylphenol (Sigma Aldrich), were then added to 3. Results
final concentrations of 10 mM, 4 mM and 6 mM, respectively
(the concentrations recommended by the household product 3.1. Bacterial strains
manufacturer). Samples (0.5 ml) were removed at timed inter-
vals as indicated in Fig. 3. Seven minutes after addition of the Eight Salmonella and nine E. coli strains were examined in
chemical, CCCP was added to half of the reaction mixture to a this study. Using the National Committee for Clinical Labo-
final concentration of 200 μM and the other half of the reaction ratory Standards (NCCLS, 2001) it was determined that 4 of the
mixture was used as a control (no CCCP). Each of the aliquots Salmonella species were sensitive to tetracycline (disk content
removed was immediately diluted in 1 ml of ice-cold PBS and 30 μg) and ofloxacin (disk content 5 μg) and 4 were resistant to
centrifuged at 5700 ×g for 5 min. The pellets were washed with tetracycline only. Four of the E. coli strains were sensitive and
1 ml of ice-cold PBS, resuspended in 1 ml of PBS and boiled for five were found to be resistant to both the antibiotics. To test the
5 min to lyse the cells. The samples were centrifuged at 5700 ×g stability of the antibiotic resistance mechanisms, the isolates
for 10 min and the absorbance of the supernatant was measured were serially passaged for 20 days. Three of the 17 strains were
using a 3 series UV spectrophotometer (BioMate, NY, USA) at not viable after 2 years of storage and could not be tested. The
wavelengths of 296 nm (for sodium salicylate), 382 nm (for viable isolates were resistant to the original antibiotic(s) of
Triclosan) and 251 nm (ortho-phenylphenol), which were de- selection after this period (Table 2). This suggests that the genes
termined to be suitable for the measurement of each aromatic coding for resistance are chromosomal, but this was not inves-
compound. The absorbance readings are indicative of the tigated further. Dissimilar banding patterns were observed in the
amount of agent internalized by the bacterial cells and therefore macro-restriction analyses of all the bacteria examined, indicat-
of accumulation. This procedure was done in duplicate on two ing that they were different strains.
318 C.A. Thorrold et al. / International Journal of Food Microbiology 113 (2007) 315–320

with a significant difference in the absorbance reading in the


presence and absence of CCCP (z = 10.76; p = 0.000), as a result
of a greater amount of cell-associated active component
(Fig. 2B). Similar results were observed with all three active
ingredients tested, demonstrating that the resistant strains had
efflux pumps that were extruding the active ingredients found in
commonly used household disinfectant products, and that these
efflux pumps were energy dependent.

3.4. In-use disinfection procedure

An in-use disinfectant test was used to assess the effect of


various concentrations of three commonly used household
Fig. 1. Representative figures of accumulation of ethidium bromide by (A) disinfectants, containing sodium salicylate (10 mM), triclosan
antibiotic sensitive E. coli # 2 and (B) resistant E. coli # 5. On addition of (4 mM), or 2-phenylphenol (6 mM), on the antibiotic sensitive
ethidium bromide, f luorescence intensity ( y-axis) increased, and then leveled and resistant bacteria.
off. (A) Addition of the eff lux pump inhibitor, CCCP, had no effect on After the disinfection procedure with all three household
f luorescence intensity indicating that no eff lux pumps were inhibited.(B)
products, there was a dramatic reduction in the number of
Addition of the eff lux pump inhibitor, CCCP, resulted in a sharp increase in
f luorescence indicating inactivation of eff lux pumps (n = 1). bacteria remaining on the cloths; as can be seen in Fig. 3A and B.
The greater the concentration of disinfectant used, the greater the
reduction in bacterial numbers; as seen in Fig. 3A and B.
3.2. Ethidium bromide accumulation With the antibiotic sensitive bacteria, the difference in the
number of surviving bacteria at different disinfectant concen-
Ethidium bromide accumulation was assayed to determine trations (Fig. 3A) was marginal. However, with the resistant
whether the bacteria had active efflux pumps. As shown in bacteria, there was a noticeable difference in the number of
Fig. 1A and B, the fluorescence intensity was seen to increase remaining bacteria at the different concentrations (Fig. 3B),
substantially on addition of ethidium bromide for each strain
examined. This was indicative of the cells taking up the dye, as
binding of ethidium bromide to intracellular components results
in increased fluorescence. By 5 min after the addition of ethid-
ium bromide a plateau in the fluorescence values was reached.
On addition of CCCP, the fluorescence intensity did not change
in the sensitive bacterial strains (Fig. 1A). These results indicate
that efflux pumps were not extruding the dye, since their in-
activation had no effect on the equilibrium. However, in each of
the resistant bacterial strains examined, addition of CCCP resulted
in a rapid and dramatic increase in fluorescence (Fig. 1B). This
was due to the inhibition of energy dependent efflux pumps,
resulting in a greater net accumulation of ethidium bromide.

3.3. Measurement of chemical accumulation

Cellular accumulation of the chemicals was measured in


order to establish whether efflux pumps could extrude the active
ingredients of three household disinfectants. In all the tetra-
cycline and ofloxacin sensitive bacterial strains, there was an
increase in the absorbance readings in both the presence and
absence of CCCP (Fig. 2A), with no significant difference be-
tween the presence and absence of CCCP (z = − 0.88; p = 0.381).
These results demonstrate the absence of efflux of the active
ingredients from the sensitive bacterial cells, and that the chemi-
cals were accumulating in the bacterial cells. This was observed
for all three of the active ingredients tested (4 mM triclosan,
10 mM sodium salicylate, and 6 mM ortho-phenylphenol). Fig. 2. Graph of accumulation of triclosan by (A) sensitive Salmonella # 2 and
(B) resistant Salmonella # 8. Absorbance on the y-axis is a measure of the levels
In the antibiotic resistant strains, the absorbance values in the of accumulated triclosan in the bacterial cells. Each data point represents the
absence of CCCP increased gradually with time. However, upon mean ± standard error of the mean of 2 independent experiments. Consistently
addition of CCCP, the absorbance values increased considerably, similar results were seen for the other strains and chemicals (n = 2).
C.A. Thorrold et al. / International Journal of Food Microbiology 113 (2007) 315–320 319

biotic resistant Salmonella and E. coli isolates, indicating that


the cells were unable to extrude the agents upon inhibition of the
pumps. The active components (sodium salicylate, triclosan and
ortho-phenylphenol) all contain aromatic groups, but are other-
wise not structurally related. The ability of the antibiotic resistant
strains to extrude all three chemicals indicates a promiscuity of
function of a single pump, or the activity of multiple pumps.
Cleaning cloths and sponges play a very important role in
domestic hygiene as they aid in the detachment of microbes and
soil present on the surface (Baranova and Neyfakh, 1997).
However, in many instances they remain moist for long periods
of time, contain residual amount of soil, and are not changed
frequently, providing ideal conditions for the survival and/or
growth of microorganisms (Bohn and Bouloc, 1998). Studies
have demonstrated that if cloths are not properly disinfected,
they can transfer bacteria to clean surfaces in the kitchen or
hands, resulting in disease (Bolhuis et al., 1995). Several brands
of household cleaning products are available that contain anti-
microbial agents, and these are marketed as being more effec-
tive at killing or inactivating bacteria than regular products. An
in-use disinfection procedure was used to test the efficacy of
several commonly used products containing salicylate, triclosan
and ortho-phenylphenol by simulating cleaning. The products
Fig. 3. Representative bar chart graph showing the effect of different
were used at the concentrations recommended by the manufac-
concentrations of household disinfectant product (containing ortho-phenylphe-
nol) on (A) sensitive E. coli # 2 and (B) resistant E. coli # 7 in artificially turer, as well as at half and quarter strengths. The rationale for
contaminated dishcloths. Each error bar represents the standard error of the using the items at concentrations other than that recommended,
mean of 2 independent experiments (n = 2). is that people may fail to adhere to the instructions and dilute the
agents to extend their use. This is particularly the case when a
with a significant difference observed between the sensitive 100% (neat) application of a product is required for anti-
and resistant bacteria at the 25% concentration (z = 30.20; germicidal action, but it is most practically used with water for
p = 0.000). cleaning purposes. As expected, the higher the concentration of
detergent product used, the greater the reduction in bacterial
4. Discussion numbers amongst both antibiotic sensitive and resistant bac-
teria. However, as the product was diluted to 25%, more anti-
The zoonotic organisms, Salmonella and E. coli, are known biotic resistant bacteria survived than sensitive isolates. This
to be associated with human gastrointestinal disease following was seen with all three products tested and for both organisms.
consumption of contaminated food. Similar to what has been The ability of tetracycline and ofloxacin resistant Salmonella
previously described, Salmonella and E. coli species resistant and E. coli to stably extrude the active ingredients of three
to tetracycline and/or ofloxacin could be isolated from poultry antimicrobial detergents used, suggests a mechanism of how
samples (Quednau et al., 1998). These organisms may find their reduced susceptibility could occur. Efflux is thus postulated to
way into the home and be ingested if proper kitchen hygiene is be involved in the reduction in efficacy of the ‘anti-germ’
not practiced. household products observed at higher dilutions. While the use
Tetracycline and ofloxacin were chosen as indicator anti- of antimicrobial household detergents may assist in reducing
biotics due to their well documented pathways of resistance due bacterial load in the kitchen and elsewhere, incorrect usage
to efflux pump activity (Heisig, 1996), which is the mechanism could result in selection of bacteria with reduced susceptibility
most likely to be involved in cross resistance. The finding that to both antibiotics and antimicrobials. This could occur where
active efflux of ethidium bromide was solely associated with there is a sufficient concentration of the antimicrobial agent to
antibiotic resistant organisms suggests that efflux mechanisms eliminate sensitive bacteria, but insufficient to reduce the num-
could be responsible for the antibiotic resistance, but this was bers of the less susceptible, antibiotic resistant strains. This is
not further investigated and other systems may also be involved. clearly undesirable in a setting such as that of South Africa,
Time-course sodium salicylate, triclosan and ortho-phenyl- where many individuals are immuno-compromised due to HIV/
phenol accumulation experiments were performed, to determine AIDS and are receiving home-based medical care.
whether active efflux of these chemicals could be observed. The
use of the pump inhibitor CCCP allowed us to block energy Acknowledgement
dependent efflux pumps, which resulted in an increase in the
cytoplasmic levels of the chemicals. A significant increase in We thank Dr Jackie Picard, Onderstepoort Veterinary Insti-
the amount of cell-associated components was seen with anti- tute, University of Pretoria, South Africa for E. coli isolates.
320 C.A. Thorrold et al. / International Journal of Food Microbiology 113 (2007) 315–320

References Kusumaningrum, H.D., van Putten, M.M., Rombouts, F.M., Beumer, R.R.,
2002. Effects of antibacterial dishwashing liquid on foodborne pathogens
Baranova, N.N., Neyfakh, A.A., 1997. Apparent involvement of a multidrug and competitive microorganisms in kitchen sponges. Journal of Food
transporter in the f luoroquinolone resistance of Streptococcus pneumoniae. Protection 65, 61–65.
Manie, T., Khan, S., Brozel, V.S., Veith, W.J., Gouws, P.A., 1998. Antimicrobial
Antimicrobial Agents and Chemotherapy 41, 1396–1398.
Bohn, C., Bouloc, P., 1998. The Escherichia coli cmlA gene encodes the resistance of bacteria isolated from slaughtered and retail chickens in South
multidrug eff lux pump Cmr/MdfA and is responsible for isopropyl-beta-D- Africa. Letters in Applied Microbiology 26, 253–258.
McDonnell, G., Russell, A.D., 1999. Antiseptics and disinfectants: activity,
thiogalactopyranoside exclusion and spectinomycin sensitivity. Journal of
Bacteriology 180, 6072–6075. action, and resistance. Clinical Microbiology Reviews 12, 147–179.
Bolhuis, H., Poelarends, G., van Veen, H.W., Poolman, B., Driessen, A.J., McEwen, S.A., Fedorka-Cray, P.J., 2002. Antimicrobial use and resistance in
Konings, W.N., 1995. The lactococcal lmrP gene encodes a proton motive animals. Clinical Infectious Diseases 34 (Suppl 3), S93–S106.
National Committee for Clinical Laboratory Standards, 2001. Performance
force-dependent drug transporter. Journal of Biological Chemistry 270,
26092–26098. standards for antimicrobial susceptibility testing, 11th ed. pp. 42–43.
Botteldoor n, N., Herman, L., Rijpens, N., Heyndrickx, M., 2004. Phenotypic Parsonnet, K.C., Kass, E.H., 1987. Does prolonged exposure to antibiotic-
and molecular typing of Salmonella strains reveals contamination sources in resistant bacteria increase the rate of antibiotic-resistant infection?
Antimicrobial Agents and Chemotherapy 31, 911–914.
two commercial pig slaughterhouses. Applied and Environmental Microbi-
ology 70, 5305–5314. Poole, K., 2000. Eff lux-mediated resistance to f luoroquinolones in gram-
Butaye, P., Devriese, L.A., Haesebrouck, F., 2003. Antimicrobial growth negative bacteria. Antimicrobial Agents and Chemotherapy 44, 2233–2241.
Quednau, M., Ahrne, S., Petersson, A.C., Molin, G., 1998. Antibiotic-resistant
promoters used in animal feed: effects of less well known antibiotics on
gram-positive bacteria. Clinical Microbiology Reviews 16, 175–188. strains of Enterococcus isolated from Swedish and Danish retailed chicken
Chapman, J.S., Georgopapadakou, N.H., 1988. Routes of quinolone permeation and pork. Journal of Applied Microbiology 84, 1163–1170.
in Escherichia coli. Antimicrobial Agents and Chemotherapy 32, 438–442. Singer, R.S., Finch, R., Wegener, H.C., Bywater, R., Walters, J., Lipsitch, M.,
2003. Antibiotic resistance—the interplay between antibiotic use in animals
Chopra, I., Roberts, M., 2001. Tetracycline antibiotics: mode of action,
applications, molecular biology, and epidemiology of bacterial resistance. and human beings. Lancet. Infectious Diseases 3, 47–51.
Microbiology and Molecular Biology Reviews 65, 232–260. Suller, M.T.E., Russell, A.D., 2000. Triclosan and antibiotic resistance in
Cruchaga, S., Echeita, A., Aladuena, A., Garcia-Pena, J., Frias, N., Usera, M.A., Staphylococcus aureus. Journal of Antimicrobial Chemotherapy 46, 11–18.
Vali, L., Wisely, K.A., Pearce, M.C., Turner, E.J., Knight, H.I., Smith, A.W.,
2001. Antimicrobial resistance in salmonellae from humans, food and
animals in Spain in 1998. Journal of Antimicrobial Chemotherapy 47, Amyes, S.G., 2004. High-level genotypic variation and antibiotic sensitivity
315–321. among Escherichia coli O157 strains isolated from two Scottish beef cattle
Geornaras, I., Hastings, J.W., von Holy, A., 2001. Genotypic analysis of farms. Applied and Environmental Microbiology 70, 5947–5954.
Webber, M.A., Piddock, L.J., 2003. The importance of eff lux pumps in bacterial
Escherichia coli strains from poultry carcasses and their susceptibilities to
antimicrobial agents. Applied and Environmental Microbiology 67, antibiotic resistance. Journal of Antimicrobial Chemotherapy 51, 9–11.
1940–1944. White, D.G., Zhao, S., Sudler, R., Ayers, S., Friedman, S., Chen, S., McDermott,
P.F., McDermott, S., Wagner, D.D., Meng, J., 2001. The isolation of
Heisig, P., 1996. Genetic evidence for a role of parC mutations in development
of high-level f luoroquinolone resistance in Escherichia coli. Antimicrobial antibiotic-resistant salmonella from retail ground meats. New England
Agents and Chemotherapy 40, 879–885. Journal of Medicine 345, 1147–1154.

Potrebbero piacerti anche