首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 859 毫秒
1.
Steady-state pharmacokinetics of ertapenem were compared in patients after 1-g intravenous and subcutaneous (s.c.) infusions. Bioavailability was 99% ± 18% after s.c. administration, but peaks were reduced by about (43 ± 29 versus 115 ± 28 μg/ml) and times to peak were delayed. Simulations based on unbound concentrations show that time over the MIC should always be longer than 30% to 40% of the dosing interval, suggesting that s.c. infusion could be an alternative in patients with reduced vascular access.Ertapenem is a recent long-acting, parenteral carbapenem antibiotic mainly indicated in the treatment of community-acquired infections or hospital-acquired infections without suspicion of Pseudomonas or Acinetobacter (5), as an alternative to penicillin-β-lactamase inhibitor combination (10, 19, 23). The pharmacokinetics of ertapenem has been extensively described (2, 3, 6-8, 15, 16, 18, 21). Ertapenem may be administered intravenously (i.v.) or intramuscularly (i.m.) for several days (13), but for many hospitalized patients the i.m. route might be contraindicated due to anticoagulant therapy. Subcutaneous (s.c.) administration is daily safely used with drugs and fluids mostly for dehydrated elderly patients or patients in palliative care when oral or i.v. administration is impossible (20) and could then appear as an interesting alternative. Advantages for the s.c. route over the i.v. route include a similar number of or even fewer complications, cost savings, greater patient comfort, and less nursing time to start and maintain the infusion (1, 22). The aim of this study was to compare the pharmacokinetics of ertapenem at steady state following 30-min i.v. and s.c. infusions in order to determine if s.c. administration of ertapenem, which is not yet approved, could be a viable alternative for i.v. infusion in patients with limited vascular sites.The study was conducted at the University Hospital of Poitiers (France) after its approval by the local ethics committee (Region Poitou-Charentes CCPPRB, protocol no. 05.12.26). Written informed consent was obtained from each subject or their closest relative if the patient was unconscious. The study enrolled 6 adult male patients suspected of having an infection due to ertapenem-susceptible bacteria (Table (Table1).1). Ertapenem was the only antibiotic used, sedation was obtained with propofol and sufentanil, and no other drugs that could have been suspected of interacting with ertapenem pharmacokinetics such as vasopressors or midazolam were used. At study enrollment, patients were mechanically ventilated and exhibited a systemic inflammatory response syndrome. Infection sites justifying ertapenem administration were early-onset ventilator-associated pneumonia (n = 5) and surgical wound infection (patient no. 2). The microorganisms isolated at those infection sites were methicillin-susceptible Staphylococcus aureus (n = 2), Haemophilus influenzae (n = 1), Escherichia coli (n = 1), and Klebsiella pneumoniae (n = 2). Local tolerance was assessed during the 24 h following subcutaneous infusion by checking for erythema, pruritus, hematoma, or necrosis at the insertion site. Ertapenem (Invanz) was purchased from the pharmaceutical company Merck Sharp & Dohme-Chibret (Paris, France) as a dry powder and reconstituted in 50 ml normal saline just before being infused with a pump (Orchestra DPS; Fresenius Vial, Brezins, France). The administration sites were a central vein (i.v.) or the anterior side of a thigh (s.c.). Initially 1 g of ertapenem was administered i.v. over 30 min once daily, and blood samples for the i.v. pharmacokinetic study were collected between the 4th day and the 7th day. The next day, treatment was shifted to the s.c. route and a second series of blood samples was collected during the following 24 h for the s.c. pharmacokinetic study. Ertapenem administration was shifted back to the i.v. route until the end of therapy. Blood samples were drawn via an arterial catheter in heparinized tubes and immediately centrifuged for 10 min at 2,500 × g and 4°C to separate plasma, which was then transferred to storage vials and diluted (1:1) with a stabilizing solution consisting of 1:1 ethylene glycol and 2-(4-morpholino) ethylsulfonic acid at 0.1 mol/liter (pH 6.5). Plasma ultrafiltrates were obtained from plasma samples collected at times 0.5 h and 24 h postdosing by centrifugation with a Centrifree system (CF50A model; Amicon, Molsheim, France). Samples were stored at −80°C until analysis. Ertapenem concentrations were measured using a liquid chromatography method with tandem mass spectrometry detection (12). Within- and between-day variability of the method at various concentrations led to coefficients of variation no greater than 16.4% (n = 11) and accuracies ranging between 98.5% and 110.9%. A compartmental pharmacokinetic analysis for total plasma concentrations was conducted with WinNonLin version 4.0.1. (Pharsight Corporation, Mountain View, CA), using a two-open-compartment model with multiple zero order infusions after i.v. administrations followed by a one-open-compartment model with one zero order infusion after s.c. administration. Duration of infusion was set at a fixed value equal to 0.5 h after i.v. administrations, but estimated by the modeling after s.c. administration. A 1/y weight was used for all the analysis. Unbound concentrations were derived from measured total concentrations using a saturable two-class binding site model with one specific binding site and one nonspecific binding site, previously validated for ertapenem (7). The rate constants for specific and nonspecific binding sites were estimated from the 12 pairs of total and unbound concentrations measured at 0.5 and 24 h and using the mean concentration values of albumin (14.1 g/liter) and the remaining proteins (38.7 g/liter) characteristic of these patients. The same compartmental pharmacokinetic analysis as for total concentrations was conducted with unbound concentrations, and simulations were derived for each subject and each route of administration in order to estimate the percentage of dosing interval during which unbound concentrations would be higher than various breakpoint values, chosen as 1, 2, 4, and 8 mg/liter, in agreement with the Clinical Laboratory Standards Institute. Results are presented as means ± standard deviation (SD), and nonparametric Wilcoxon''s rank test was used for statistical comparisons, with P < 0.05 considered as significant.

TABLE 1.

Patient characteristics and individual and mean ± SD pharmacokinetic parameters based on total ertapenem concentrations measured after i.v. and s.c. 30-min infusions of ertapenem (1 g/24 h) to 6 adult patients
ParameteraResult for patient:
Mean ± SDb
123456
Patient characteristics
    Age (yr)67521963558156 ± 19
    Body wt (kg)726076901006677 ± 14
    Height (m)1.671.721.801.801.851.701.76 ± 0.07
    BMI (kg/m2)25.820.323.527.829.222.824.9 ± 3.0
    SAPS II on admission20182159213028 ± 14
    Albumin concn (g/liter)10.014.416.814.39.519.614.1 ± 3.6
    Creatinine concn (μmol/liter)27374366867255 ± 21
    Total concn of proteins (g/liter)48576549405853 ± 9
    Fluid balance (ml)
        i.v.+ 350+ 950+ 200+ 350+ 800+ 100+458 ± 340
        s.c.+ 500+ 700+ 200+ 150+ 650+ 150+392 ± 255
    ICU outcomeSurvivedSurvivedSurvivedSurvivedSurvivedSurvived
Pharmacokinetics
    Cmax (μg/ml)
        i.v.9412410475148142115 ± 28
        s.c.25257719288443 ± 29*
    tmax (h)
        i.v.0.50.50.50.50.50.50.5
        s.c.2.71.32.84.62.32.32.7 ± 1.1*
    t1/2 (h)
        i.v.3.42.34.73.04.95.33.9 ± 1.2
        s.c.5.16.32.86.66.45.35.4 ± 1.4*
    AUC0-24 s.c./AUC0-24 i.v.0.871.021.031.180.701.140.99 ± 0.18
    CL (liters/h) i.v.4.74.43.16.22.61.73.8 ± 1.6
    Vss (liters) i.v.19.412.615.023.115.310.916.1 ± 4.5
    fu (%)
        i.v.45.8 ± 4.443.4 ± 4.340.9 ± 3.945.8 ± 3.351.4 ± 5.546.8 ± 4.6
        s.c.43.4 ± 1.741.4 ± 1.540.4 ± 2.844.1 ± 0.948.3 ± 2.245.3 ± 2.9
Open in a separate windowaBMI, body mass index; SAPS, simplified acute physiology score; ICU, intensive care unit; Cmax, maximal concentration of ertapenem; tmax, time to obtain maximal concentration; t1/2, half-life of elimination; AUC0-24, area under the curve from 0 to 24 h; CL, clearance; Vss, volume of distribution at steady state; fu, unbound fraction of ertapenem.b*, P < 0.05.All patients completed the study without any local or systemic adverse effect attributable to ertapenem administration, and signs of infection had disappeared by the end of treatment. Ertapenem plasma concentration-time profiles were shifted to the right after s.c. infusion, with an approximately 3-fold reduction of peak concentrations (Cmax) and 5-fold increase of time to peak concentration (tmax) (Table (Table1).1). However, after 3 h postdosing on average, plasma concentrations became higher following s.c. infusion (Fig. (Fig.1),1), and AUCs were virtually identical after both routes of administration, attesting for complete bioavailability following s.c. infusion. However, because only unbound drug has the ability to distribute and to exert antimicrobial activity at the target site of infection, unbound concentrations should be considered to predict efficacy (4, 14). In this study, ertapenem protein binding demonstrated no sign of nonlinearity and was relatively limited, with unbound fractions (fu) ranging from 40.4% ± 2.8% to 51.4% ± 5.5% (Table (Table1),1), consistent with the value (fu = 54.8% ± 19.1%) recently reported by Burkhardt et al. in critical care patients (7), but much higher than the average value (16% unbound corresponding to 84% bound) currently reported in healthy volunteers (15, 20). Because ertapenem antimicrobial activity is considered to be time dependent (9), a peak reduction after s.c. administration may not have major consequences on its clinical efficacy. Instead, the dosing interval during which unbound drug concentration exceeds the MIC (t > MIC), represents the most relevant pharmacokinetics/pharmacodynamics parameter (17), and a t > MIC of 30 to 40% of the dosing interval should be effective (11). Conducted simulations suggested that for susceptible and intermediately susceptible microorganisms (MIC ≤ 4 mg/liter), t > MIC based upon unbound ertapenem concentrations should always be longer than 30% to 40% of the dosing interval, independently of the route of administration. In conclusion, this study suggests that s.c. infusion of ertapenem should be equivalent to i.v. infusions in terms of efficacy and could therefore represent an interesting alternative for patients with reduced vascular access, such as dehydrated elderly patients or patients in palliative care. However, this should be confirmed in a larger population of such patients.Open in a separate windowFIG. 1.Mean ± SD total ertapenem concentrations in plasma after multiple daily intravenous infusions (1 g over 30 min) followed by a subcutaneous infusion (1 g over 30 min) in 6 patients. Closed symbols and the solid line correspond to intravenous infusion, and open symbols and the dashed line correspond to subcutaneous infusion.  相似文献   

2.
LCB01-0371 is a new oxazolidinone with cyclic amidrazone. In vitro activity of LCB01-0371 against 624 clinical isolates was evaluated and compared with those of linezolid, vancomycin, and other antibiotics. LCB01-0371 showed good activity against Gram-positive pathogens. In vivo activity of LCB01-0371 against systemic infections in mice was also evaluated. LCB01-0371 was more active than linezolid against these systemic infections. LCB01-0371 showed bacteriostatic activity against Staphylococcus aureus.The emergence of multidrug-resistant (MDR) pathogens, such as methicillin-resistant Staphylococcus aureus (MRSA), methicillin-resistant coagulase-negative staphylococci (MRCNS), penicillin-resistant Streptococcus pneumoniae (PRSP), and vancomycin-resistant enterococci (VRE), has generated worldwide concern in the medical community (11). The requirement for effective new antimicrobial agents to treat infections caused by Gram-positive organisms is becoming urgent as resistance to existing agents arises and spreads around the world.The oxazolidinones, a totally synthetic class of novel antibiotics, have strong activity against nearly all Gram-positive organisms, including those resistant to other agents (1, 10). They inhibit protein synthesis by binding to domain V of the 23S rRNA and thereby blocking formation of the initiation complex (6). Linezolid is the first member of the oxazolidinone class approved by the FDA in the United States. The success of linezolid and the occurrence of strains resistant to linezolid in clinical isolates of Enterococcus faecium (4, 5) and S. aureus (12) have inspired further efforts toward developing new oxazolidinones with improved safety and antibacterial activity.LCB01-0371 (Fig. (Fig.1),1), a novel oxazolidinone with cyclic amidrazone, was synthesized by LegoChem BioSciences Inc. (Daejeon, Republic of Korea). In this study, in vitro activity of LCB01-0371 was compared with those of eight different antibacterial agents against 624 clinical isolates that were collected from several general hospitals in the Republic of Korea. In vivo activity of LCB01-0371 against systemic infections in mice and time-kill studies of LCB01-0371 against S. aureus giorgio (methicillin-susceptible S. aureus [MSSA]) and S. aureus p125 (MRSA) were also investigated.Open in a separate windowFIG. 1.Chemical structure of LCB01-0371.(This study was presented in part at the 49th Annual Interscience Conference on Antimicrobial Agents and Chemotherapy, San Francisco, CA, 2009 [7].)In vitro MICs were determined by the 2-fold agar dilution method as described by the Clinical and Laboratory Standards Institute (CLSI) (3). Mueller-Hinton agar (MHA) medium was used for testing aerobic and facultative organisms. Streptococcus pneumoniae, Streptococcus pyogenes, and Moraxella catarrhalis were grown on Mueller-Hinton agar supplemented with 5% defibrinated sheep blood (Hanil Komed Ltd., Sungnam City, Republic of Korea). Mueller-Hinton agar supplemented with 3% Fildes enrichment (Oxoid Ltd., Basingstoke, Hampshire, England) was used for Haemophilus influenzae. Bacteria (104 to 105 CFU) were spotted onto plates containing the appropriate concentration of drug. Plates were incubated at 35°C for 18 h and examined for growth. The MIC was considered to be the lowest concentration that completely inhibited growth on agar plates, disregarding a single colony or a faint haze caused by the inoculum.Time-kill studies were performed by the M26-A method of the NCCLS (8). Test organisms incubated on tryptic soy agar (TSA) for 18 h at 37°C were diluted with fresh Mueller-Hinton broth to ∼105 CFU/ml, and the diluted cultures were preincubated for 2 h. Each drug was added to the cultures at concentrations of 0.25×, 0.5×, 1×, 2×, 4×, and 8× MIC. Aliquots (0.1 ml) of the cultures were removed at 0, 2, 4, 6, and 24 h of incubation, and serial 10-fold dilutions were prepared in saline as needed. Drug carryover effects were reduced by 100-fold dilution of the sample with agar. The numbers of viable cells on drug-free MHA plates after 24 h of incubation were determined. The compound was considered bactericidal at the concentration that reduced the original inoculum by 3 log10 CFU/ml (99.9%) at each of the time periods or considered bacteriostatic if the inoculum was reduced by ∼0 to 3 log10 CFU/ml.In vivo activity of LCB01-0371 against systemic infections caused by S. aureus giorgio (MSSA), S. aureus p125 (MRSA), Enterococcus faecalis u810, S. pneumoniae ATCC 6305, and Haemophilus influenzae hd2 in mice was determined. Four-week-old male ICR mice weighing 18 to 22 g (Daehan Bio Link Co., Ltd., Eum-sung Gun, Republic of Korea) were used for the systemic infection model. They were maintained in animal rooms kept at 23 ± 2°C with 55% ± 20% relative humidity. Test organisms for infection were cultured in Mueller-Hinton agar medium (Difco) at 37°C for 18 h. For S. pneumoniae, Muller-Hinton agar medium was supplemented with 5% defibrinated sheep blood. For use as inocula, bacterial strains were suspended in 0.9% saline solution containing 5% gastric mucin (Sigma), except for S. pneumoniae, which was suspended in 0.9% saline solution. Mice were used in groups of six for each dose and were challenged intraperitoneally with a single 0.5-ml portion of the bacterial suspension, corresponding to an inoculum range of 10 to 100 times the minimal lethal dose of bacteria. Four dose levels were used for each antibiotic, depending on the in vitro antimicrobial activity of the compound. Antibiotics at various dose regimens were administered orally twice, at 1 and 4 h postinfection. Mortality was recorded for 7 days, and the median effective dose needed to protect 50% of the mice (ED50) was calculated by the Probit method (2). The challenge inoculum was sufficient to kill 100% of the untreated control mice, which died within 48 h postinfection.All animal experiments were approved by the Ethics Review Committee of Handong Global University, Republic of Korea.The comparative in vitro antibacterial activities of LCB01-0371 are shown in Table Table1.1. The MIC90 of LCB01-0371 for MSSA and MRSA was 2 μg/ml. LCB01-0371 was as active as linezolid. Against methicillin-susceptible coagulase-negative staphylococci (MSCNS) (MIC90, 0.5 μg/ml) and MRCNS (MIC90, 0.5 μg/ml), LCB01-0371 was at least 2-fold more active than linezolid. LCB01-0371 was equally active irrespective of whether the strains were methicillin susceptible or resistant. Against S. pneumoniae (MIC90, 1 μg/ml) and S. pyogenes (MIC90, 2 μg/ml), LCB01-0371 showed antibacterial activity comparable to that of linezolid. LCB01-0371 was as active as linezolid against E. faecalis (MIC90, 2 μg/ml) and E. faecium (MIC90, 2 μg/ml). Against VRE (MIC90, 1 μg/ml), LCB01-0371 was 2-fold more active than linezolid. LCB01-0371 showed weak activity against the fastidious Gram-negative aerobes H. influenzae and M. catarrhalis. Against H. influenzae, LCB01-0371 yielded a MIC90 of 16 μg/ml, while slightly better activity against M. catarrhalis (MIC90, 8 μg/ml) was seen. The MIC90 of LCB01-0371 against H. influenzae was 2-fold lower than that of linezolid, but the MICs of LCB01-0371 against H. influenzae and M. catarrhalis were too high for clinical efficacy.

TABLE 1.

In vitro antibacterial activities of LCB01-0371 against clinical isolates
Microorganism (no. of strains) and compoundMIC (μg/ml)
Range50%90%
MSSA (69)
    LCB01-0371∼0.5-212
    Linezolid∼2-422
    Oxacillin∼0.06-10.250.5
    Erythromycin∼0.125->640.25>64
    Ciprofloxacin∼0.06->640.250.5
    Moxifloxacin∼0.015-640.060.125
    Gemifloxacin∼0.008-640.0150.06
    Vancomycin∼0.25-211
    Quinupristin-dalfopristin∼0.125-0.50.250.5
∼0.125->64>64>64
MRSA (202)
    LCB01-0371∼0.5-412
    Linezolid∼2-222
    Oxacillin∼2->64>64>64
    Erythromycin∼0.25->64>64>64
    Ciprofloxacin∼0.125->6432>64
    Moxifloxacin∼0.03->64464
    Gemifloxacin∼0.008->64264
    Vancomycin∼0.5-412
    Quinupristin-dalfopristin∼0.125-10.51
∼0.125->64>64>64
MSCNS (20)
    LCB01-0371∼0.5-10.50.5
    Linezolid∼1-212
    Oxacillin∼0.03-10.1251
    Erythromycin∼0.06->640.25>64
    Ciprofloxacin∼0.06-80.1258
    Moxifloxacin∼0.03-40.1254
    Gemifloxacin∼0.008-0.50.0150.5
    Vancomycin∼1-424
    Quinupristin-dalfopristin∼0.125-10.251
∼>64->64>64>64
MRCNS (33)
    LCB01-0371∼0.5-10.50.5
    Linezolid∼1-211
    Oxacillin∼2->64>64>64
    Erythromycin∼0.06->64>64>64
    Ciprofloxacin∼0.06-64832
    Moxifloxacin∼0.06-1628
    Gemifloxacin∼0.008-80.51
    Vancomycin∼1-422
    Quinupristin-dalfopristin∼0.125-80.251
S. pneumoniae (97)
    LCB01-0371∼0.125-20.51
    Linezolid∼0.5-111
    Oxacillin∼0.008->64816
    Erythromycin∼0.008->6464>64
    Ciprofloxacin∼0.5-3212
    Moxifloxacin∼0.06-40.250.5
    Gemifloxacin∼0.008-0.250.030.06
    Vancomycin∼0.25-10.51
    Quinupristin-dalfopristin∼0.5-412
S. pyogenes (46)
    LCB01-0371∼0.5-212
    Linezolid∼1-222
    Oxacillin∼0.03-160.58
    Erythromycin∼0.008-80.062
    Ciprofloxacin∼0.5-212
    Moxifloxacin∼0.125-0.250.1250.25
    Gemifloxacin∼0.015-0.1250.030.06
    Vancomycin∼0.5-111
    Quinupristin-dalfopristin∼0.25-412
E. faecalis (109)
    LCB01-0371∼1-222
    Linezolid∼1-222
    Oxacillin∼8->6416>64
    Erythromycin∼0.125->64>64>64
    Ciprofloxacin∼0.06->64264
    Moxifloxacin∼0.06-64132
    Gemifloxacin∼0.008-160.1254
    Vancomycin∼0.5-6424
    Quinupristin-dalfopristin∼0.25-16416
E. faecium (29)
    LCB01-0371∼1-222
    Linezolid∼1-222
    Oxacillin∼16->64>64>64
    Erythromycin∼0.125->64>64>64
    Ciprofloxacin∼1-64464
    Moxifloxacin∼0.25->64432
    Gemifloxacin∼0.03-64216
    Vancomycin∼0.5-812
    Quinupristin-dalfopristin∼0.25-320.54
VRE (16)
    LCB01-0371∼1-111
    Linezolid∼2-222
    Oxacillin∼32->64>64>64
    Erythromycin∼>64->64>64>64
    Ciprofloxacin∼0.5-444
    Moxifloxacin∼0.25-424
    Gemifloxacin∼0.015-20.52
    Vancomycin∼>64->64>64>64
    Quinupristin-dalfopristin∼0.5-222
M. catarrhalis (20)
    LCB01-0371∼2-848
    Linezolid∼4-888
    Oxacillin∼0.25-32816
    Ciprofloxacin∼<0.008-0.060.030.06
    Moxifloxacin∼0.015-0.060.060.06
    Gemifloxacin∼<0.008-0.03<0.0080.015
    Vancomycin∼64->6464>64
    Quinupristin-dalfopristin∼0.5-211
H. influenzae (13)
    LCB01-0371∼2-16816
    Linezolid∼8-321632
    Oxacillin∼>32->32>32>32
    Erythromycin∼0.5-828
    Ciprofloxacin∼<0.008-<0.008<0.008<0.008
    Moxifloxacin∼0.008-0.0150.0080.008
    Gemifloxacin∼<0.008-<0.008<0.008<0.008
    Vancomycin∼>64->64>64>64
    Quinupristin-dalfopristin∼2-848
Open in a separate windowThe time-kill analyses of LCB01-0371 against S. aureus giorgio (MSSA) and S. aureus p125 (MRSA) are presented in Fig. Fig.2.2. LCB01-0371 and linezolid showed similar patterns of the time-kill effect irrespective of whether the strain was methicillin susceptible or resistant. LCB01-0371, at concentrations of 1× MIC and 2× MIC, had bacteriostatic activity against MSSA and MRSA after 24 h. At concentrations of 4× MIC and 8× MIC, LCB01-0371 showed bacteriostatic activity, but there was no regrowth at concentrations of 4× MIC and 8× MIC after 24 h of incubation.Open in a separate windowFIG. 2.Time-kill curves of LCB01-0371 and linezolid against S. aureus giorgio (MSSA) and S. aureus p125 (MRSA).The protective efficacy of LCB01-0371 against systemic infections in mice was compared with that of linezolid (Table (Table2).2). When administered orally, LCB01-0371 showed more-potent protective effects than linezolid against systemic infections caused by Gram-positive and Gram-negative bacteria. Against infection caused by S. aureus giorgio (MSSA), the ED50s of LCB01-0371 and linezolid were 4.53 and 8.05 mg/kg of body weight, respectively. Against S. aureus p125 (MRSA), LCB01-0371 (ED50, 2.96 mg/kg) was more active than linezolid (ED50, 4.84 mg/kg). Against E. faecalis u810, the ED50s of LCB01-0371 and linezolid were 4.53 and 5.97 mg/kg, respectively. LCB01-0371 (ED50, 2.28 mg/kg) was also more active than linezolid (ED50, 9.10 mg/kg) against S. pneumoniae ATCC 6305. Against H. influenzae hd2, the ED50s of LCB01-0371 and linezolid were 9.96 and 21.43 mg/kg, respectively. In general, the ED50s of LCB01-0371 were well correlated with in vitro MICs.

TABLE 2.

In vivo activities of LCB01-0371 against systemic infection in mice
Microorganism (inoculum, CFU/mousea)Antimicrobial agentbMIC (μg/ml)ED50, mg/kg (95% confidence limit)
S. aureus giorgio, MSSALCB01-037114.53 (∼2.26-7.87)
    (1 × 107)Linezolid28.05 (∼4.70-13.85)
S. aureus p125, MRSALCB01-037112.96 (∼0.00-5.81)
    (1 × 108)Linezolid24.84 (∼0.01-12.66)
E. faecalis u810LCB01-037124.53 (∼2.26-7.87)
    (2 × 108)Linezolid25.97 (∼2.23-7.87)
S. pneumoniae ATCC 6305LCB01-03710.52.28 (∼0.00-4.49)
    (1 × 104)Linezolid19.10 (∼4.92-23.72)
H. influenzae hd2LCB01-037189.96 (∼4.26-16.75)
    (7.5 × 108)Linezolid1621.43 (∼9.99-450.60)
Open in a separate windowaBacterial strains were suspended in 0.9% saline solution containing 5% mucin, except for S. pneumoniae ATCC 6305, which was suspended in 0.9% saline solution.bAntimicrobial agents were administered orally at 1 and 4 h postinfection.Although linezolid has been recognized as an effective antibiotic against infections with Gram-positive bacteria, such as MRSA and VRE, it produced side effects, such as myelosuppression and peripheral neuropathy, in long-term applications (9). Therefore, it is important to develop new oxazolidinones with good safety profiles, broad antibacterial spectrum, improved pharmacokinetic (PK) parameters, and good water solubility for parenteral administration. LCB01-0371 showed good in vitro and in vivo activities against Gram-positive bacteria and had high aqueous solubility and good absorption, distribution, metabolism, excretion, and toxicity (ADMET) and PK profiles (7). In view of its improved antibacterial activities against Gram-positive bacteria and good pharmacokinetic profiles in animals, the clinical usefulness of LCB01-0371 should be established by further studies.  相似文献   

3.
We tested the effect of probenecid and verapamil in chemosensitizing Plasmodium falciparum to 14 antimalarials using the multidrug-resistant strain V1S and the drug-sensitive 3D7. Verapamil chemosensitizes V1S to quinine and chloroquine. Interestingly, probenecid profoundly chemosensitizes V1S to piperaquine. Thus, probenecid could be used to increase piperaquine efficacy in vivo.The modulation of chloroquine (CQ) resistance with the calcium channel blocker verapamil (VPM), antipsychotic drugs, histamine receptor antagonists, and antidepressant agents among others was the first case of resistance modulation to be reported (reviewed in references 6, 8, and 28).Clinical evaluation of some of these agents has been carried out in areas where CQ resistance is moderate (22, 24-26). However, these agents have the disadvantage of being pharmacologically active, with systemic effects that may result in a variety of side effects. In addition, the minimum concentrations of these agents needed to chemosensitize parasites to CQ (usually more than 1 μM of free drug) (1, 4, 11-13) are not achievable in vivo when normal doses are used; therefore, high doses have to be used, with all of the attendant risks of toxicity. All of these limitations may explain why the reversal of CQ resistance has never attained widespread application.We have demonstrated that the uricosuric drug probenecid (PBN) chemosensitizes parasites to antifolate and CQ (16, 18). In this study, we tested the effect of PBN against the aminoquinolines CQ, piperaquine (PPRQ), primaquine (PMQ), desethylamodiaquin (DEAQ), and amodiaquin (AQ); the amino alcohols lumefantrine (LM), mefloquine (MFQ), halofantrine (HLF), and quinine (QN); the antifolates pyrimethamine (PM), chlorcycloguanil (CCG), and methotrexate (MTX); the benzonaphthyridine pyronaridine (PRN); and the sesquiterpene dihydroartemisinin (DHA). We used the chemosensitizer VPM as a comparator.CQ, PMQ, AQ, MFQ, QN, MTX, PBN, and VPM were purchased from Sigma (Poole, Dorset, United Kingdom). PPRQ was a gift from Universal Corporation Limited, Kikuyu, Kenya. DEAQ, DHA, LM, PRN, and HLF were gifts from Steve Ward, Liverpool School of Tropical Medicine, Liverpool, United Kingdom. Drugs were dissolved as suggested by manufacturers. For PBN, after dilution in dimethyl sulfoxide (500 mg/ml), a subsequent serial dilution was carried out; this consisted of fivefold dilution in absolute ethanol followed by twofold dilution in 2% sodium bicarbonate and then dilution in RPMI 1640 medium (PBN crystallizes when diluted directly from dimethyl sulfoxide to RPMI 1640 medium).We employed two reference P. falciparum laboratory strains: V1S, a multidrug-resistant strain (resistant to CQ, PM, and QN) and 3D7, a strain fully sensitive to all tested antimalarials, except LM and PMQ. Cultures were carried out in RPMI 1640 (GIBCO BRL, United Kingdom), and antimalarial activities were expressed as the drug concentration required for 50% inhibition of [3H]hypoxanthine incorporation (IC50) during the 66-h assay (19).Chemosensitization consisted of testing the activity of the antimalarials in the presence of PBN and VPM at the following ratios: 0.8:0.2, 0.6:0.4, 0.4:0.6, and 0.2:0.8. Chemosensitization was measured geometrically by construction of isobolograms and algebraically, by calculating the sum of the minimum fractional inhibitory concentrations (FICs) (2). FIC values <0.5 and falling between 0.5 and 1 denote pronounced and moderate chemosensitization, respectively. FIC values >1.0 indicate the absence of chemosensitization, and antagonism is denoted by FIC values >4.Our data show that FIC of VPM/CQ, VPM/PMQ, and VPM/LM for V1S ranged between 0.5 and 0.71, a clear indication of chemosensitization or a VPM effect, though moderate (Table (Table1).1). The most pronounced effect was observed with DEAQ and QN, with a FIC of <0.5 (Table (Table1).1). No chemosensitization was observed when 3D7 was used.

TABLE 1.

In vitro analysis, using the multidrug-resistant strain V1S, of the combination of VPM and PBN with 14 antimalarial drugsa
Chemosensitizing drugAntimalarial drug type and drug (IC50 [nM]b)FIC at drug ratio of:
Mean FIC (nM)Score
0.8:0.20.6:0.40.4:0.60.2:0.8
VPMAminoquinolines
    CQ (205 ± 17)0.57 ± 0.070.41 ± 0.050.48 ± 0.140.61 ± 0.060.52 ± 0.08Moderate
    PMQ (326 ± 93)0.59 ± 0.200.64 ± 0.030.77 ± 0.090.83 ± 0.460.71 ± 0.19Moderate
    AQ (8 ± 0.4)0.98 ± 0.611.09 ± 0.621.38 ± 0.012.13 ± 1.041.40 ± 0.57No effect
    PPRQ (88 ± 2)1.17 ± 0.161.21 ± 0.251.05 ± 0.271.08 ± 0.171.13 ± 0.21No effect
    DEAQ (42 ± 12)0.50 ± 0.030.51 ± 0.050.41 ± 0.090.46 ± 0.170.47 ± 0.09Pronounced
Amino alcohols
    HLF (9 ± 0.4)1.12 ± 0.271.05 ± 0.201.02 ± 0.351.19 ± 01.10 ± 0.20No effect
    LM (53 ± 8)0.36 ± 0.060.50 ± 0.200.54 ± 0.090.81 ± 0.150.55 ± 0.13Moderate
    MFQ (11 ± 2)0.99 ± 0.061.20 ± 0.261.05 ± 0.171.26 ± 0.641.12 ± 0.29No effect
    QN (133 ± 44)0.50 ± 0.150.34 ± 0.030.34 ± 0.090.36 ± 0.170.39 ± 0.11Pronounced
Antifolates
    CCG (79 ± 9)1.06 ± 0.050.91 ± 0.191.00 ± 0.191.01 ± 0.071.00 ± 0.12No effect
    PM (5 × 103 ± 103)1.25 ± 0.081.03 ± 0.050.92 ± 0.201.09 ± 0.181.07 ± 0.13No effect
    MTX (63 ± 18)0.59 ± 0.362.18 ± 0.581.82 ± 0.090.73 ± 0.381.33 ± 0.35No effect
Benzonaphthyridine PRN (11 ± 5)1.27 ± 0.081.19 ± 0.010.75 ± 0.211.07 ± 0.061.07 ± 0.09No effect
Sesquiterpene DHA (2 ± 1)0.86 ± 0.251.15 ± 0.341.24 ± 0.511.02 ± 0.141.07 ± 0.31No effect
PBNAminoquinolines
    CQ0.77 ± 0.010.68 ± 0.170.83 ± 0.080.87 ± 0.100.78 ± 0.08Moderate
    PMQ0.89 ± 0.010.87 ± 0.010.81 ± 0.030.77 ± 0.060.83 ± 0.03Moderate
    AQ1.10 ± 0.080.98 ± 0.071.91 ± 0.052.29 ± 0.381.57 ± 0.15No effect
    PPRQ0.08 ± 0.020.09 ± 0.010.11 ± 0.010.18 ± 0.010.12 ± 0.01Pronounced
    DEAQ0.56 ± 0.060.72 ± 0.141.23 ± 0.141.95 ± 0.051.12 ± 0.10No effect
Amino alcohols
    HLF0.53 ± 0.050.51 ± 0.020.51 ± 0.050.74 ± 0.350.57 ± 0.12Moderate
    LM1.04 ± 0.111.33 ± 0.421.14 ± 0.081.35 ± 0.281.21 ± 0.22No effect
    MFQ0.88 ± 0.020.90 ± 0.060.74 ± 00.62 ± 0.310.78 ± 0.10Moderate
    QN1.11 ± 0.091.06 ± 0.021.32 ± 0.321.30 ± 0.081.20 ± 0.13No effect
Antifolates
    CCG0.69 ± 0.110.61 ± 0.060.58 ± 0.110.69 ± 0.160.64 ± 0.11Moderate
    PM0.35 ± 0.370.14 ± 0.090.14 ± 0.050.19 ± 0.050.20 ± 0.14Pronounced
    MTX1.08 ± 0.020.98 ± 0.130.98 ± 0.281.05 ± 0.021.02 ± 0.11No effect
Benzonaphthyridine PRN2.19 ± 0.824.93 ± 0.387.28 ± 0.529.21 ± 0.575.90 ± 0.57Antagonism?
Sesquiterpene DHA1.18 ± 0.361.20 ± 0.411.05 ± 0.381.03 ± 0.501.12 ± 0.41No effect
Open in a separate windowaData are sums of FICs. Data for combinations with moderate or pronounced chemosensitization are in boldface.bInhibitory concentrations that kill 50% of parasites when drugs are used alone against V1S.The ability of VPM to chemosensitize parasites to CQ and CQ-related drugs in CQ-resistant strains is well established (20, 27); however, this effect is commonly found to be moderate (mean FICs > 0.5) (10, 12, 13, 15), in line with our data. The molecular mechanism of this chemosensitization is well understood. It is the result of the interaction of VPM with the pfcrt mutant (codon 76) allele, blocking the efflux of CQ from the cell (7, 20, 27). Since pfcrt modulates QN and DEAQ susceptibility (5, 9), the chemosensitization of parasites to QN and DEAQ could also result from the interaction of VPM with pfcrt.In our previous work, we showed that the uricosuric PBN increases the activity of antifolates in antifolate-sensitive and -resistant strains and chemosensitizes V1S to CQ (PBN effect) (18). We have confirmed these effects (Table (Table1),1), and PBN chemosensitizes 3D7 to PM and CCG (FIC < 1). In addition, PBN moderately chemosensitizes V1S to CQ, PMQ, HLF, and MFQ. The PBN effect on CQ is likely to be associated with the pfcrt mutant allele (18). Apart from the PBN effect on antifolate, chemosensitization occurs in parasites with higher drug IC50s and thus in parasites with drug resistance or reduced-drug-susceptibility phenotypes. However, this seems to be not always the case. For instance, the IC50 of HLF against 3D7 (14 ± 1 nM) was approximately 50% times higher than that against V1S (9 ± 0.4 nM). Yet VPM chemosensitization occurred in V1S but not in 3D7. More interestingly, the IC50 of LM against 3D7 was two times higher, and that of PMQ was five times higher, than those against V1S (104 ± 0.41 versus 53 ± 8 and 1,689 ± 107 versus 326 ± 93, respectively); thus, one would expect chemosensitization to occur in 3D7, yet it occurred in V1S. Thus, chemosensitization is not primarily associated with the parasite IC50 range. We conclude that the preexistence of a multidrug resistance phenotype is the determinant, supporting the observation that pfcrt and pfmdr1 mutants are, at least partly, involved in this chemosensitization (6, 8, 28).Interestingly, we have also found that PNB chemosensitizes V1S to PPRQ, with a mean FIC of 0.12 ± 0.0, the most pronounced chemosensitization effect observed in our study. This is an interesting finding. Indeed, for PBN, unlike VPM, the concentration required to increase drug efficacy in vivo can be achieved when a safe and normal dose is used. Indeed, the use of a normal dose of 2 g of PBN can yield a free and pharmacologically active PBN concentration between 50 and 150 μM (http://www.medscape.com). On the other hand, VPM, when used at normal dose (80 and 160 mg/day), gives rise to around 25 to 50 nM only of free drug (http://www.medscape.com) (3, 14). The relationship between PBN and VPM concentrations and in vitro chemosensitization (Fig. (Fig.1)1) indicates that up to 200 nM of VPM does not exert any effect on QN (the drug harboring the most pronounced chemosensitization with VPM), yet PBN concentrations that are >50 μM are associated with a pronounced decrease in PPRQ IC50 values (a 20-fold increase in activity) (Fig. (Fig.1).1). These observations show the potential of PBN as a chemosensitizer of malaria parasites to antimalarials in vivo, as already demonstrated when this agent was tested with the antifolate pyrimethamine-sulfadoxine (21, 23). Thus, the pronounced chemosensitization of parasites to PPRQ by PBN could be of clinical significance. PPRQ has been combined with DHA, and the drug is known as Artekin. This new combination is undergoing clinical evaluation and is likely to become an alternative to artemether-LM (Coartem), the current drug of choice in the treatment of malaria (17). Thus, if these data on V1S can be extended to field isolates, the chemosensitization of PPRQ by PBN opens up the possibility of using this agent to increase the efficacy of Artekin in clinical situations.Open in a separate windowFIG. 1.Isobolograms representing the in vitro activity of the combination of VPM and QN (A) and of PBN and PPRQ (B) against the multidrug-resistant strain V1S. In panel A, the arrow on the x axis represents the concentration of VPM achieved in vivo when a normal dose is used, which is four to eight times higher than the concentration of unbound VPM. At this concentration, the IC50 of QN is reduced only from 157.5 to 137 nM. In panel B, the arrow represents the achievable free PBN concentration (150 ìM), and at this concentration, PBN increases PPQR activity almost 20-fold, from 78 to 4 nM.  相似文献   

4.
Amphotericin B (AMB) concentrations were determined in pulmonary epithelial lining fluid (ELF) of 44 critically ill patients, who were receiving treatment with liposomal AMB (LAMB) (n = 11), AMB colloidal dispersion (ABCD) (n = 28), or AMB lipid complex (ABLC) (n = 5). Mean AMB levels (± standard errors of the means) in ELF amounted to 1.60 ± 0.58, 0.38 ± 0.07, and 1.29 ± 0.71 μg/ml in LAMB-, ABCD-, and ABLC-treated patients, respectively (differences are not significant).Invasive pulmonary mycoses exhibit a high mortality, particularly in critically ill patients (19). Amphotericin B (AMB) lipid formulations—liposomal AMB (AmBisome; Gilead) (LAMB), AMB colloidal dispersion (Amphotec [Three Rivers] and Amphocil [Torrex-Chiesi]) (ABCD), and AMB lipid complex (Abelcet; Zeneus) (ABLC)—display differences in plasma pharmacokinetics and tissue distribution (15, 26, 28). During treatment with AMB lipid formulations, AMB concentrations were investigated in epithelial lining fluid (ELF), which is a well-established model for pulmonary drug penetration (1-4, 6-10, 12, 17).This study was approved by the local ethics committee. Patients on lipid-formulated AMB requiring bronchoalveolar lavage (BAL) were enrolled (Table (Table1).1). AMB concentrations were assessed in 8-ml aliquots of BAL samples obtained by a standard procedure (16). BAL fluid was concentrated by evaporation, and AMB was quantified by high-performance liquid chromatography as described previously, with modifications for BAL samples (13). The concentrations were assessed by using a linear standard curve (R between 0.995 and 0.999), obtained from standards comprising 0.9% saline solution spiked with AMB. The lower detection limit of AMB in BAL fluid was 0.005 μg/ml. The assay has been found to be linear over the concentration range of 0.005 to 2.5 μg/ml for AMB in BAL fluid. The intraday and interday precisions were 3.2% and 4.7%, respectively. AMB concentrations in ELF were calculated by the urea dilution method (23), AMBELF = AMBBAL × (ureaPLA/ureaBAL), where AMBELF is the AMB concentration in ELF, AMBBAL is the AMB concentration in BAL fluid, ureaPLA is the urea concentration in plasma, and ureaBAL is the urea concentration in BAL fluid (23). Two ml of the BAL fluid was separated for urea quantification, which was performed using an enzymatic assay (urea/blood urea nitrogen; Roche) as with plasma.

TABLE 1.

Demographic and clinical characteristics of patientsa
CharacteristicValue for treatment group
LAMBABCDABLC
Total subjects11285
Mean age in yr46 ± 450 ± 355 ± 5
Sex
    Male9144
    Female2141
Mean wt (kg)68 ± 463 ± 271 ± 8
Main diagnosis
Hematological disorder6182
    Acute myeloid leukemia23
    Other hem. malignancy282
    Lymphoma27
Solid-organ transplantation241
    Liver23
    Heart1
    Kidney1
Solid tumor211
    Carcinoma of lung1
    Brain tumor1
    Skin tumor1
    Pharynx cancer1
Liver cirrhosis41
Other21
Laboratory values
    Creatinine (mg/dl)0.94 ± 0.101.33 ± 0.171.06 ± 0.37
    Bilirubin (mg/dl)6.51 ± 2.8110.56 ± 2.5611.65 ± 6.26
    Prothrombin time (%)76 ± 762 ± 473 ± 10
AMB treatment
    Duration (days)6.1 ± 0.98.8 ± 1.55.6 ± 2.7
    Daily dose (mg)309 ± 22279 ± 16300 ± 47
    Daily dose (mg/kg)4.55 ± 0.234.46 ± 0.194.25 ± 0.58
    Cumulative dose (mg)1,688 ± 2852,176 ± 3402,061 ± 1,259
    Time from start of last infusion to sampling (h)22.0 ± 12.712.6 ± 2.57.3 ± 3.1
Open in a separate windowaMeans ± standard errors of the means. Creatinine, plasma creatinine; normal range, 0.70 to 1.20 mg/dl. Bilirubin, plasma bilirubin; normal range, 0.00 to 1.28 mg/dl. Prothrombin time, normal range, 70 to 130 %. Duration, duration of treatment with lipid-formulated AMB. The time from start of last infusion to sampling was variable, since BALs were scheduled according to clinical requirements. The infusion time was 4 h. When AMB treatment was started at the intensive care unit, the choice of AMB formulation was made by randomization. In patients already on AMB at admission, the respective therapy was continued. Hem., hematological.Arterial blood samples were simultaneously taken for measurement of plasma AMB and urea concentrations. In patients on LAMB or ABCD therapy, the lipid-associated fractions were separated from AMB that had been liberated from its lipid encapsulation. AMB was measured with high-performance liquid chromatography as described previously (13).Statistical analysis was performed using the Statistica software program, version 5. The differences between total AMB concentrations in plasma and in ELF were analyzed by using the Wilcoxon matched pairs test. For comparisons between the lipid formulations, the Mann-Whitney U test was applied.Forty-four patients were enrolled: 11 patients on LAMB, 28 on ABCD, and 5 on ABLC. Table Table22 displays the ELF and plasma concentrations of AMB and the penetration ratios. In the entire study population and in LAMB-treated patients, ELF concentrations correlated with plasma levels (r = 0.68, P < 0.001, and r = 0.66, P = 0.04, respectively). In the LAMB group, this correlation was even more significant when liberated AMB was considered (r = 0.89; P < 0.001). A positive correlation between the time from last infusion to sampling and the penetration ratio was found during LAMB (r = 0.75; P = 0.01) and ABLC (r = 0.95; P = 0.01) treatments.

TABLE 2.

Concentrations of AMB in plasma and in ELFa
ParameterValue for treatment groupb
LAMBABCDABLC
Mean concn in ELF ± SEM (μg/ml)1.60 ± 0.58**0.38 ± 0.07*1.29 ± 0.71
Mean concn in plasma ± SEM (μg/ml)
    Liberated1.08 ± 0.310.57 ± 0.09NA
    Lipid associated4.11 ± 1.61‡0.54 ± 0.15‡NA
    Total5.17 ± 1.89**1.12 ± 0.21*0.48 ± 0.18
Mean penetration ratio ± SEM (%)
    ELF/total plasma61 ± 25†125 ± 52†447 ± 224†
    ELF/liberated plasma154 ± 44153 ± 53NA
Highest ELF concn (μg/ml)6.011.706.97
Respective penetration ratio (%)701,371942
Respective time from start of AMB infusion to BAL (h)6.2524.005.50
Respective cumulative dose (mg)2,60090011,700
Concn measured at maximum time from start of AMB infusion to BAL0.350.280.84
    Time from start of AMB infusion to BAL (h)146.0048.0019.50
    Penetration ratio (%)2421801,276
    Cumulative dose (mg)1,7751,375150
Open in a separate windowaNA, not available. For the ABLC group, the chromatographic separation of lipid-associated and liberated AMB in plasma was not feasible. For patients who underwent more than one BAL, the mean concentration in ELF and penetration ratio were applied for statistical calculations. A P value of <0.05 was regarded as statistically significant. The penetration ratio was defined as the AMB concentration in ELF/simultaneous total AMB plasma level (%). The differences in concentrations in ELF between the treatment groups did not reach significance (LAMB vs. ABCD, P = 0.21; ABCD vs. ABLC, P = 0.08; LAMB vs. ABLC, P = 0.95).b**, concentrations in ELF in were significantly lower than the respective total levels in plasma (P = 0.001); *, AMB concentrations in ELF were significantly lower than the respective total levels in plasma (P = 0.01); ‡, in LAMB therapy, the levels of the lipid-encapsulated AMB fraction exceeded those in the ABCD group highly significantly (P < 0.001); †, the penetration ratio was significantly higher for patients on ABLC therapy than for those in the ABCD and LAMB groups (P < 0.05).Inhalation of fungal conidiae is the most common route of infection with molds. During treatment with AMB lipid formulations at standard doses, mean AMB levels in ELF were below 2 μg/ml. For Aspergillus species, the MIC of AMB has been reported to range from 0.25 to 4 μg/ml (14). Thus, in some cases, the MIC exceeds the AMB concentration in ELF. This may contribute to unsatisfying responses sometimes observed, though the impact of target site concentrations in relation to MICs is controversial. ELF concentrations are markedly lower than AMB levels in whole lung tissue (32.6 μg/g after ABCD treatment) (26). Whole tissue samples, however, comprise various compartments and potential targets of fungal invasion, such as different cells, extracellular matrix, and blood vessels.The differences in the underlying diseases, the limited number of patients that differed between the groups, slight differences in doses, and various intervals between AMB infusion and BAL are limitations of our study. In the LAMB group and in the ABLC group, penetration of ELF increased with this interval. Similarly, a slow increase in concentrations in lung tissue over 25 h was observed after LAMB infusion (11).A study of rabbits revealed ELF concentrations comparable to our human data (2.28, 0.68, and 0.90 μg/ml after LAMB, ABCD, and ABLC treatment, respectively) (17).In pleural effusion and ascites, where mainly liberated AMB is found, concentrations were even lower than those in ELF (27, 28). In vitro investigations suggest an influence of phosphatidylcholine liposomes within ELF on membrane oxidation and nitration that could potentially affect the activity of lipid-associated antimicrobial agents in vivo (25). Unlike the case with plasma and with body fluids, separation of liberated and lipid-encapsulated AMB was not feasible in ELF. For LAMB and ABCD, the penetration ratios of liberated AMB were similar, suggesting that mainly liberated AMB penetrates ELF.Lung transplant recipients on prophylaxis with nebulized LAMB (several 25-mg doses) displayed concentrations in ELF of ∼10 μg/ml 2 days after inhalation and 3 to 4 μg/ml after 2 weeks. AMB was undetectable in plasma of all but one patient, suggesting a poor systemic absorption and penetration into deeper lung compartments (20).Penetration of ELF by voriconazole was studied for lung transplant recipients on prophylactic treatment, revealing various concentrations (0.29 to 83.32 μg/ml; mean penetration ratio, 1,100%) (5). In healthy volunteers who had received posaconazole at the standard dosage for 8 days, a mean concentration in ELF of 1.86 μg/ml was measured (10). Treatment with the high-molecular-weight lipopeptide micafungin (150 mg daily for 3 days) resulted in concentrations in ELF of ∼0.5 μg/ml and an accumulation in alveolar macrophage cells (8.4 to 14.6 μg/ml) (21). Similarly, AMB in either a deoxycholate or a lipid formulation accumulates in cells of the reticuloendothelial system, particularly in alveolar macrophage cells, as shown in animal and in vitro experiments (17, 18, 22, 24). In the present study, AMB was not separately quantified in alveolar macrophages.In conclusion, treatment with AMB lipid formulations at standard doses yields ELF concentrations moderately above or even below MICs of relevant fungal pathogens. ELF levels are much lower than AMB concentrations in lung tissue samples. Further investigations should address the impact of target site penetration of antifungals on the therapeutic outcome in invasive pulmonary mycoses.  相似文献   

5.
Antimicrobial susceptibilities of 23 strains of Desulfovibrio spp. were tested by Etest. Generally, Desulfovibrio spp. were highly susceptible to sulbactam-ampicillin, meropenem, clindamycin, metronidazole, and chloramphenicol: MIC90s of 6, 4, 0.19, 0.25, and 8 μg/ml, respectively. In addition, these strains generally showed high MICs to piperacillin and piperacillin-tazobactam. Desulfovibrio fairfieldensis (eight strains) was the species least susceptible to most agents, especially β-lactams, and was the only species resistant to fluoroquinolones. Desulfovibrio desulfuricans strain Essex 6 isolates were less susceptible to β-lactams than D. desulfuricans strain MB isolates.Desulfovibrio spp. are gram-negative anaerobes and a type of dissimilatory sulfate-reducing bacteria. Most established species of Desulfovibrio are distributed in the environment, but some Desulfovibrio spp. reside in oral cavities and intestinal tracts of animals, including humans (1, 17). In 1996, Tee et al. first reported the isolation of Desulfovibrio species from a blood culture of a patient with cholecystitis and suggested that Desulfovibrio species might act as an opportunistic pathogen (16). Since then, several case reports that suggested Desulfovibrio may be the causative organism have been published (5, 7, 8, 10, 11, 14, 15). In the published case reports, most Desulfovibrio strains were isolated from the blood of patients who suffered from a brain abscess, appendicitis, intra-abdominal abscess, or abdominal wall abscess, while some were isolated from peritoneal fluid or the pus from abscesses of various origins. Gibson et al. have also reported that Desulfovibrio spp. might be associated with ulcerative colitis (4). Furthermore, Langendijk et al. found that some Desulfovibrio spp. may be associated with the early onset of periodontitis, rapidly progressive periodontitis, adult periodontitis, and refractory periodontitis (6).Presently, four Desulfovibrio spp. (D. fairfieldensis, D. desulfuricans, D. piger, and D. vulgaris) are recognized to be associated with humans (5, 8). They are slow growers and relatively difficult to isolate from clinical specimens by a conventional approach. Identification to the species level without molecular techniques is considerably difficult. Lozniewski et al. tested the susceptibility of 16 clinical isolates of Desulfovibrio spp. and showed the broad MIC range of some antimicrobial agents (9). However, they did not identify their isolates to the species level. Furthermore, Warren et al. tested 18 clinical Desulfovibrio isolates, which were identified to the species level; however, only scattered strains of D. desulfuricans were included in their study (18). Therefore, the information available regarding the antibiogram of human Desulfovibrio isolates is extremely limited and incomplete.Considering the present situation, it is important and necessary to obtain additional information concerning the antibiogram of Desulfovibrio spp. for the empirical treatment of anaerobic infections in which Desulfovibrio spp. might be involved. Our laboratory has collected Desulfovibrio strains from various human specimens, both clinical and nonclinical, over the past several years. After molecular identification by 16S rRNA sequencing, 13 isolates of D. desulfuricans strains Essex 6 and MB were included in our collection. Therefore, in this study, we performed the Etest to obtain additional knowledge regarding the antimicrobial susceptibilities of Desulfovibrio species.Twenty-three strains of Desulfovibrio spp. were tested. These strains were isolated from human specimens and identified by classical phenotypic and molecular methods such as 16S rRNA gene sequencing. Specimens from which isolates were obtained and the number of isolates are as follows: mucosal swab of a patient with pouchitis, 6; stool specimens,7; tongue coating, 6; appendicitis, 3; and blood culture, 1. The following four Desulfovibrio strains were used as reference strains: D. fairfieldensis ATCC 700045, D. desulfuricans Essex 6 (ATCC 29577T), D. desulfuricans MB (ATCC 27774), and D. piger ATCC 29098T. Bacteroides fragilis ATCC 25285T was used as the quality control strain. A special agar medium for Desulfovibrio, which was formulated and named “Desulfovibrio agar” (DA) by the author, was also used as a susceptibility test medium. The ingredients of DA were as follows: polypeptone, 15 g; soya-peptone, 7.5 g; yeast extract, 7.5 g; beef extract, 7.5 g; l-cysteine HCl, 0.75 g; ferric ammonium citrate, 0.75 g; dextran sodium sulfate, 10 g; and agar powder, 15 g per liter of distilled water (pH 7.0). In our preliminary experiment, it was confirmed that Desulfovibrio spp. grew more rapidly on the surface of DA than on the surface of Brucella blood agar (BBA) and formed a blackish or black halo around the colonies on this medium due to H2S production followed by FeS formation. DA was also used for a pour plate method as described below.Nine antianaerobic antimicrobial agents were used: sulbactam-ampicillin, piperacillin, piperacillin-tazobactam, cefoxitin, cefotaxime, meropenem, clindamycin, chloramphenicol, and metronidazole. Eleven non-antianaerobic antimicrobial agents were also included in this study: ampicillin, cefoperazone, sulbactam-cefoperazone, cephalothin, ceftazidime, cefpirome, erythromycin, minocycline, ciprofloxacin, levofloxacin, and sulfamethoxazole-trimethoprim.The Etest was first performed according to a commonly used protocol. Namely, the bacterial suspension was inoculated on the surface of BBA by swabbing. The bacterial suspension was adjusted to McFarland no. 1 standard solution. However, this protocol is not always appropriate for testing Desulfovibrio species because achieving confluent growth of the inoculated organisms is difficult on the surface of BBA. Then we used the DA medium. Although Desulfovibrio spp. grew better on the surface of this medium, it was still difficult to achieve confluent growth by swab inoculation. Finally, we used the pour plate method described by Wilkins et al. (19). Briefly, 100 μl of bacterial suspension (McFarland no. 1 standard) was mixed with 20 ml DA medium, which was maintained at 50°C and then poured into petri dishes. When this pour plate method was used, a clear transparent elliptical zone of growth inhibition was observed within 48 h of incubation, with another zone of blackened medium surrounding this clear zone due to H2S formation. After 48 h of anaerobic incubation (atmosphere of 80 to 85% N2, 5 to 10% CO2, and 10% H2), the MIC was read as the concentration at which the border of the elliptical zone of growth inhibition intersected the scale in the Etest strips. Results for reference strains using the DA agar method were comparable to those of the standard BBA method (data not shown).β-Lactamase production was tested using the nitrocefin hydrolysis test (Cefinase; Becton-Dickinson Co., Ltd.) according to the manufacturer''s instructions.Regardless of the species, all of the Desulfovibrio spp. tested in this study were susceptible to five antianaerobic agents with low MIC90s, i.e., sulbactam-ampicillin (6 μg/ml), clindamycin (0.19 μg/ml), meropenem (4 μg/ml), metronidazole (0.75 μg/ml), and chloramphenicol (8 μg/ml). On the other hand, these strains showed high MIC90s toward the other antianaerobic agents: piperacillin (>256 μg/ml), piperacillin-tazobactam (>256 μg/ml), cefoxitin (>256 μg/ml), and cefotaxime (>256 μg/ml) (Table (Table1).1). The high susceptibilities of Desulfovibrio spp. to clindamycin, metronidazole, chloramphenicol, and imipenem have been noted in previously published reports (9, 18). In this study, one D. fairfieldensis strain that was intermediately resistant to meropenem was recognized. With regard to the susceptibility of Desulfovibrio spp. to carbapenems, previous reports have demonstrated their uniform susceptibility to imipenem (9, 18). However, another report demonstrated that D. fairfieldensis strains were less susceptible to ertapenem, with fairly high MIC90 (>32 μg/ml) and MIC50 (>32 μg/ml) values (18). Further studies are required to determine the susceptibilities of D. fairfieldensis strains to an array of carbapenems, which are now very often chosen as empirical treatment for serious infections.

TABLE 1.

Susceptibilities of Desulfovibrio isolates from humans to 20 antimicrobial agents
Species and antimicrobial agent (no. of isolates)MIC (μg/ml)a
Range50%90%
Desulfovibrio spp. (n = 23)
    Antianaerobic
        Sulbactam-ampicillin0.064-60.386
        Piperacillin32->256>256>256
        Piperacillin-tazobactam16->256>256>256
        Cefoxitin16->256>256>256
        Cefotaxime0.047->2561.5>256
        Meropenem0.023-120.194
        Clindamycin0.016-0.250.1250.19
        Chlorampenicol1.5-1248
        Metronidazole<0.016-1.50.1250.25
    Others
        Ampicillin0.19-240.758
        Cefoperazone12->256>256>256
        Sulbactam-cefoperazone1->2564>256
        Cefalothin16->256>256>256
        Ceftazidime2->2566>256
        Cefpirome0.5->2564>256
        Erythromycin0.064-40.51.5
        Minocycline0.19-24624
        Ciprofloxacin0.125->320.5>32
        Levofloxacin0.094->320.75>32
        Sulfamethoxazole-trimethoprim>32>32>32
D. fairfieldensis (n = 8)b
    Antianaerobic
        Sulbactam-ampicillin4-6
        Piperacillin>256
        Piperacillin-tazobactam>256
        Cefoxitin64->256
        Cefotaxime0.047->256
        Meropenem2-12
        Clindamycin0.032-0.19
        Chloramphenicol3-6
        Metronidazole<0.016-0.38
    Others
        Ampicillin4-8
        Cefoperazone>256
        Sulbactam-cefoperazone24->256
        Cefalothin>256
        Ceftazidime>256
        Cefpirome>256
        Erythromycin0.5-4
        Minocycline4-24
        Ciprofloxacin>32
        Levofloxacin>32
        Sulfamethoxazole-trimethoprim>32
D. desulfuricans Essex 6 (n = 7)c
    Antianaerobic
        Sulbactam-ampicillin0.125-1.5
        Piperacillin>256
        Piperacillin-tazobactam16->256
        Cefoxitin>256
        Cefotaxime0.047-6
        Meropenem0.064-0.25
        Clindamycin0.094-0.25
        Chloramphenicol1.5-12
        Metronidazole0.064-1.5
    Others
        Ampicillin0.19-2
        Cefoperazone16->256
        Sulbactam-cefoperazone3->256
        Cefalothin3->256
        Ceftazidime3->256
        Cefpirome0.5-32
        Erythromycin0.38-0.75
        Minocycline2-12
        Ciprofloxacin0.19-2
        Levofloxacin0.25-3
        Sulfamethoxazole-trimethoprim>32
D. desulfuricans MB (n = 6)d
    Antianaerobic
        Sulbactam-ampicillin0.064-0.25
        Piperacillin32->256
        Piperacillin-tazobactam32-96
        Cefoxitin16->256
        Cefotaxime0.5-1.5
        Meropenem0.023-0.064
        Clindamycin0.094-0.19
        Chloramphenicol1.5-4
        Metronidazole0.016-0.38
    Others
        Ampicillin0.19-0.38
        Cefoperazone12-24
        Sulbactam-cefoperazone1-3
        Cefalothin2-6
        Ceftazidime2-4
        Cefpirome0.5-2
        Erythromycin0.38-1
        Minocycline3-8
        Ciprofloxacin0.125-0.38
        Levofloxacin0.25-0.5
        Sulfamethoxazole-trimethoprim>32
D. piger (n = 2)e
    Antianaerobic
        Sulbactam-ampicillin0.5-1
        Piperacillin>256
        Piperacillin-tazobactam>256
        Cefoxitin16
        Cefotaxime1-1
        Meropenem0.032-0.19
        Clindamycin0.016-0.016
        Chloramphenicol1.5-3
        Metronidazole0.032-0.5
    Others
        Ampicillin12-32
        Cefoperazone12-32
        Sulbactam-cefoperazone1-1.5
        Cefalothin96->256
        Ceftazidime3-3
        Cefpirome0.75-4
        Erythromycin0.016-0.125
        Minocycline0.19-1
        Ciprofloxacin0.125-0.19
        Levofloxacin0.095-0.125
        Sulfamethoxazole-trimethoprim>32
Open in a separate windowa50% and 90%, MIC50 and MIC90, respectively.bOne isolate from pouchitis, five from stool specimens, and two from tongue coating.cOne isolate from appendicitis, two from stool specimens, and four from tongue coating.dOne isolate from blood, three from pouchitis, and two from appendicitis.eBoth isolates from pouchitis.Fifteen Desulfovibrio strains, including 8 D. fairfieldensis strains and 7 D. desulfuricans Essex 6 isolates, were highly resistant to both piperacillin and cefoxitin: MICs of >256 μg/ml for piperacillin and 64 to >256 μg/ml for cefoxitin, respectively. D. desulfuricans MB isolates were susceptible or intermediately susceptible to piperacillin and cefoxitin, except for one resistant isolate. Most of the D. fairfieldensis strains were cefotaxime resistant (MIC range, >265 μg/ml), but the other species showed low MICs to cefotaxime (MIC range, 0.5 to 6 μg/ml).All of the Desulfovibrio strains tested showed relatively low MIC90s to erythromycin (MIC90, 1.5 μg/ml), slightly higher MIC90s to ampicillin (MIC90, 8 μg/ml) and minocycline (MIC90, 24 μg/ml), and consistently higher MIC90s to sulfamethoxazole-trimethoprim (MIC90, >32 μg/ml). Their susceptibilities to the remaining seven non-antianaerobic agents were strain dependent with broad MIC ranges.Eight D. fairfieldensis strains were obviously highly resistant to four cephems (cephalothin, cefoperazone, ceftazidime, and cefpirome), sulbactam-cefoperazone, and two fluoroquinolones (ciprofloxacin and levofloxacin). This study, together with the previous report, demonstrated that among three species of the genus Desulfovibrio, D. fairfieldensis was the species least susceptible to antimicrobial agents. D. fairfieldensis strains from our collections were highly resistant to narrow- and broad-spectrum cephems (cephalothin, cefoxitin, cefoperazone, cefotaxime, ceftazidime, and cefpirome) with no detectable β-lactamase by the nitrocefin tests. The mechanism underlying the trend of resistance of this species to β-lactams remains to be elucidated. D. desulfuricans Essex 6 isolates showed relatively lower MICs to fluoroquinolones (0.19 to 3 μg/ml). D. desulfuricans Essex 6 isolates were more resistant to β-lactam antibiotics than were D. desulfuricans MB isolates. Morin et al. have demonstrated that some D. desulfuricans strains possess a β-lactamase gene (blaDES-1) and suggested that the gene is associated with a certain subtype of D. desulfuricans, although they did not discriminate D. desulfuricans Essex 6 from D. desulfuricans MB isolates in their study (12). In our study, it is notable that all four β-lactamase producers from among the D. desulfuricans strains were the D. desulfuricans Essex 6 isolates. Therefore, the existence of blaDES-1 in our D. desulfuricans strains should be investigated.The conventional Etest method is simple and easy to perform and may be suitable for all anaerobes, including some slow growers (2, 3). Desulfovibrio spp. have a tendency to form tiny, viscous, and pitting colonies on the agar surface: some modification was required to obtain a good performance of the Etest. We proposed to use the DA agar medium and to incorporate the inocula in the pour plate instead of swabbing them. In these bacterium-impregnated agar plates, after only 2 days of incubation, all of the Desulfovibrio strains grew homogeneously and formed clear, black-edged, elliptical zones of inhibition. The reproducibility of the MIC determinations by this method was well acceptable, since the MICs were identical or varied within a twofold dilution range for several strains tested, including for the reference strains. The MICs obtained for the quality control strains (B. fragilis ATCC 25285) were within the ranges of the CLSI reference values (13; data not shown).It is essential to increase research efforts to isolate Desulfovibrio strains from clinical specimens in order to establish a more useful antibiogram of the Desulfovibrio species. However, in the present situation, in addition to the published information concerning the antimicrobial susceptibilities of Desulfovibrio, the findings of this study may also be useful for clinicians in treating patients with Desulfovibrio bacteremia and patients with mixed infections associated with endogenous Desulfovibrio spp.  相似文献   

6.
Klebsiella pneumoniae isolates frequently contain complex mixtures of blaSHV alleles. A high-resolution melting-based method for interrogating the extended-spectrum activity conferring codon 238 and 240 polymorphisms was developed. This detects minority extended-spectrum β-lactamase-encoding alleles, allows estimation of allele ratios, and discriminates between single and double mutants.High-resolution melting (HRM) analysis is showing considerable promise as a method for rapid and cost-effective interrogation of single nucleotide polymorphisms (SNPs) (5). There have been numerous reports of the successful use of HRM to discriminate homozygotes and heterozygotes in humans (5). However, the use of HRM to analyze allele mixtures that are not 1:1 and to reveal the presence of minority alleles has been little explored.The increasing prevalence of bacteria producing extended-spectrum β-lactamases (ESBLs) has become a significant problem facing health care across the world. Many ESBLs are derived from plasmid-encoded SHV or TEM β-lactamases (12-14). The conversion of a non-ESBL SHV enzyme into an ESBL is nearly always associated with a G238S (GGC→AGC) substitution (numbering according to that of Ambler et al. [1]), while a further extension of the spectrum of activity is mediated by an E240K (GAG→AAG) substitution (12). Although there are many other substitutions reported, mutations of codons 238 and 240 are by far the most significant for ESBL activity of the blaSHV family (http://www.lahey.org/Studies/).An unusual aspect of the biology of the SHV-encoding gene, blaSHV, is that it is present on the chromosome of most Klebsiella pneumoniae strains (2, 3). However, it is also disseminated on plasmids, and most SHV ESBLs are plasmid encoded. It appears that there have been two recent mobilizations of non-ESBL blaSHV from the K. pneumoniae chromosome onto plasmids, the mobilized genes being blaSHV-1 and blaSHV-11, which differ only at codon 35 (7). The G238S mutant-encoding derivatives of these are blaSHV-2 and blaSHV-2a, respectively, while the G238S and E240K double-mutant-encoding derivatives are blaSHV-5 and blaSHV-12, respectively. These six blaSHV variants are all abundant.The existence of both chromosome and plasmid-borne blaSHV means that many K. pneumoniae strains harbor mixtures of blaSHV alleles, and the ratios between alleles can vary widely (6, 8, 9). It has also been determined that the presence of a minority ESBL-encoding allele confers an ESBL-positive phenotype (9). This provides a diagnostic challenge.An allele-specific real-time PCR (sometimes known as kinetic PCR) method for interrogating the blaSHV codon 238 and 240 SNPs has previously been reported (8, 9). This has proven effective, but it requires four separate PCRs. Also, it is difficult to calibrate for interlab comparison because of the effects of small batch-to-batch variations in primer concentration and quality (unpublished data). In contrast, the allele discrimination in an HRM-based assay takes place at the end of the PCR, so the format is inherently more robust.Here, we demonstrate a single-tube HRM-based assay for codon 238 and 240 mutations in the blaSHV gene and its comparison with allele-specific real-time PCR. To support this assay, data analysis methods were developed for the inference of confidence limits regarding whether or not an analyte contains a mutant allele. This procedure is similar in some respects to the probe-based melting analyses of Chia et al. and Randegger and Hachler (4, 15), but it is inherently simpler and provides more information.Two isolate collections of K. pneumoniae were used in this study (Table (Table1).1). Isolates A1 to L1 are clinical isolates from the Princess Alexandra Hospital (PAH) in Brisbane, Australia, have previously been characterized, and contain various mixtures of blaSHV-11, blaSHV-2a, and blaSHV-12 (8, 9, 11, 16). Isolates 14 to 121 are derived from the SENTRY collection (10). All strains were cultured in Luria-Bertani (LB) broth and stored in cryovials with 12% glycerol at −80°C. The isolates that are asterisked in Table Table11 are derivatives of the clinical isolates that have been subjected to selection for resistance to 128 μg/ml cefotaxime (8).

TABLE 1.

blaSHV allele distributions and results from allele-specific real-time PCR and HRM assays
Origin and isolateESBL phenotype% blaSHV alleles
Kinetic PCRΔCTb
WT or single- or double-mutant strainDifference in graph amplitude
WT238 mutation240 mutationCodon 238Codon 240Combined
PAH
    B2Negative100004.9710.0715.04WTc3.27
    K2Negative100005.385.8611.24WTc2.52
    J3Negative100005.227.0212.24WTc1.97
    A1Positive50500−0.326.606.28Singlec−10.79
    D1Positive257510−1.196.265.07Singlec−12.96
    J2Positive8020201.071.362.43Doublec−10.29
    L1Positive604035−0.27−0.11−0.38Doublec−17.40
    I1Positive109065−1.85−3.01−4.86Doublec−25.91
    B1Positive−1.106.715.61Singled−12.07
    F1Positive−1.187.366.18Singled−12.19
    B1aPositive−1.946.985.04Singled−16.16
    F1aPositive−1.927.935.01Singled−14.92
SENTRY
    30Negative5.396.3511.74WTd1.10
    54Negative4.915.9610.87WTd0.13
    70Negative5.246.1811.42WTd−0.24
    85Negative3.739.6813.41WTd−1.03
    102Negative4.7111.0015.71WTd0.13
    104Negative4.137.3511.48WTd1.46
    105Negative4.107.4311.53WTd−1.01
    106Negative3.9011.8015.70WTd0.40
    107Negative3.4510.6514.10WTd−0.11
    108Negative3.407.7511.15WTd−2.19
    109Negative3.6910.2313.92WTd0.95
    110Negative4.917.0411.95WTd−0.24
    113Negative4.1712.7216.89WTd−0.73
    114Negative3.9212.7916.71WTd−1.87
    115Negative3.3510.9314.28WTd−2.23
    116Negative4.8812.7917.67WTd−1.44
    120Negative4.677.0711.74WTd−0.22
    121Negative5.306.8211.58WTd−0.51
    14Positive0.740.240.98Doubled−15.28
    18Positive−0.92−4.50−5.42Doubled−24.88
    58Positive0.514.835.34Singled−16.40
    54aPositive0.166.776.93Singled−9.02
    110aPositive−1.857.615.76Singled−13.13
    120aPositive1.987.008.98Singled−4.86
    121aPositive2.066.898.95Singled−4.34
Open in a separate windowaIsolates that were subjected to stepwise selection for resistance to increasing cefotaxime concentrations and are resistant to 128 μg/ml cefotaxime.bShown are the mean ΔCT measurements from three separate experiments.cAllele content was directly demonstrated by analysis of cloned PCR products.dAllele content was deduced on the basis of consistency between the allele-specific real-time PCR, HRM data, and ESBL phenotype.DNA was extracted, using a DNeasy tissue kit (Qiagen), from 2.5-ml cultures grown overnight in LB broth. The resuspended DNA was stored at −20°C. All reactions for both the allele-specific real-time PCR assay and HRM assay were performed in a Rotor-Gene 6000 real-time PCR device (Qiagen).Allele-specific PCR was carried out essentially as previously described by Hammond et al. (9). The reactions were performed in 10-μl volumes, containing SensiMix NoRef master mix at a 1× concentration (Quantace), Sybr green I at a 1× concentration, 5 pmol of common primer (Shv238reverse [5′-CGGCGTATCCCGCAGATAA-3′] or Shv240reverse [5′-CCGGCGGGCTGGTTTAT-3′]), 5pmol of either the wild-type (WT)-specific primer (Shv238wt [5′-CGCCGATAAGACCGGAGCTG-3′] or Shv240wt [5′-GCGCGCACCCCGCTC-3′]) or mutant-specific primer (Shv238mt [5′-CGCCGATAAGACCGGAGCTA-3′] or Shv240mt [5′-GCGCGCACCCCGCTT-3′]), and 50 ng of DNA template. Thermocycling parameters were as described by Hammond et al. (9). The ΔCT values were calculated by subtracting the CT for the WT-specific reaction from the CT for the mutant-specific reaction. It has previously been shown by Hammond and coworkers that ESBL-negative isolates have codon 238 ΔCT values of >2.5 and codon 240 ΔCT values of >5 (8), and these are the cutoffs used for presumptive mutant allele detection by allele-specific real-time PCR.The HRM assay was carried out using primers SHVmutHRM_F (5′-CGCCGATAAGACCGGAGCT-3′) and SHVmutHRM_R (5′-CCGCGCGCACCCCGCT-3′). These were designed to anneal very close to the SNPs, and the amplified fragment was only 39 bp. The small size of this fragment maximizes the differences in melting temperature (Tm) conferred by the SNPs and allows for short cycling times. PCR was performed in a 10-μl reaction, containing SensiMix NoRef master mix at a 1× concentration (Quantace), Sybr green I at 1× concentration, 5 pmol of each primer, Q-solution at a 1× concentration (Qiagen), and 1 ng of DNA template. Thermocycling parameters were as follows: 50°C for 2 min; 95°C for 10 min; 40 cycles of 95°C for 5 s, 65°C for 5 s, and 72°C for 10 s; 95°C for 2 min; and 50°C for 30s. The amplification was followed by HRM from 70 to 86°C at 0.05°C increments, remaining at each step for 2 s. In order to calculate confidence limits for conclusions regarding the sequence of the amplified fragment, the numerical data defining the normalized HRM curves were exported using the “export” function of the software supplied with the Corbett 6000 device, and the mean and standard deviation (SD) were defined at each temperature of the melting protocol. From this, the 95% confidence limits were calculated as the mean ± 1.96 × SD. These numbers were used to define a 95% confidence limit area on the difference graph that is obtained by setting the mean of all of the WT HRM data as the baseline. The amplitudes at the nadirs of the HRM difference graphs were calculated with reference to the mean for all WT isolates.The fully characterized samples for the development of this assay were K. pneumoniae clinical isolates with blaSHV allele ratios that had previously been determined by allele-specific real-time PCR and also by the cloning and analysis of PCR products (5, 10). These were subjected to HRM analysis, and in addition, the allele-specific real-time PCR assays were repeated so as to control for the effect on primer extension efficiency of different primer batches (Table (Table11 and Fig. 1A and B). The ESBL-positive isolates were clearly discriminated from the ESBL-negative isolates, with the mutant samples having lower Tm values, consistent with the G→A substitutions. The ESBL-positive isolates included J2, which has 80% WT alleles and 20% double-mutant (blaSHV-12) alleles. This yielded a very different HRM curve from the WT isolates. Therefore, this method can detect mutant alleles when they in the minority. The clear identification of a minority allele probably reflects the formation of heteroduplexes which depress the Tm. In the difference graph (Fig. (Fig.1B),1B), it can be seen that the double-mutant (blaSHV-12)-containing isolates J2, L1, and I1 can be easily discriminated by eye because the curve nadirs are displaced to the left. However, isolate D1, which contains just 10% double-mutant (blaSHV-12) alleles and a high proportion of single-mutant (blaSHV-2a) alleles, was not discriminated from A1, in which only the WT and single-mutant alleles have been detected. In addition, there was correlation between the difference graph amplitudes and the ratio between the WT and mutant alleles. Because the difference graph amplitudes are derived from melt curves that are always normalized to the same number of arbitrary fluorescence units by the Corbett 6000 software, they can be compared between runs.Open in a separate windowFIG. 1.HRM data. (A) Normalized HRM graphs from the PAH isolates. Isolates containing mutated alleles (A1, D1, J2, L1, and I1) are clearly separated from isolates containing only WT alleles (B2, J3, and K2). (B) Difference graph from the PAH isolates, with the fluorescence of isolate J3 set as the baseline. For each isolate, the number and allelic distribution are stated. (C) Difference graph of the additional 27 SENTRY isolates (18 WT, 7 single mutant, 2 double mutant) isolates. The baseline used was the average fluorescence data of 21 WT (ESBL negative) isolates. The area representing the 95% confidence limit for the WT sequence is delineated with dotted lines. All ESBL-positive isolates yielded difference graph curves outside the 95% confidence limit area. The two curves with nadirs displaced to the left are the two double-mutant isolates, 14 and 18.To further test this method, 27 diverse K. pneumoniae isolates (18 WT, 7 single mutants, and 2 double mutants), from the Asia-Pacific component of the SENTRY program (10) were subject to analysis. All were analyzed by HRM and allele-specific real-time PCR (Table (Table11 and Fig. Fig.1C).1C). Some of these isolates had been subjected to allele-specific real-time PCR previously (8, 9), but all isolates were subjected to this procedure during the course of this study. Once again, the ESBL-positive isolates were discriminated from the ESBL-negative isolates, and isolates with two mutations were discriminated from those with one mutation on the basis of displacement of the nadir of the difference graph to the left. The greater number of samples enabled the power of the HRM assay to indicate allele ratios to be examined more rigorously. The correlation coefficient for the linear regression of the combined kinetic PCR ΔCT values versus the amplitude at the nadirs of the HRM difference graphs for the “single-mutant” ESBL-positive isolates is 0.94 (P < 0.0001). The value for “double-mutant” isolates is 0.98 (P = 0.0023). This confirms that this assay can provide an indication of allele ratios. The data from the 18 SENTRY WT isolates together with the 3 PAH WT isolates (B2, J3, and K2) were used to obtain 95% confidence limits for the HRM curve corresponding to the WT allele. This not only indicates when the presence of a mutation should be called but also facilitates the portability of this method. The Tm of the WT sequence using our Rotor-Gene 6000 device is 79.00 ± 0.15°C. It is our experience that an individual Corbett 6000 device is highly accurate with respect to relative temperatures during runs, but absolute temperature calibration between different Corbett 6000 devices can differ by up to 0.5°C. The 95% confidence limit curves for the WT sequence can easily be adjusted with reference to a Tm determined from a small number of runs of the procedure against a WT sequence on a particular machine.In conclusion, an HRM-based method for interrogating the clinically significant codon 238 and codon 240 SNPs of the blaSHV gene has been developed. The method requires no probes, makes use of a low-cost Sybr green-based master mix, detects mutant alleles if they are in the minority, and provides an indication of the allele ratio. The method costs <$0.60 per sample in materials, even when carried out in duplicate, and takes approximately 1 h.  相似文献   

7.
The in vitro activities of ceftaroline, a novel, parenteral, broad-spectrum cephalosporin, and four comparator antimicrobials were determined against anaerobic bacteria. Against Gram-positive strains, the activity of ceftaroline was similar to that of amoxicillin-clavulanate and four to eight times greater than that of ceftriaxone. Against Gram-negative organisms, ceftaroline showed good activity against β-lactamase-negative strains but not against the members of the Bacteroides fragilis group. Ceftaroline showed potent activity against a broad spectrum of anaerobes encountered in respiratory, skin, and soft tissue infections.With the continuing emergence of novel patterns of resistance to commonly used antimicrobial agents, alternative therapies are needed to treat serious infections. Ceftaroline is a novel, parenteral, broad-spectrum cephalosporin that exhibits bactericidal activity against Gram-positive organisms, such as methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-intermediate S. aureus, and multidrug-resistant Streptococcus pneumoniae (MDRSP) strains, as well as common Gram-negative pathogens (8, 12, 14, 16, 18-22). Ceftaroline is currently in development for the treatment of complicated skin and skin structure infections and community-acquired pneumonia.Anaerobic bacteria are common pathogens in a variety of pleuropulmonary infections, including aspiration pneumonia, lung abscesses, and empyema (1, 3, 6, 15). However, many laboratories do not culture for anaerobes (9), diminishing awareness of the role of anaerobes in these infections. The main anaerobic pathogens isolated from these infections include Prevotella melaninogenica (∼25%), Prevotella intermedia (∼30%), Fusobacterium species (∼39%), Gram-positive cocci (∼30%), and Veillonella species (∼35%) (7). Cephalosporins such as cefoxitin have been used for the therapy of aspiration pneumonias. Although cefoxitin is active against most respiratory anaerobes, it has poor activity against the newer resistant strains of members of the family Enterobacteriaceae and MRSA. The activity of ceftaroline against Gram-positive anaerobes is similar to that of amoxicillin-clavulanate, and non-β-lactamase-producing Gram-negative strains generally have low ceftaroline MICs (present study), suggesting that ceftaroline might have an adequate spectrum of activity for therapy for some cases of aspiration pneumonia.To investigate the broader potential of ceftaroline, we compared its in vitro activity against 623 unique clinical isolates of anaerobic bacteria representing 5 Gram-negative bacterial genera and 17 Gram-positive bacterial genera to the activities of ceftriaxone, metronidazole, clindamycin, and amoxicillin-clavulanate.The reference agar dilution procedure described in CLSI document M11-A7 was used (5). The organisms were recovered from a variety of clinical specimens and were stored at −70°C in 20% skim milk. Identification was accomplished by standard phenotypic methods or by partial 16S rRNA gene sequencing for strains that could not be identified phenotypically (13, 17). Quality control strains Bacteroides fragilis ATCC 25285, Clostridium difficile ATCC 700057, and Staphylococcus aureus ATCC 29213 were included on each day of testing.The antimicrobial agents were obtained as follows: ceftaroline was from Forest Laboratories, Inc. (New York, NY); ceftriaxone, vancomycin, and metronidazole were from Sigma-Aldrich, Inc. (St. Louis, MO); and amoxicillin and clavulanate were from GlaxoSmithKline (Research Triangle Park, NC). The agar dilution plates were prepared on the day of testing.The strains were taken from the freezer and transferred twice to ensure purity and good growth. Cell paste from 48-h cultures was suspended in brucella broth to achieve the turbidity of a 0.5 McFarland standard, and the mixture was applied to plates with a Steers replicator to deliver approximately 105 CFU/spot. The plates were incubated for 44 h at 37°C in an anaerobic chamber. The MIC was the lowest concentration that completely inhibited growth or that resulted in a marked reduction in growth compared with that for the drug-free growth control (5).A summary showing the MIC range, MIC50, MIC90, and percent susceptibility is presented in Table Table1.1. The cumulative ceftaroline MIC distributions for all groups of strains are displayed in Table Table22.

TABLE 1.

Summary of ceftaroline and comparator agent MICs, by species or group
OrganismNo. of isolatesMIC (μg/ml)
% susceptible% resistant
Range50%90%
Gram-negative bacteria
    Bacteroides fragilis30
        Ceftaroline4->641664NAaNA
        Ceftriaxone (≤16, ≥64)b4->6432642743
        Clindamycin (≤2, ≥8)0.06->12811286337
        Metronidazole (≤8, ≥32)0.25-2121000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.5-640.52937
    Bacteroides thetaiotaomicron20
        Ceftaroline32->6464>64NANA
        Ceftriaxone (≤16, ≥64)64->64>64>640100
        Clindamycin (≤2, ≥8)0.06->12841284545
        Metronidazole (≤8, ≥32)0.5-1111000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.5-824950
    Bacteroides fragilis group spp.c26
        Ceftaroline2->6464>64NANA
        Ceftriaxone (≤16, ≥64)4->64>64>642358
        Clindamycin (≤2, ≥8)0.06->1284>1284250
        Metronidazole (≤8, ≥32)0.5-2121000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.125-3228774
    Prevotella bivia20
        Ceftaroline0.125->64264NANA
        Ceftriaxone (≤16, ≥64)0.125->642>647515
        Clindamycin (≤2, ≥8)0.03->128≤0.03>1288515
        Metronidazole (≤8, ≥32)≤0.03-4121000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-40.2541000
    Prevotella buccae20
        Ceftaroline0.125->640.564NANA
        Ceftriaxone (≤16, ≥64)0.125->640.25645030
        Clindamycin (≤2, ≥8)≤0.03->128≤0.03>1288020
        Metronidazole (≤8, ≥32)0.25-10.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.06-40.0611000
    Prevotella melaninogenica18
        Ceftaroline≤0.008-32232NANA
        Ceftriaxone (≤16, ≥64)0.03-32232780
        Clindamycin (≤2, ≥8)≤0.03->128≤0.03>1287228
        Metronidazole (≤8, ≥32)0.06-20.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-20.12521000
    Prevotella intermedia20
        Ceftaroline≤0.008-64116NANA
        Ceftriaxone (≤16, ≥64)0.03-641168010
        Clindamycin (≤2, ≥8)≤0.03->128≤0.03168515
        Metronidazole (≤8, ≥32)0.125-20.2511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-10.060.51000
    Prevotella spp.d20
        Ceftaroline≤0.008-32232NANA
        Ceftriaxone (≤16, ≥64)≤0.008-6418905
        Clindamycin (≤2, ≥8)≤0.03->128≤0.031287030
        Metronidazole (≤8, ≥32)0.06-80.521000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-20.12511000
    Porphyromonas asaccharolytica21
        Ceftaroline≤0.008-0.50.0150.03NANA
        Ceftriaxone (≤16, ≥64)≤0.008-10.060.061000
        Clindamycin (≤2, ≥8)≤0.03->128≤0.03>1288119
        Metronidazole (≤8, ≥32)≤0.03-0.250.060.1251000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-≤0.03≤0.03≤0.031000
    Porphyromonas somerae10
        Ceftaroline≤0.008-160.01516NANA
        Ceftriaxone (≤16, ≥64)≤0.008-640.015648020
        Clindamycin (≤2, ≥8)≤0.03->128≤0.03>1288020
        Metronidazole (≤8, ≥32)0.25-0.50.50.51000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.5≤0.030.1251000
    Fusobacterium nucleatum22
        Ceftaroline≤0.008-0.125≤0.0080.125NANA
        Ceftriaxone (≤16, ≥64)0.015-10.1250.51000
        Clindamycin (≤2, ≥8)≤0.03-0.50.060.061000
        Metronidazole (≤8, ≥32)≤0.03-0.25≤0.030.251000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.5≤0.030.061000
    Fusobacterium necrophorum22
        Ceftaroline0.015-0.060.030.06NANA
        Ceftriaxone (≤16, ≥64)≤0.008-0.1250.0150.031000
        Clindamycin (≤2, ≥8)≤0.03-0.25≤0.030.061000
        Metronidazole (≤8, ≥32)0.06-0.250.1250.251000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-10.1250.51000
    Fusobacterium mortiferum10
        Ceftaroline1-64832NANA
        Ceftriaxone (≤16, ≥64)16->64>64>641090
        Clindamycin (≤2, ≥8)≤0.03-0.250.0611000
        Metronidazole (≤8, ≥32)0.25-20.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.25-848800
    Fusobacterium varium10
        Ceftaroline0.015-0.50.250.5NANA
        Ceftriaxone (≤16, ≥64)0.15-8181000
        Clindamycin (≤2, ≥8)0.06-64249010
        Metronidazole (≤8, ≥32)0.25-0.50.250.51000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.125-2121000
    Veillonella spp.19
        Ceftaroline0.015-10.1250.5NANA
        Ceftriaxone (≤16, ≥64)0.03-8487916
        Clindamycin (≤2, ≥8)≤0.03->1280.1251287921
        Metronidazole (≤8, ≥32)1-8281000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-80.254950
Gram-positive bacteria
    Anaerococcus prevotii-Anaerococcus tetradiuse20
        Ceftaroline≤0.008-20.030.125NANA
        Ceftriaxone (≤16, ≥64)0.03-320.250.5950
        Clindamycin (≤2, ≥8)≤0.03->1280.51286040
        Metronidazole (≤8, ≥32)0.125-4121000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-8≤0.030.125950
    Finegoldia magna19
        Ceftaroline0.03-10.250.5NANA
        Ceftriaxone (≤16, ≥64)2-8481000
        Clindamycin (≤2, ≥8)0.06->1282>1285337
        Metronidazole (≤8, ≥32)0.06-10.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.250.1250.251000
    Parvimonas micra22
        Ceftaroline0.015-0.50.060.25NANA
        Ceftriaxone (≤16, ≥64)0.125-20.511000
        Clindamycin (≤2, ≥8)0.06-1280.25168614
        Metronidazole (≤8, ≥32)0.125-10.250.251000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-10.1250.51000
    Peptoniphilus asaccharolyticus21
        Ceftaroline≤0.008-0.250.060.25NANA
        Ceftriaxone (≤16, ≥64)0.03-10.1250.251000
        Clindamycin (≤2, ≥8)≤0.03->1280.125>1287624
        Metronidazole (≤8, ≥32)0.125-2111000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.06≤0.030.061000
    Peptostreptococcus anaerobius-Peptostreptococcus stomatisf23
        Ceftaroline0.125-80.54NANA
        Ceftriaxone (≤16, ≥64)0.5-16281000
        Clindamycin (≤2, ≥8)≤0.03-32≤0.030.25964
        Metronidazole (≤8, ≥32)0.125-10.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-320.1250.5919
    Anaerobic Gram-positive coccig22
        Ceftaroline≤0.008-80.061NANA
        Ceftriaxone (≤16, ≥64)0.03-640.2516915
        Clindamycin (≤2, ≥8)≤0.03->1280.125647327
        Metronidazole (≤8, ≥32)0.25->6414919
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-40.060.51000
    Actinomyces spp.h13
        Ceftaroline≤0.008-0.250.0150.25NANA
        Ceftriaxone (≤16, ≥64)≤0.008-0.50.1250.51000
        Clindamycin (≤2, ≥8)≤0.03->1280.061287723
        Metronidazole (≤8, ≥32)>32->32>32>320100
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.50.060.51000
    Propionibacterium acnes20
        Ceftaroline≤0.008-0.125≤0.0080.06NANA
        Ceftriaxone (≤16, ≥64)≤0.008-0.1250.0150.061000
        Clindamycin (≤2, ≥8)0.125->1280.1250.125955
        Metronidazole (≤8, ≥32)>32->32>32>320100
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.25≤0.030.061000
    Propionibacterium avidum11
        Ceftaroline0.015-0.250.250.25NANA
        Ceftriaxone (≤16, ≥64)0.03-0.50.250.51000
        Clindamycin (≤2, ≥8)0.125-0.50.250.251000
        Metronidazole (≤8, ≥32)>32->32>32>320100
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.250.250.251000
    Eggerthella lenta17
        Ceftaroline2-16816NANA
        Ceftriaxone (≤16, ≥64)16->64>64>64694
        Clindamycin (≤2, ≥8)0.06-80.52946
        Metronidazole (≤8, ≥32)0.5-10.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.5-1111000
    “Eubacterium” groupi25
        Ceftaroline0.015-0.250.1250.25NANA
        Ceftriaxone (≤16, ≥64)0.03-160.521000
        Clindamycin (≤2, ≥8)≤0.03->1280.062928
        Metronidazole (≤8, ≥32)0.125-40.511000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.50.1250.251000
    Lactobacillus casei-Lactobacillus rhamnosus groupj10
        Ceftaroline0.25-80.51NANA
        Ceftriaxone (≤16, ≥64)8->6432644030
        Clindamycin (≤2, ≥8)0.25-2121000
        Metronidazole (≤8, ≥32)>64->64>64>640100
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.25-20.511000
    Clostridium perfringens20
        Ceftaroline≤0.008-0.50.1250.25NANA
        Ceftriaxone (≤16, ≥64)≤0.008-40.521000
        Clindamycin (≤2, ≥8)≤0.03-20.2511000
        Metronidazole (≤8, ≥32)0.5-4241000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.1250.030.1251000
    Clostridium ramosum21
        Ceftaroline1-211NANA
        Ceftriaxone (≤16, ≥64)0.25-0.50.250.51000
        Clindamycin (≤2, ≥8)1->128482443
        Metronidazole (≤8, ≥32)0.5-2111000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-0.250.060.251000
    Clostridium innocuum21
        Ceftaroline0.5-412NANA
        Ceftriaxone (≤16, ≥64)8-32816950
        Clindamycin (≤2, ≥8)0.125->1280.5>1288614
        Metronidazole (≤8, ≥32)0.5-4141000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.125-10.50.51000
    Clostridium clostridioforme groupk20
        Ceftaroline0.25-212NANA
        Ceftriaxone (≤16, ≥64)2->644327510
        Clindamycin (≤2, ≥8)≤0.03-40.52950
        Metronidazole (≤8, ≥32)≤0.03-0.250.060.251000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)0.25-10.50.51000
    Clostridium spp., otherl24
        Ceftaroline0.015-160.516NANA
        Ceftriaxone (≤16, ≥64)0.015->642647521
        Clindamycin (≤2, ≥8)≤0.03->12821285438
        Metronidazole (≤8, ≥32)0.125-40.541000
        Amoxicillin-clavulanate (≤4/2, ≥16/8)≤0.03-20.12511000
Open in a separate windowaNA, not applicable.bValues in parentheses are the breakpoints for susceptibility, resistance (in μg/ml).cBacteroides caccae (n = 6), B. distasonis (n = 3), B. merdae (n = 1), B. ovatus (n = 5), B. uniformis (n = 4), and B. vulgatus (n = 7).dPrevotella bergensis (n = 2), P. corporis (n = 1), P. denticola (n = 5), P. disiens (n = 5), P. loescheii (n = 3), P. nanceiensis (n = 2), P. oris (n = 1), and P. tannerae (n = 1).eAnaerococcus prevotii (n = 12) and A. tetradius (n = 8).fPeptostreptococcus anaerobius (n = 17) and P. stomatis (n = 6).gAnaerococcus lactolyticus (n = 1), Anaerococcus murdochii (n = 1), Anaerococcus octavius (n = 1), Anaerococcus vaginalis (n = 5), Anaerococcus species, no PCR match (n = 3), Gemella morbillorum (n = 1), Gemella sanguinis (n = 1), Peptoniphilus harei (n = 7), and Peptoniphilus lacrimalis (n = 2).hActinomyces israelii (n = 1), A. meyeri (n = 2), A. neuii subsp. anitratus (n = 2), A. odontolyticus (n = 3), and A. turicensis (n = 5).iAtopobium parvulum (n = 1), Collinsella aerofaciens (n = 4), Eubacterium contortum (n = 1), Eubacterium cylindroides (n = 1), Eubacterium limosum (n = 8), Eubacterium saburreum (n = 2), Mogibacterium timidum (n = 3), Slackia exigua (n = 4), and Solobacterium moorei (n = 1).jLactobacillus casei (n = 3) and L. rhamnosus (n = 7).kClostridium aldenense (n = 4), C. bolteae (n = 5), C. citroniae (n = 3), C. hathewayi (n = 4), and C. clostridioforme (n = 4).lClostridium barati (n = 1), C. bifermentans (n = 1), C. butyricum (n = 2), C. cadaveris (n = 2), C. celerecrescens (n = 1), C. difficile (n = 4), C. glycolicum (n = 2), C. hylemonae (n = 2), C. paraputrificum (n = 2), C. sordellii (n = 1), C. sphenoides (n = 1), C. subterminale (n = 1), C. symbiosum (n = 2), and C. tertium (n = 2).

TABLE 2.

Ceftaroline MIC distributions for Gram-negative and Gram-positive anaerobes
Organism group and organismTotalCumulative % of isolates with the following ceftaroline MIC (μg/ml):
≤0.0080.0150.030.060.1250.250.51248163264>64
Gram-negative anaerobes
    Bacteroides fragilis307376373100
    Bacteroides fragilis group, othera46479203757100
    Prevotella speciesb983.14.1121827374350556374829196100
    Porphyromonas speciesc31137181848790100
    Fusobacterium nucleatum/Fusobacterium necrophorumd4425507789100
    Fusobacterium mortiferum101020708090100
    Fusobacterium varium10203080100
    Veillonella species19532848995100
        Total288
Gram-positive anaerobes
    All Gram-positive coccie12710203047618292969798100
    Propionibacterium and Actinomyces speciesf444357647782100
    Lactobacillus casei-Lactobacillus rhamnosus groupg10208090100
    Eggerthella lenta1761288100
    “Eubacterium” group, otherh258202892100
    Clostridium perfringens2015356090100
    Clostridium ramosum2190100
    Clostridium innocuum21296795100
    Clostridium clostridioforme groupi20153580100
    Clostridium species, otherj2448214654677583100
        Total329
Open in a separate windowaBacteroides thetaiotaomicron (n = 20), B. caccae (n = 6), B. distasonis (n = 3), B. merdae (n = 1), B. ovatus (n = 5), B. uniformis (n = 4), and B. vulgatus (n = 7).bPrevotella bivia (n = 20), P. buccae (n = 20), P. melaninogenica (n = 18), P. intermedia (n = 20), P. bergensis (n = 2), P. corporis (n = 1), P. denticola (n = 5), P. disiens (n = 5), P. loescheii (n = 3), P. nanceiensis (n = 2), P. oris (n = 1), and P. tannerae (n = 1).cPorphyromonas asaccharolytica (n = 21) and P. somerae (n = 10).dFusobacterium nucleatum (n = 22) and F. necrophorum (n = 22).eFinegoldia magna (n = 19), Parvimonas micra (n = 22), Peptostreptococcus anaerobius (n = 17), Peptostreptococcus stomatis (n = 6), Anaerococcus prevotii (n = 12), Anaerococcus tetradius (n = 8), Anaerococcus lactolyticus (n = 1), Anaerococcus murdochii (n = 1), Anaerococcus octavius (n = 1), Anaerococcus vaginalis (n = 5), Anaerococcus species, no PCR match (n = 3), Gemella morbillorum (n = 1), Gemella sanguinis (n = 1), Peptoniphilus asaccharolyticus (n = 21), Peptoniphilus harei (n = 7), and Peptoniphilus lacrimalis (n = 2).fPropionibacterium acnes (n = 21), Propionibacterium avidum (n = 11), Actinomyces israelii (n = 1), Actinomyces meyeri (n = 2), Actinomyces neuii subsp. anitratus (n = 2), Actinomyces odontolyticus (n = 3), and Actinomyces turicensis (n = 5).gLactobacillus casei (n = 3) and L. rhamnosus (n = 7).hAtopobium parvulum (n = 1), Collinsella aerofaciens (n = 4), Eubacterium contortum (n = 1), Eubacterium cylindroides (n = 1), Eubacterium limosum (n = 8), Eubacterium saburreum (n = 2), Mogibacterium timidum (n = 3), Slackia exigua (n = 4), and Solobacterium moorei (n = 1).iClostridium aldenense (n = 4), C. bolteae (n = 5), C. citroniae (n = 3), C. hathewayi (n = 4), and C. clostridioforme (n = 4).jClostridium barati (n = 1), C. bifermentans (n = 1), C. butyricum (n = 2), C. cadaveris (n = 2), C. celerecrescens (n = 1), C. difficile (n = 4), C. glycolicum (n = 2), C. hylemonae (n = 2), C. paraputrificum (n = 2), C. sordellii (n = 1), C. sphenoides (n = 1), C. subterminale (n = 1), C. symbiosum (n = 2), and C. tertium (n = 2).The ceftaroline MIC50s for B. fragilis and other B. fragilis group species were 16 and 64 μg/ml, respectively, and the MIC90s were >64 μg/ml for both for B. fragilis and other B. fragilis group species. Ceftaroline was effective against all other Gram-negative, non-β-lactamase-producing strains and had activity similar to that of ceftriaxone. For Prevotella species, the ceftaroline MICs varied according to β-lactamase production, with the MIC50 and the MIC90 being 1 and 32 μg/ml, respectively. Most Porphyromonas species were susceptible to ceftaroline at ≤0.5 μg/ml; four β-lactamase-positive strains of Porphyromonas somerae (previously Porphyromonas levii), however, had ceftaroline MICs of 8 to 16 μg/ml. Fusobacterium nucleatum and Fusobacterium necrophorum, including two β-lactamase-positive strains, had a ceftaroline MIC50 and a ceftaroline MIC90 of 0.015 and 0.125 μg/ml, respectively. The bile-resistant Fusobacterium varium strains were susceptible to ceftaroline, with the highest MIC observed being 0.5 μg/ml, whereas Fusobacterium mortiferum had high MICs of ceftaroline (MIC90, 32 μg/ml), ceftriaxone (MIC90, >64 μg/ml), and amoxicillin-clavulanate (MIC90, 8 μg/ml). All Veillonella species were inhibited by ≤1 μg/ml ceftaroline.Almost all of the Gram-negative species were susceptible to metronidazole; four strains of Veillonella species and one strain of Prevotella nanceiensis, however, showed elevated MICs of 4 to 8 μg/ml. Clindamycin resistance was present in 37% of B. fragilis strains, 43% of Bacteroides thetaiotaomicron strains, 45% of B. fragilis group species, 21% of Prevotella species, and 19% of Porphyromonas asaccharolytica strains. Resistance to amoxicillin-clavulanate at >8/4 μg/ml was present in one B. fragilis strain and one Bacteroides ovatus strain, both of which were also resistant to imipenem; however, 19% of the B. fragilis group species showed an intermediate-susceptible amoxicillin-clavulanate MIC.Ceftaroline exhibited excellent activity against Gram-positive strains. The MIC50 and MIC90 for 127 strains of Gram-positive cocci were 0.125 and 0.5 μg/ml, respectively; and the MIC50 and MIC90 for 44 strains of Propionibacterium acnes, Propionibacterium avidum, and Actinomyces species were 0.015 and 0.25 μg/ml, respectively. The MIC50 and MIC90 for 106 strains of Clostridium species were 0.5 and 2 μg/ml, respectively, with higher MICs of 8 to 16 μg/ml being noted for 4 strains of Clostridium difficile, 1 strain of Clostridium celerecrescens, and 1 strain of Clostridium tertium. The MIC50 and MIC90 for 10 strains of vancomycin-resistant lactobacilli were 0.5 and 1 μg/ml, respectively. All “Eubacterium” group Gram-positive rods except Eggerthella lenta were inhibited by ≤0.25 μg/ml; the MIC50 and MIC90 for Eggerthella lenta were 8 and 16 μg/ml, respectively. Ceftaroline was four- to eightfold more active than ceftriaxone against Gram-positive organisms, with the MICs being the most similar to those of amoxicillin-clavulanate.Clindamycin resistance was present in 37% of the Finegoldia magna strains and 40% of the strains in the Anaerococcus prevotii and Anaerococcus tetradius groups. All strains of Actinomyces, Propionibacterium, and Lactobacillus were resistant to metronidazole, as were one strain of anaerobic Gemella morbillorum and one strain of Gemella sanguinis. All except two Gram-positive strains were susceptible to amoxicillin-clavulanate; the exceptions were two strains of Peptostreptococcus anaerobius (MICs, 32 μg/ml).Ceftaroline has been demonstrated to have excellent activity against strains commonly encountered in skin and respiratory infections, including MRSA, group A Streptococcus, MDRSP, and non-extended-spectrum β-lactamase (ESBL)-producing members of the family Enterobacteriaceae (8, 12, 14, 16, 18-22). The present study is the first to focus on the activity of ceftaroline against anaerobes and expands the known spectrum of species against which ceftaroline shows activity. The findings reported here are consistent with those of a limited study by Sader et al. (21).Although ceftaroline has a low level of activity against most Bacteroides isolates, its use in combination with a β-lactamase inhibitor might overcome this resistance and increase the clinical potential of the use of ceftaroline against intra-abdominal infections and some skin and soft tissue infections. Many skin infections contain anaerobes that are predominantly Gram-positive anaerobic cocci and relatively few Bacteroides species (2, 10), suggesting that ceftaroline may have activity in these instances as well.Our study confirmed the increasing resistance to clindamycin currently being reported by many investigators. Of particular interest was the resistance demonstrated by 2 of 19 strains of P. asaccharolytica, a species previously thought to be very susceptible to clindamycin (11). Additionally, four strains of P. somerae were β-lactamase producers, which is of interest because most studies do not report MICs for Porphyromonas and, to date, β-lactamase-producing strains have been a rare finding. We also noted an increase in the number of B. fragilis group strains with amoxicillin-clavulanate MICs reaching the intermediate level, similar to the increase in the ampicillin-sulbactam MICs reported in the CLSI M11-A7 supplement, which includes an antibiogram for the B. fragilis group (4).Except for Bacteroides species and β-lactamase-producing Prevotella isolates, ceftaroline showed potent activity against a broad spectrum of anaerobic bacteria frequently recovered from a variety of clinical infections.  相似文献   

8.
The in vitro antibacterial activities of nemonoxacin (TG-873870), a novel nonfluorinated quinolone, against 770 clinical isolates were investigated. Nemonoxacin (tested as its malate salt, TG-875649) showed better in vitro activity than ciprofloxacin and levofloxacin against different species of staphylococci, streptococci, and enterococci, Neisseria gonorrhoeae, and Haemophilus influenzae. The in vitro activity of TG-875649 was also comparable to or better than that of moxifloxacin against these pathogens, which included ciprofloxacin-resistant, methicillin-resistant Staphylococcus aureus and levofloxacin-resistant Streptococcus pneumoniae.Antimicrobial resistance is a global public health threat (6, 7). In Taiwan, multidrug and fluoroquinolone resistances are common in both Gram-negative and Gram-positive pathogens from inpatients as well as outpatients (3, 4, 8-10). For example, as many as 80% of nosocomial Staphylococcus aureus strains in Taiwan are methicillin-resistant S. aureus (MRSA) strains, of which 80% are fluoroquinolone resistant (9). Based on a national surveillance program of >1,200 pneumococcal isolates from recent years, the levofloxacin resistance level was 10% among non-penicillin-susceptible strains (isolates with penicillin MIC > 2 μg/ml) (3).Development of new antibiotics is one of the means of combating multidrug-resistant bacteria (6). Nemonoxacin (TG-873870) is a novel nonfluorinated quinolone, and TG-875649 is a malate salt of nemonoxacin. The chemical structure of TG-875649 is shown in Fig. Fig.1.1. The present study examined the in vitro antibacterial activity of TG-875649 against a spectrum of Gram-positive and Gram-negative clinical isolates in Taiwan.Open in a separate windowFIG. 1.Chemical structure of TG-875649, a malate salt of nemonoxacin (TG-873870). Me, methyl.(This study was presented in part at the 47th Interscience Conference on Antimicrobial Agents and Chemotherapy, Chicago, IL, 17 to 20 September 2007 [5].)Nonduplicate bacterial strains were selected from a pool of clinical isolates previously studied under a national surveillance program in Taiwan (4, 9). Isolates were selected to include those with known resistance to different classes of antibiotics in addition to fluoroquinolones (levofloxacin and/or ciprofloxacin); thus, multidrug-resistant bacteria comprised a larger proportion than the wild type. A total of 770 isolates were tested, among which 688 (89.4%) were isolated in the year 2004, with the remainder isolated from 1998 to 2002. MICs were determined by the broth microdilution (BMD) method, using custom-made 96-well microtiter panels containing different antimicrobial agents at various concentrations (Trek Diagnostics, West Essex, England). The test procedure followed the instructions of the MIC panel manufacturer and guidelines of the Clinical and Laboratory Standards Institute (CLSI) (1, 2). For most species, 50 μl of the 0.5 McFarland standard organism suspension was transferred to 10 ml of cation-adjusted Mueller-Hinton broth (CAMHB) to obtain a final inoculum of 5 × 105 CFU/ml at 100 μl/well. For Proteus mirabilis, a 1 × 105-CFU/ml inoculum was used to avoid an inoculum effect. For fastidious organisms, 100 μl of the 0.5 McFarland standard suspension was transferred to 10 ml of MHB containing 3% lysed horse blood for Neisseria gonorrhoeae and streptococci and to 10 ml of Haemophilus Test Medium (HTM) broth for Haemophilus influenzae. All MIC plates were incubated in 35°C ambient air overnight except those for N. gonorrhoeae, which were incubated in 5% CO2. Appropriate quality control strains were included during each test run. The CAMHB and HTM were purchased from Trek Diagnostics. All other media were purchased from BBL (Becton Dickinson Microbiology System, Cockeysville, MD).Table Table11 lists the antibacterial activities (MIC range, MIC50, and MIC90) of TG-875649 against various species of Gram-positive and Gram-negative pathogens in comparison with those of 3 fluoroquinolones (FQs) (ciprofloxacin, levofloxacin, and moxifloxacin) and nonquinolone agents. TG-875649 was more active than ciprofloxacin and levofloxacin against all staphylococcal isolates, including 150 S. aureus, with 4- to >32-fold lower MIC90s. In addition, against ciprofloxacin-resistant (CIPr) MRSA, the MIC90 of TG-875649 (1 μg/ml) was 4-fold lower than that of moxifloxacin. Of the 47 CIPr MRSA isolates, only 2 (4.3%) and 3 (6.4%) isolates had levofloxacin and moxifloxacin MICs of ≤1 μg/ml, respectively (Table (Table2),2), while the majority (45 isolates; 95.7%) had TG-875649 MICs of ≤1 μg/ml (Table (Table2).2). Among the 50 enterococci tested, 17 (7 Enterococcus faecalis strains and 10 E. faecium strains) were vancomycin resistant. The MIC90s of TG-875649 were at least 2-fold lower than those of the 3 FQs for both E. faecalis and E. faecium.

TABLE 1.

Antibacterial activities of TG-875649, a malate salt of nemonoxacin (TG-873870), and reference compounds against Gram-positive and Gram-negative bacteria
Organism (n) and compoundMIC (μg/ml)
Range50%90%
Staphylococcus aureus
    Methicillin susceptible (59)
        TG-8756490.015-10.030.12
        Ciprofloxacin0.12->160.52
        Levofloxacin≤0.06-160.251
        Moxifloxacin≤0.015-40.060.12
        Ceftriaxone2-844
        Oxacillin≤0.25-20.51
        Linezolid0.5-222
        Tigecycline0.06-0.250.120.25
        Vancomycin0.5-211
    Methicillin resistant, ciprofloxacin
            susceptible (44)
        TG-875649≤0.008-0.060.030.06
        Ciprofloxacin≤0.06-10.50.5
        Levofloxacin≤0.06-0.250.250.25
        Moxifloxacin≤0.015-0.120.060.06
        Ceftriaxone8->8>8>8
        Oxacillin4->8>8>8
        Linezolid222
        Tigecycline0.06-0.50.250.25
        Vancomycin1-211
    Methicillin resistant, ciprofloxacin
            resistant (47)
        TG-8756490.06-411
        Ciprofloxacin2->16>16>16
        Levofloxacin0.5->161616
        Moxifloxacin0.12-844
        Ceftriaxone8->8>8>8
        Oxacillin>8>8>8
        Linezolid1-222
        Tigecycline0.12-0.50.250.5
        Vancomycin1-222
Coagulase negative staphylococci
    Methicillin resistant (68)a
        TG-8756490.03-80.120.5
        Ciprofloxacin0.12->160.5>16
        Levofloxacin0.12->160.58
        Moxifloxacin0.03->80.122
        Ceftriaxone4->8>8>8
        Oxacillin0.5->8>8>8
        Linezolid≤0.25-412
        Tigecycline0.06-0.50.250.5
        Vancomycin0.25-422
Enterococcus faecalis (31)b
    TG-8756490.12-40.254
    Ciprofloxacin1->162>16
    Levofloxacin1->162>16
    Moxifloxacin0.25->80.5>8
    Linezolid1-212
    Tigecycline0.06-0.250.250.25
    Vancomycin1->162>16
Enterococcus faecium (19)b
    TG-8756490.5-848
    Ciprofloxacin2->16>16>16
    Levofloxacin2->16>16>16
    Moxifloxacin1->8>8>8
    Linezolid1-222
    Tigecycline0.03-0.250.060.12
    Vancomycin0.5-168>16
Escherichia coli
    Ciprofloxacin resistant (43)
        TG-8756492->16>16>16
        Ciprofloxacin2->16>16>16
        Levofloxacin1->16>16>16
        Moxifloxacin1->8>8>8
        Ceftazidime≤1->1281664
        Ceftriaxone0.06->8>8>8
        Cefepime≤0.06->80.5>8
        Imipenem≤0.12-20.50.5
        Piperacillin2->128>128>128
        Tigecycline0.12-0.50.250.5
    Ciprofloxacin susceptible (37)
        TG-8756490.015-412
        Ciprofloxacin≤0.06-10.251
        Levofloxacin≤0.06-20.51
        Moxifloxacin≤0.015-20.51
        Ceftazidime≤1-32116
        Ceftriaxone≤0.03->80.06>8
        Cefepime≤0.06->8≤0.068
        Imipenem≤0.12-10.250.5
        Piperacillin2->128>128>128
        Tigecycline0.12-0.50.250.5
Klebsiella pneumoniae (30)
    TG-8756490.25->164>16
    Ciprofloxacin≤0.06->164>16
    Levofloxacin≤0.06->162>16
    Moxifloxacin0.12->84>8
    Ceftazidime≤1->12816>128
    Ceftriaxone≤0.03->8>8>8
    Cefepime≤0.06->82>8
    Imipenem0.25-320.54
    Piperacillin4->128>128>128
    Tigecycline0.25-20.51
Enterobacter cloacae (30)
    TG-8756490.5->164>16
    Ciprofloxacin0.12->162>16
    Levofloxacin0.12->16216
    Moxifloxacin0.25->82>8
    Ceftazidime≤1->128128>128
    Ceftriaxone0.12->8>8>8
    Cefepime≤0.06-84>8
    Imipenem0.25-814
    Piperacillin4->128>128>128
    Tigecycline0.25-40.52
Proteus mirabilis (30)
    TG-8756490.5->168>16
    Ciprofloxacin≤0.06->162>16
    Levofloxacin≤0.06->162>16
    Moxifloxacin0.25->88>8
    Ceftazidime≤1-8≤1≤1
    Ceftriaxone≤0.03->8≤0.03>8
    Cefepime≤0.06->8≤0.068
    Imipenem0.25->3228
    Piperacillin1->12832>128
    Tigecycline1-424
Citrobacter freundii (30)
    TG-8756490.12->160.54
    Ciprofloxacin≤0.06->16≤0.062
    Levofloxacin≤0.06->160.122
    Moxifloxacin0.12->80.58
    Ceftazidime≤1->128≤164
    Ceftriaxone≤0.03->80.25>8
    Cefepime≤0.06->8≤0.06>8
    Imipenem0.25-412
    Piperacillin2->12816>128
    Tigecycline0.25-10.250.5
Streptococcus pneumoniae
    Levofloxacin susceptible (71)
        TG-8756490.06-0.250.120.12
        Ciprofloxacin1-1624
        Levofloxacin0.5-212
        Moxifloxacin0.06-0.50.120.25
        Ceftriaxone≤0.06->812
        Linezolid0.5-211
        Tigecycline≤0.08-0.120.030.06
        Vancomycin≤0.12-10.50.5
    Levofloxacin-resistant (29)
        TG-8756490.5-812
        Ciprofloxacin8->16>16>16
        Levofloxacin8->1616>16
        Moxifloxacin2->828
        Ceftriaxone0.12-812
        Linezolid0.5-211
        Tigecycline0.015-0.120.060.06
        Vancomycin0.25-0.50.50.5
Streptococcus pyogenes (30)
    TG-8756490.06-0.120.120.12
    Ciprofloxacin0.25-414
    Levofloxacin0.25-212
    Moxifloxacin0.06-0.50.250.5
    Ceftriaxone≤0.06≤0.06≤0.06
    Linezolid0.5-111
    Tigecycline0.015-0.060.030.06
    Vancomycin0.25-0.50.50.5
Streptococcus agalactiae (30)
    TG-8756490.12-20.120.25
    Ciprofloxacin1->1612
    Levofloxacin0.5->1612
    Moxifloxacin0.12->80.250.5
    Ceftriaxone≤0.06->8≤0.060.12
    Linezolid1-211
    Tigecycline0.03-0.060.030.06
    Vancomycin0.5-10.51
Viridans group streptococci (30)
    TG-9756490.06-0.250.120.25
    Ciprofloxacin0.25-414
    Levofloxacin0.25-212
    Moxifloxacin0.06-9,250.120.25
    Ceftriaxone≤0.06-10.250.5
    Linezolid0.5-212
    Tigecycline0.03-10.060.12
    Vancomycin0.5-10.51
Acinetobacter baumannii (30)
    TG-8756490.06->164>16
    Ciprofloxacin≤0.06-164>16
    Levofloxacin≤0.06->16216
    Moxifloxacin0.06->81>8
    Ceftazidime8->12864>128
    Ceftriaxone8->8>8>8
    Cefepime4->8>8>8
    Imipenem0.5->322>32
    Piperacillin16->128>128>128
    Tigecycline0.12-212
Pseudomonas aeruginosa (30)
    TG-8756491->1616>16
    Ciprofloxacin0.12->164>16
    Levofloxacin0.25->168>16
    Moxifloxacin1->8>8>8
    Ceftazidime2->12832>128
    Ceftriaxone>8>8>8
    Cefepime1->8>8>8
    Imipenem1->328>32
    Piperacillin4->128128>128
    Tigecycline2->88>8
Haemophilus influenzae (30)
    TG-875649≤0.008-80.124
    Ciprofloxacin≥0.06-160.2516
    Levofloxacin≥0.06-80.258
    Moxifloxacin≥0.015->80.25>8
    Ceftazidime≤1≤1≤1
    Ceftriaxone≤0.03-0.06≤0.03≤0.03
    Cefepime≤0.06-0.50.250.5
    Imipenem≤0.12-814
    Tigecycline0.12-10.51
Open in a separate windowaIncluding 44 S. epidermidis coagulase-negative staphylococci and 24 non-S. epidermidis coagulase-negative staphylococci.bIncluding 17 (7 E. faecalis strains and 10 E. faecium strains) vancomycin-resistant enterococci.

TABLE 2.

MIC distribution of TG-875649 and comparator fluoroquinolone agents against ciprofloxacin-susceptible and -resistant, methicillin-resistant S. aureus (MRSA)
Open in a separate window
Open in a separate windowaMICs before vertical solid lines, between solid and dotted lines, and after dotted lines indicate susceptible, intermediate, and resistant breakpoints, respectively, based on CLSI interpretive criteria (2). The white fields denote range of dilutions tested for each agent. Values above the range denote MICs greater than the highest concentration tested. MICs equal to or lower than the lowest concentration tested are given as the lowest concentration.Against levofloxacin-susceptible Streptococcus pneumoniae isolates, the MIC90 of TG-875649 (0.12 μg/ml) was severalfold lower than those of ciprofloxacin (4 μg/ml) and levofloxacin (2 μg/ml) and 2-fold lower than that of moxifloxacin (0.25 μg/ml). Against the 29 levofloxacin-resistant S. pneumoniae isolates studied, TG-875649 also had the lowest MIC90 (2 μg/ml), lower than those of the 3 FQs (8 to >16 μg/ml). The activity of TG-875649 was comparable to that of moxifloxacin against viridans group streptococci (MIC90 of 0.25 μg/ml for both drugs) and better than that of moxifloxacin against group A and group B streptococci (S. pyogenes and S. agalactiae, respectively), with MIC90s of 0.12 to 0.25 μg/ml for TG-875649 and 0.5 μg/ml for moxifloxacin. Three group B streptococcal isolates were FQ resistant, with ciprofloxacin and levofloxacin MICs of >16 μg/ml and a moxifloxacin MIC of 2 to >8 μg/ml, while the MICs of TG-875649 were lower, at 1 to 2 μg/ml (data not shown).TG-875649 was less active against Gram-negative pathogens Enterobacteriaceae, Acinetobacter baumannii, and Pseudomonas aeruginosa. The MIC90s of TG-875649 and the 3 FQs against these species (except Citrobacter freundii and CIP-susceptible Escherichia coli) were nearly all equal to or greater than the highest concentrations tested. A look at the MIC50s revealed that the in vitro activities of TG-875649 were 2- to 4-fold less than the 3 FQs against most of these species. However, against Haemophilus influenzae, TG-875649 showed slightly better activity than the 3 FQs, with an MIC50 of 0.12 μg/ml (those of the 3 FQs were all 0.25 μg/ml), and its MIC90 was also ≥2-fold lower than those of the 3 FQs. It needs to be pointed out that the 30 H. influenzae isolates tested included 11 non-FQ-susceptible isolates. TG-875649 also showed slightly better activity against N. gonorrhoeae. Of the 10 N. gonorrhoeae isolates studied, only 2 were susceptible to ciprofloxacin. The ciprofloxacin, levofloxacin, and moxifloxacin MICs of the other 8 non-FQ-susceptible N. gonorrhoeae isolates ranged from 2 to 4, 1 to 4, and 0.5 to 2 μg/ml, respectively, but the MIC for TG-875649 was lower, at 0.25 to 1 μg/ml (data not shown).In conclusion, nemonoxacin (tested as its malate salt, TG-875649) demonstrated better antibacterial activities than ciprofloxacin and levofloxacin against different species of Gram-positive bacteria, including staphylococci, streptococci, and enterococci, and against H. influenzae and N. gonorrhoeae. The in vitro activity of TG-875649 was also comparable to or better than that of moxifloxacin against these pathogens, including ciprofloxacin-resistant MRSA and levofloxacin-resistant S. pneumoniae. Based on these data, further studies of nemonoxacin pharmacology and development of bacterial resistance to nemonoxacin are warranted.  相似文献   

9.
Resistance of Gram-positive pathogens to first-line antimicrobial agents has been increasing in many parts of the world. We compared the in vitro activities of torezolid with those of other antimicrobial agents, including linezolid, against clinical isolates of major aerobic and anaerobic bacteria. Torezolid had an MIC90 of ≤0.5 μg/ml for the Gram-positive bacterial isolates tested and was more potent than either linezolid or vancomycin.Antimicrobial resistance in Gram-positive cocci has become a major problem in recent years. Oxazolidinones, a new therapeutic class of synthetic drugs, are active against Gram-positive pathogens. Linezolid, the only marketed oxazolidinone, inhibits the initiation of bacterial protein translation by binding to the 23S rRNA peptidyl transferase region (15). The widely used drug linezolid is effective against most Gram-positive pathogens, including methicillin-resistant Staphylococcus aureus (MRSA), vancomycin-resistant Enterococcus spp., and penicillin-resistant Streptococcus pneumoniae (1, 2). However, several recent studies have reported the emergence of linezolid-resistant staphylococci and enterococci in Brazil, China, France, Germany, Italy, and Sweden. The dominant resistance mechanisms are mutations of the 23S rRNA gene and the recently described mobile chloramphenicol-florfenicol resistance (cfr) methyltransferase gene (9).The antibacterial activity of oxazolidinones depends on their affinity for the site of action on the ribosome. Therefore, by modifying their chemical structure, novel oxazolidinones with improved antimicrobial activity can be obtained. Accordingly, it is important to find more useful and less toxic oxazolidinones. Torezolid [TR-700, DA-7157; R-3-(4-(2-(2-methyltetrazol-5-yl)pyridine-5-yl)-3-fluorophenyl)-5-hydroxymethyl oxazolidin-2-on] is the active moiety of the prodrug torezolid phosphate (TR-701, DA-7218) (Fig. (Fig.1).1). In a recent study, torezolid was 4- to 8-fold more active than linezolid against Gram-positive bacteria collected from the United States (3). In another study, torezolid demonstrated an 8- to 16-fold increase in potency against all of the linezolid-resistant isolates tested, including MRSA, MRSA carrying the mobile cfr methyltransferase gene, and vancomycin-resistant enterococci (14). However, as far as we know, the activities of torezolid against anaerobic bacteria have not been reported.Open in a separate windowFIG. 1.Chemical structure of torezolid.Human plasma protein binding of torezolid was about 80% (data not shown), and the MIC was unaffected by the presence of 20% human plasma (4). Torezolid has a better pharmacokinetic profile than linezolid. After oral administration of torezolid at 200 mg once a day, the maximum concentration of the drug in serum, half-life, and area under the curve were 2.0 μg/ml, 11.2 h, and 25.4 μg·h/ml, respectively (13). In another study, torezolid phosphate was safe and effective with once-daily 200-mg dosing over 5 to 7 days of treatment for severe complicated skin and skin structure infections caused by Gram-positive bacteria (16). In this study, we compared the in vitro activities of torezolid with those of other antimicrobial agents, including linezolid, against clinical isolates of major aerobic and anaerobic Gram-positive and Gram-negative bacteria.(Part of this study was presented at the 44th Interscience Conference on Antimicrobial Agents and Chemotherapy, Washington, DC, 2004 [12]).Five hundred ten nonduplicate aerobic and anaerobic bacterial isolates were collected between 2002 and 2004 from patients at a South Korean tertiary-care hospital. The species were identified by conventional methods or by using either the ID 32 GN or the ATB 32A system (bioMérieux, Marcy-l''Etoile, France). Antimicrobial susceptibility was tested by the CLSI agar dilution method (5, 6, 7). The media used were Mueller-Hinton agar (Becton Dickinson, Sparks, MD) for testing of Staphylococcus spp., Enterococcus spp., and Moraxella catarrhalis; Mueller-Hinton agar supplemented with 5% sheep blood for Streptococcus spp.; Haemophilus test medium for Haemophilus influenzae; and brucella agar (Becton Dickinson) supplemented with 5 μg hemin, 1 μg vitamin K1 per ml, and 5% laked sheep blood for anaerobic bacteria.The antimicrobial agents used were torezolid and linezolid (Dong-A, Seoul, South Korea); erythromycin, tetracycline, oxacillin, penicillin G, and cefuroxime (Sigma Chemical, St. Louis, MO); piperacillin and tazobactam (Yuhan, Seoul, South Korea); azithromycin and sulbactam (Pfizer Korea, Seoul, South Korea); clindamycin (Korea Upjohn, Seoul, South Korea); levofloxacin (Daiichi, Tokyo, Japan); ampicillin, gentamicin, and chloramphenicol (Chong Kun Dang, Seoul, South Korea); cefotaxime (Han-Dok, Seoul, South Korea); cefoxitin and imipenem (Merck Sharp & Dohme, Rahway, NJ); cefotetan (Je Il, Seoul, South Korea); metronidazole (Choong Wae, Seoul, South Korea); trimethoprim and sulfamethoxazole (Dong Wha, Seoul, South Korea); cefaclor and vancomycin (Daewoong, Seoul, South Korea); and teicoplanin (Sanofi Aventis, Bridgewater, NJ).American Type Culture Collection strains of S. aureus (ATCC 29213), Enterococcus faecalis (ATCC 29212), S. pneumoniae (ATCC 49619), H. influenzae (ATCC 49247), Bacteroides fragilis (ATCC 25285), and Bacteroides thetaiotaomicron (ATCC 29741) were used as reference strains. The meningeal breakpoints of penicillin G and cefotaxime were used for S. pneumoniae.MRSA continues to be prevalent in South Korea, accounting for 64% of the S. aureus strains in one study (10). In this study, all of the isolates of staphylococci tested were inhibited by torezolid at ≤1 μg/ml and the MIC for 90% of the strains tested (MIC90) was 4- to 8-fold lower than that of linezolid (Table (Table1).1). The majority of the MRSA isolates was resistant to erythromycin, clindamycin, gentamicin, levofloxacin, and tetracycline.

TABLE 1.

Comparative antimicrobial activities of torezolid and other antimicrobial agents against aerobic and anaerobic bacteria
Organism (no. of isolates tested) and antimicrobial agentBreakpoint (μg/ml)f
MIC (μg/ml)
Susceptibility (%)f
SIRRange50%90%SIR
Methicillin-susceptible S. aureus (30)
    TorezolidNAgNANA0.5-10.50.5NANANA
    Linezolid≤4≥82-444100NA0
    Erythromycin≤0.51-4≥80.5->1280.5>12863730
    Clindamycin≤0.51-2≥4≤0.06-10.250.259730
    Cotrimoxazole≤2/38≥4/76≤0.06-320.25290NA10
    Gentamicin≤48≥160.06->1280.512870327
    Levofloxacin≤12≥40.5-80.519703
    Tetracycline≤48≥160.25-640.53283017
    Oxacillin≤2≥40.06-0.50.50.5100NA0
    Vancomycin≤24-8≥160.5-10.5110000
MRSA (30)
    TorezolidNANANA0.50.50.5NANANA
    Linezolid≤4≥82-424100NA0
    Erythromycin≤0.51-4≥80.5->128>128>1283393
    Clindamycin≤0.51-2≥40.25->128>128>12817083
    Cotrimoxazole≤2/38≥4/760.25->1280.5>12873NA27
    Gentamicin≤48≥160.25->12864>12813087
    Levofloxacin≤12≥40.5->12816>12817083
    Tetracycline≤48≥160.5-128646433067
    Oxacillin≤2≥432->128>128>1280NA100
    Vancomycin≤24-8≥160.5-21110000
Methicillin-susceptible, coagulase negative staphylococci (29)
    TorezolidNANANA0.25-0.50.50.5NANANA
    Linezolid≤4≥81-424100NA0
    Erythromycin≤0.51-4≥80.25->1280.512876024
    Clindamycin≤0.51-2≥40.12->1280.2519073
    Cotrimoxazole≤2/38≥4/76≤0.06-320.251690NA10
    Gentamicin≤48≥160.06-1280.126469724
    Levofloxacin≤12≥40.25-320.50.59703
    Tetracycline≤48≥160.5-1280.53276024
    Oxacillin≤0.25≥0.50.06-0.250.120.25100NA0
    Vancomycin≤24-8≥160.5-21110000
Methicillin-resistant, coagulase negative staphylococci (26)
    TorezolidNANANA0.12-0.50.50.5NANANA
    Linezolid≤4≥80.5-422100NA0
    Erythromycin≤0.51-4≥8≤0.06->1286412842058
    Clindamycin≤0.51-2≥40.12->1280.25>12862038
    Cotrimoxazole≤2/38≥4/76≤0.06-3223250NA50
    Gentamicin≤48≥160.06-1281664271558
    Levofloxacin≤12≥40.12-160.516731215
    Tetracycline≤48≥160.5->128412869427
    Oxacillin≤0.25≥0.50.5->1284640NA100
    Vancomycin≤24-8≥160.25-21210000
Vancomycin-susceptible Enterococcus faecalis (49)
    TorezolidNANANA0.12-0.50.250.5NANANA
    Linezolid≤24≥80.5-22210000
    Ampicillin≤8≥160.25-814100NA0
    Erythromycin≤0.51-4≥80.12->1284>12894249
    Levofloxacin≤24≥80.5-6426469031
    Tetracycline≤48≥160.5-128646420080
    Vancomycin≤48-16≥321-42210000
    Teicoplanin≤816≥32≤0.12-0.50.250.510000
Vancomycin-resistant E. faecalis (12)
    TorezolidNANANA0.25-0.50.250.5NANANA
    Linezolid≤24≥80.5-11110000
    Ampicillin≤8≥161-424100NA0
    Erythromycin≤0.51-4≥8>128>128>12800100
    Levofloxacin≤24≥816-128646400100
    Tetracycline≤48≥160.5-6432648092
    Vancomycin≤48-16≥32>128>128>12800100
    Teicoplanin≤816≥3232-128646400100
Vancomycin-susceptible Enterococcus faecium (30)
    TorezolidNANANA0.06-0.250.250.25NANANA
    Linezolid≤24≥80.5-22210000
    Ampicillin≤8≥161->128>128>1287NA93
    Erythromycin≤0.51-4≥80.25->128>128>1283790
    Levofloxacin≤24≥82-12864643790
    Tetracycline≤48≥160.12-320.519703
    Vancomycin≤48-16≥320.5-40.50.510000
    Teicoplanin≤816≥320.25-20.50.510000
Vancomycin-resistant E. faecium (29)
    TorezolidNANANA0.06-0.250.120.25NANANA
    Linezolid≤24≥80.5-11110000
    Ampicillin≤8≥1664->128>128>1280NA100
    Erythromycin≤0.51-4≥864->128128>12800100
    Levofloxacin≤24≥816-1286412800100
    Tetracycline≤48≥16≤0.06-1280.2512890010
    Vancomycin≤48-16≥3264->128128>12800100
    Teicoplanin≤816≥322-641664213148
S. pneumoniae (29)
    TorezolidNANANA0.12-0.50.250.25NANANA
    Linezolid≤20.5-211100NANA
    Penicillin G≤0.06≥0.120.015-21217NA83
    Cefotaximec≤0.51≥20.015-212315514
    Clindamycin≤0.250.5≥10.25->128>128>12828072
    Erythromycin≤0.250.5≥10.25->128>128>12814086
    Cotrimoxazole≤0.5/9.51/19-2/38≥4/760.5-1281664241066
    Levofloxacin≤24≥81-22210000
    Tetracycline≤24≥8≤0.12-32163210090
S. pyogenes (15)
    TorezolidNANANA0.06-0.250.120.25NANANA
    Linezolid≤21-212100NANA
    Penicillin G≤0.12≤0.008-0.0150.0150.015100NANA
    Cefotaxime≤0.5≤0.008-0.030.0150.03100NANA
    Clindamycin≤0.250.5≥10.12-0.250.120.2510000
    Erythromycin≤0.250.5≥10.12-0.250.120.2510000
    Levofloxacin≤24≥80.5-41480200
S. agalactiae (15)
    TorezolidNANANA0.12-0.50.250.5NANANA
    Linezolid≤21-222100NANA
    Penicillin G≤0.120.03-0.060.060.06100NANA
    Cefotaxime≤0.50.03-0.060.060.06100NANA
    Clindamycin≤0.250.5≥10.25->1280.25>12853047
    Erythromycin≤0.250.5≥10.25->1280.5>128134740
    Levofloxacin≤24≥81-21210000
M. catarrhalis (27)
    TorezolidNANANA0.5-211NANANA
    LinezolidNANANA2-844NANANA
    Penicillin GNANANA0.03-321632NANANA
    Cefaclor≤816≥320.25-32289604
    Clindamycin≤0.51-2≥41-42405941
    Erythromycin≤0.51-4≥80.12-0.50.250.510000
    Levofloxacin≤20.060.060.06100NANA
    Tetracycline≤24≥80.25-160.50.59604
H. influenzae (25)
    TorezolidNANANA2-1624NANANA
    LinezolidNANANA4-16816NANANA
    Ampicillin≤12≥40.5->128>128>12816876
    Ampicillin-sulbactam≤2/1≥4/20.5-84836NA64
    Cefaclor≤816≥322->1284>12860040
    Cefuroxime≤48≥160.25->1281>12880416
    Cefotaxime≤2≤0.008-0.50.030.5100NANA
    Azithromycin≤42-444100NANA
    Cotrimoxazole≤0.5/9.51/19-2/38≥4/76≤0.06-3243248052
    Levofloxacin≤20.015-0.50.030.06100NANA
    Tetracycline≤24≥80.25-320.5884412
Peptostreptococcus spp. (59)a
    TorezolidNANANA0.03-0.250.060.25NANANA
    LinezolidNANANA0.25-20.51NANANA
    Ampicillin≤0.51≥2≤0.06-160.1219028
    Ampicillin-sulbactam≤8/416/8≥32/16≤0.06-80.12110000
    Piperacillin≤3264≥128≤0.06-16≤0.06810000
    Piperacillin-tazobactam≤32/464/4≥128/4≤0.06-16≤0.06810000
    Cefoxitin≤1632≥64≤0.06-160.25410000
    Cefotetan≤1632≥64≤0.06-1280.5169227
    Imipenem≤48≥16≤0.06-1≤0.060.1210000
    Clindamycin≤24≥8≤0.06->1280.56480020
    Metronidazole≤816≥32≤0.06-41210000
    VancomycinNANANA≤0.12-10.250.5NANANA
Clostridium perfringens (15)
    TorezolidNANANA0.12-0.250.250.25NANANA
    LinezolidNANANA1-222NANANA
    Ampicillin≤0.51≥2≤0.06-0.5≤0.060.1210000
    Ampicillin-sulbactam≤8/416/8≥32/16≤0.06-0.5≤0.060.2510000
    Piperacillin≤3264≥128≤0.06-1≤0.060.2510000
    Piperacillin-tazobactam≤32/464/4≥128/4≤0.06≤0.06≤0.0610000
    Cefoxitin≤1632≥640.25-10.5110000
    Cefotetan≤1632≥64≤0.06-0.5≤0.060.1210000
    Imipenem≤48≥16≤0.06-0.12≤0.06≤0.0610000
    Clindamycin≤24≥8≤0.06-21210000
    Metronidazole≤816≥321-44410000
    VancomycinNANANA0.5-20.50.5NANANA
Other Clostridium spp. (15)b
    TorezolidNANANA≤0.06-0.250.250.25NANANA
    LinezolidNANANA0.5-424NANANA
    Ampicillin≤0.51≥2≤0.06-10.25187130
    Ampicillin-sulbactam≤8/416/8≥32/16≤0.06-20.25110000
    Piperacillin≤3264≥128≤0.06-161810000
    Piperacillin-tazobactam≤32/464/4≥128/4≤0.06-161810000
    Cefoxitin≤1632≥640.25-12886460040
    Cefotetan≤1632≥64≤0.06->1282>12853740
    Imipenem≤48≥16≤0.06-41410000
    Clindamycin≤24≥8≤0.06->1281>128531333
    Metronidazole≤816≥320.12-16489370
    VancomycinNANANA0.25-848NANANA
Other anaerobic Gram-positive bacilli (13)c
    TorezolidNANANA0.06-0.50.060.5NANANA
    LinezolidNANANA≤0.06-40.52NANANA
    Ampicillin≤0.51≥2≤0.06-2≤0.0618588
    Ampicillin-sulbactam≤8/416/8≥32/16≤0.06-20.12110000
    Piperacillin≤3264≥128≤0.06-80.5810000
    Piperacillin-tazobactam≤32/464/4≥128/4≤0.06-8≤0.06810000
    Cefoxitin≤1632≥64≤0.06->1281>12810000
    Cefotetan≤1632≥640.12->1284>12862831
    Imipenem≤48≥16≤0.06-20.12210000
    Clindamycin≤24≥8≤0.06-4≤0.0629280
    Metronidazole≤816≥320.25->128>128>12838854
    VancomycinNANANA0.25->320.5>32NANANA
B. fragilis (30)
    TorezolidNANANA1-422NANANA
    LinezolidNANANA2-444NANANA
    Ampicillin≤0.51≥216->12832>12800100
    Ampicillin-sulbactam≤8/416/8≥32/161-3221683710
    Piperacillin≤3264≥1284->25632256531730
    Piperacillin-tazobactam≤32/464/4≥128/40.12-80.25110000
    Cefoxitin≤1632≥644-648328777
    Cefotetan≤1632≥644-12883283710
    Imipenem≤48≥16≤0.06-40.25110000
    Clindamycin≤24≥8≤0.06->128128>12843057
    Metronidazole≤816≥320.5-84410000
B. thetaiotaomicron (15)
    TorezolidNANANA1-222NANANA
    LinezolidNANANA444NANANA
    Ampicillin≤0.51≥216->12832>12800100
    Ampicillin-sulbactam≤8/416/8≥32/161-32132731313
    Piperacillin≤3264≥12816->25632>25673027
    Piperacillin-tazobactam≤32/464/4≥128/42-164810000
    Cefoxitin≤1632≥6416-32163273270
    Cefotetan≤1632≥6432->128128>12801387
    Imipenem≤48≥160.12-20.25210000
    Clindamycin≤24≥82->1288>12874053
    Metronidazole≤816≥322-44410000
Other Bacteroides spp. (14)d
    TorezolidNANANA1-412NANANA
    LinezolidNANANA1-424NANANA
    Ampicillin≤0.51≥22->128>128>12800100
    Ampicillin-sulbactam≤8/416/8≥32/161-32832572914
    Piperacillin≤3264≥1282->25664>256431443
    Piperacillin-tazobactam≤32/464/4≥128/42-164810000
    Cefoxitin≤1632≥644-64163279.147
    Cefotetan≤1632≥644->12864>128291457
    Imipenem≤48≥16≤0.06-20.5110000
    Clindamycin≤24≥84->128>128>1280793
    Metronidazole≤816≥32≤0.25-44410000
Other anaerobic Gram-negative rods (27)e
    TorezolidNANANA0.03-40.252NANANA
    LinezolidNANANA≤0.12-814NANANA
    Ampicillin≤0.51≥2≤0.03-128164223344
    Ampicillin-sulbactam≤8/416/8≥32/16≤0.03-41410000
    Piperacillin≤3264≥128≤0.06-1284329344
    Piperacillin-tazobactam≤32/464/4≥128/4≤0.06-8≤0.06410000
    Cefoxitin≤1632≥64≤0.06-81410000
    Cefotetan≤1632≥64≤0.06-322169370
    Imipenem≤48≥16≤0.06-1≤0.06110000
    Clindamycin≤24≥8≤0.06->128≤0.066478715
    Metronidazole≤816≥32≤0.06-40.5410000
    Chloramphenicol≤816≥320.5-82410000
Open in a separate windowaFinegoldia magna (19 strains), Peptoniphilus asaccharolyticus (15 strains), Peptostreptococcus anaerobius (12 strains), Peptostreptococcus micros (7 strains), and Anaerococcus prevotii (6 strains).bClostridium clostridiiforme (3 strains), C. sordellii (1 strain), C. innocuum (5 strains), C. tertium (2 strains), C. ramosum (2 strains), C. sporogenes (1 strain), and C. bifermentans (1 strain).cBifidobacterium adolescentis (2 strains), Propionibacterium acnes (4 strains), Eubacterium lentum (3 strains), Lactobacillus acidophilus (2 strains), and Actinomyces sp. (2 strains).dBacteroides distasonis (5 strains), B. vulgatus (7 strains), and B. ovatus (2 strains).ePrevotella bivia (6 strains), P. buccae (3 strains), P. intermedia (4 strains), P. oralis (2 strains), Fusobacterium mortiferum (3 strains), F. necrophorum (2 strains), F. varium (6 strains), and Fusobacterium sp. (1 strain).fS, susceptible; I, intermediate; R, resistant.gNA, not applicable.Vancomycin-resistant Enterococcus faecium has become prevalent in the United States (18). The vancomycin resistance rate of E. faecium has been 20% or higher in South Korean hospitals since 2003 (10). The MIC ranges of torezolid were 0.06 to 0.25 μg/ml for all of the enterococci, including vancomycin-resistant ones, while those of linezolid were 0.5 to 2 μg/ml (Table (Table1),1), which are similar to prior reports (8, 17). All of the isolates were susceptible to linezolid.Penicillin-nonsusceptible S. pneumoniae strains were very prevalent (69%) in South Korean hospitals in 2007, when the meningeal breakpoint was applied. In this study, most of the pneumococcal isolates tested were nonsusceptible to penicillin G or cefotaxime, but the MIC range of torezolid was 0.12 to 0.5 μg/ml and the MIC90 was 4-fold lower than that of linezolid (Table (Table1).1). All of the isolates of Streptococcus pyogenes and Streptococcus agalactiae were inhibited by torezolid at ≤0.5 μg/ml.β-Lactamase-producing M. catarrhalis and H. influenzae were prevalent in South Korea (11). The MIC ranges of torezolid for M. catarrhalis and H. influenzae were 0.5 to 2 and 2 to 16 μg/ml, respectively. The MIC90s for both of these organisms were 4-fold lower than those of linezolid.Intraabdominal and soft-tissue infections are often due to aerobic and anaerobic bacteria. Torezolid had excellent activity against Gram-positive anaerobes (Table (Table1).1). All of the peptostreptococci and anaerobic Gram-positive bacilli were inhibited by torezolid at ≤0.5 μg/ml, and the MIC90s for these organisms were 4- to 16-fold lower than those of linezolid. The MIC90 of torezolid, 2 μg/ml, for anaerobic Gram-negative bacilli, was slightly lower than that of linezolid, 4 μg/ml (Table (Table11).In conclusion, torezolid is a new antimicrobial agent with high in vitro activity against common aerobic and anaerobic Gram-positive bacteria, including multidrug-resistant isolates. Further studies are warranted to determine the clinical utility of torezolid as a therapeutic agent.  相似文献   

10.
We determined the antimicrobial susceptibilities of 255 clinical isolates of anaerobic bacteria collected in 2007 and 2008 at a tertiary-care hospital in South Korea. Piperacillin-tazobactam, cefoxitin, imipenem, and meropenem were highly active β-lactam agents against most of the isolates tested. The rates of resistance of Bacteroides fragilis group organisms and anaerobic Gram-positive cocci to moxifloxacin were 11 to 18% and 0 to 27%, respectively.Anaerobic bacterial resistance trends may vary greatly, depending on regions or institutions (1). The Clinical and Laboratory Standards Institute (CLSI) does not recommend routine susceptibility testing of all clinical isolates of anaerobic bacteria, except for the management of patients with serious infections (4). A recent survey indicated that only a few laboratories in the United States performed antimicrobial susceptibility testing of anaerobic bacteria due to the complex techniques and predictable susceptibilities involved (5). However, regional susceptibility patterns are pivotal in the empirical treatment of infected patients because these patterns are related to clinical outcomes (13). Therefore, periodic monitoring of the regional or institutional resistance trends of clinically important anaerobe isolates is recommended (4). Our investigation of resistance trends of Bacteroides fragilis group organisms from South Korea has been taking place since 1989 (9, 15). However, few studies have focused on the susceptibilities of other anaerobes. Therefore, the aim of this study was to determine the recent antimicrobial resistance patterns of frequently isolated anaerobes at a tertiary-care hospital in South Korea.Anaerobes were isolated from blood, normally sterile body fluid, and abscess specimens, but Clostridium difficile was isolated from stool specimens of suspected C. difficile-associated disease patients at Severance Hospital in 2007 and 2008. The isolates were identified by either conventional methods (19) or the ATB 32A system (bioMérieux, Marcy l''Etoile, France). A total of 255 nonduplicated isolates were used in this study, including 63 of B. fragilis, 57 of other B. fragilis group species, 28 of Prevotella spp., 9 of other Gram-negative bacilli, 15 of Anaerococcus prevotii, 15 of Peptoniphilus asaccharolyticus, 15 of Finegoldia magna, 13 of Peptostreptococcus spp., 15 of C. perfringens, 12 of C. difficile, and 13 of other Gram-positive bacilli.Antimicrobial susceptibility testing was performed using the CLSI agar dilution method (4). The medium used was Brucella agar (Becton Dickinson, Cockeysville, MD) supplemented with 5 μg hemin and 1 μg vitamin K1 per ml and 5% laked sheep blood. The antimicrobial powders used were piperacillin and tazobactam (Yuhan, Seoul, South Korea), cefoxitin (Merck Sharp & Dohme, West Point, PA), cefotetan (Daiichi Pharmaceutical, Tokyo, Japan), clindamycin (Korea Upjohn, Seoul, South Korea), imipenem, metronidazole (Choong Wae, Seoul, South Korea), chloramphenicol (Chong Kun Dang, Seoul, South Korea), meropenem (Sumitomo, Tokyo, Japan), moxifloxacin (Bayer Korea, Seoul, South Korea), and vancomycin (Eli Lilly & Co., Indianapolis, IN). For the combination of piperacillin and tazobactam, a constant tazobactam concentration of 4 μg/ml was added.An inoculum of 105 CFU was applied with a Steers replicator (Craft Machine Inc., Woodline, PA), and the plates were incubated in an anaerobic chamber (Forma Scientific, Marietta, OH) for 48 h at 37°C. The MIC of each antimicrobial agent was defined as the concentration at which there was a marked reduction in growth, such as from confluent colonies to a haze, <10 tiny colonies, or 1 to 3 normal-sized colonies. B. fragilis ATCC 25285 and B. thetaiotaomicron ATCC 29741 were used as controls.β-Lactamase production by anaerobic Gram-negative bacilli, with the exception of B. fragilis group organisms, was determined by applying test organisms to the Cefinase disks and recording the results after 30 min (Becton Dickinson, Cockeysville, MD).Table Table11 shows the MICs of antimicrobial agents and the resistance rates of the anaerobes tested. Among the 255 isolates, B. fragilis group organisms were the most prevalent (47%). These organisms are more virulent and more resistant to antimicrobial agents than most other anaerobes (3). In this study, piperacillin-tazobactam, cefoxitin, imipenem, and meropenem were highly active against B. fragilis group organisms, with resistance rates of less than 7%. The rates of resistance to imipenem and piperacillin-tazobactam were 4% and 7%, respectively, for other B. fragilis group organisms. However, much higher resistance rates were observed for piperacillin (27 to 51%), cefotetan (14 to 68%), and clindamycin (33 to 86%). These values were similar to those observed in 1997 to 2004 in the same hospital: piperacillin, 33 to 49%; cefotetan, 14 to 60%; clindamycin, 51 to 76% (15). A higher prevalence of resistance, in particular to clindamycin, was observed than in the United States, i.e.,19 to 35% (17). CLSI added a recommendation to test susceptibility to moxifloxacin in 2004 and 2007. In this study, the moxifloxacin resistance rates were 11% for B. fragilis and 18% for other B. fragilis group organisms. These rates were slightly higher than the 7 to 9% reported in Taiwan (11) but lower than those in Greece (14) and the United States (16 to 75% and 26 to 55%, respectively) (17).

TABLE 1.

Antimicrobial activities against 255 anaerobic bacteria isolated in 2007 to 2008
Organism (no. of isolates) and antimicrobial agentBreakpoint (μg/ml)a
MIC (μg/ml)
Susceptibility (%)a
SIRRange50%90%SIR
Bacteroides fragilis (63)
    Piperacillin≤3264≥1284->2568>25667627
    Piperacillin-tazobactam≤3264≥1280.25-128149722
    Cefoxitin≤1632≥648-128163279165
    Cefotetan≤1632≥644->128864711414
    Imipenem≤48≥160.06-320.12519802
    Meropenem≤48≥160.06-1280.12549253
    Clindamycin≤24≥8≤0.06->1280.5>12867033
    Moxifloxacin≤24≥80.25->1280.5884511
    Chloramphenicol≤816≥322-16449820
    Metronidazole≤816≥320.5-82210000
B. fragilis group, other species (57)b
    Piperacillin≤3264≥1288->256128>25642751
    Piperacillin-tazobactam≤3264≥1281->1288648947
    Cefoxitin≤1632≥644->128323225687
    Cefotetan≤1632≥644->128>128>12889586
    Imipenem≤48≥160.13-320.549524
    Meropenem≤48≥160.13-80.2529820
    Clindamycin≤24≥80.06->128>128>128161668
    Moxifloxacin≤24≥80.13->128216721018
    Chloramphenicol≤816≥324-16489820
    Metronidazole≤816≥320.5-42210000
Prevotella intermedia (10)
    Piperacillin≤3264≥1282-1681610000
    Piperacillin-tazobactam≤3264≥128≤0.03≤0.03≤0.0310000
    Cefoxitin≤1632≥640.5-42410000
    Cefotetan≤1632≥640.13-1621610000
    Imipenem≤48≥160.02-0.060.030.0610000
    Meropenem≤48≥160.03-0.060.060.0610000
    Clindamycin≤24≥8≤0.06-2≤0.06≤0.0610000
    Moxifloxacin≤24≥80.50.50.510000
    Chloramphenicol≤816≥320.5-10.5110000
    Metronidazole≤816≥320.5-20.5210000
Prevotella spp. (18)c
    Piperacillin≤3264≥1280.5-25616128781111
    Piperacillin-tazobactam≤3264≥128≤0.03-16≤0.03410000
    Cefoxitin≤1632≥640.5-3213289110
    Cefotetan≤1632≥640.5-64464721117
    Imipenem≤48≥160.03-10.060.510000
    Meropenem≤48≥160.03-10.1250.510000
    Clindamycin≤24≥8≤0.06-128≤0.0612850050
    Moxifloxacin≤24≥80.5-1628563311
    Chloramphenicol≤816≥320.5-84810000
    Metronidazole≤816≥320.5-84810000
Other Gram-negative bacilli (9)d
    Piperacillin≤3264≥1280.06-32NAgNANANANA
    Piperacillin-tazobactam≤3264≥128≤0.03-4NANANANANA
    Cefoxitin≤1632≥64≤0.06-8NANANANANA
    Cefotetan≤1632≥64≤0.06-8NANANANANA
    Imipenem≤48≥160.02-4NANANANANA
    Meropenem≤48≥16≤0.008-4NANANANANA
    Clindamycin≤24≥8≤0.06-128NANANANANA
    Moxifloxacin≤24≥80.25-128NANANANANA
    Chloramphenicol≤816≥320.13-4NANANANANA
    Metronidazole≤816≥320.13-1NANANANANA
Peptostreptococcus spp. (13)e
    Piperacillin≤3264≥1280.06-160.251610000
    Piperacillin-tazobactam≤3264≥128≤0.03-160.251610000
    Cefoxitin≤1632≥640.25-1611610000
    Cefotetan≤1632≥64≤0.06-12846462831
    Imipenem≤48≥16≤0.008-40.125210000
    Meropenem≤48≥160.01-40.25410000
    Clindamycin≤24≥8≤0.06-1280.1256477023
    Moxifloxacin≤24≥8≤0.06-80.1250.259208
    Chloramphenicol≤816≥321-22210000
    Metronidazole≤816≥320.25-10.5110000
Anaerococcus prevotii (15)
    Piperacillin≤3264≥128≤0.06-0.50.1250.2510000
    Piperacillin-tazobactam≤3264≥128≤0.03-10.1250.12510000
    Cefoxitin≤1632≥64≤0.06-40.5110000
    Cefotetan≤1632≥64≤0.06-41210000
    Imipenem≤48≥16≤0.008-0.250.060.2510000
    Meropenem≤48≥16≤0.008-0.250.060.12510000
    Clindamycin≤24≥8≤0.06-128212860040
    Moxifloxacin≤24≥8≤0.06-80.25887013
    Chloramphenicol≤816≥321-16489370
    Metronidazole≤816≥320.25-11110000
Peptoniphilus asaccharolyticus (15)
    Piperacillin≤3264≥128≤0.06-0.25≤0.06≤0.0610000
    Piperacillin-tazobactam≤3264≥128≤0.03-0.25≤0.030.0610000
    Cefoxitin≤1632≥64≤0.06-1≤0.060.510000
    Cefotetan≤1632≥640.13-20.25110000
    Imipenem≤48≥16≤0.008-0.13≤0.0080.0310000
    Meropenem≤48≥16≤0.008-0.06≤0.0080.0310000
    Clindamycin≤24≥8≤0.06-320.1253267033
    Moxifloxacin≤24≥80.13-20.25210000
    Chloramphenicol≤816≥321-42410000
    Metronidazole≤816≥320.5-21110000
Finegoldia magna (15)
    Piperacillin≤3264≥128≤0.06-0.25≤0.060.12510000
    Piperacillin-tazobactam≤3264≥128≤0.03-0.250.060.12510000
    Cefoxitin≤1632≥64≤0.06-10.5110000
    Cefotetan≤1632≥640.12-41210000
    Imipenem≤48≥16≤0.008-0.130.060.12510000
    Meropenem≤48≥160.03-0.130.060.12510000
    Clindamycin≤24≥8≤0.06-1280.2564731313
    Moxifloxacin≤24≥80.13-320.58601327
    Chloramphenicol≤816≥322-44410000
    Metronidazole≤816≥320.5-10.5110000
Clostridium perfringens (15)
    Piperacillin≤3264≥128≤0.06-0.50.250.510000
    Piperacillin-tazobactam≤3264≥128≤0.03-10.250.510000
    Cefoxitin≤1632≥640.5-21210000
    Cefotetan≤1632≥64≤0.06-10.25110000
    Imipenem≤48≥160.03-0.250.1250.12510000
    Meropenem≤48≥16≤0.008-0.030.0150.01510000
    Clindamycin≤24≥8≤0.06-1282480137
    Moxifloxacin≤24≥80.25-160.50.59307
    Chloramphenicol≤816≥322-84410000
    Metronidazole≤816≥320.02-0.060.030.0610000
    VancomycinNANANA0.25-10.50.5NANANA
Clostridium difficile (12)
    Piperacillin≤3264≥1282-84810000
    Piperacillin-tazobactam≤3264≥1281-164810000
    Cefoxitin≤1632≥6464->12864>12800100
    Cefotetan≤1632≥648-128812883017
    Imipenem≤48≥160.25-164858338
    Meropenem≤48≥160.25-22210000
    Clindamycin≤24≥82-128641288883
    Moxifloxacin≤24≥81->128163225075
    Chloramphenicol≤816≥321-1641683170
    Metronidazole≤816≥320.5-21110000
    VancomycinNANANA0.25-20.52NANANA
Other Gram-positive bacilli (13)f
    Piperacillin≤3264≥128≤0.06-64189280
    Piperacillin-tazobactam≤3264≥128≤0.03-640.589280
    Cefoxitin≤1632≥640.13-322169280
    Cefotetan≤1632≥640.13->12846485015
    Imipenem≤48≥16≤0.008-20.060.510000
    Meropenem≤48≥16≤0.008-160.12529208
    Clindamycin≤24≥8≤0.06->1280.06>12885015
    Moxifloxacin≤24≥8≤0.06-4129280
    Chloramphenicol≤816≥321-16129280
    Metronidazole≤816≥320.5->12816>12846846
Open in a separate windowaS, susceptible; I, intermediate; R, resistant.bBacteroides thetaiotaomicron (n = 25), B. caccae (n = 3), B. distasonis (n = 9), B. ovatus (n = 8), and B. vulgatus (n = 12).cPrevotella bivia (n = 10), P. buccae (n = 5), and P. oralis (n = 3).dPorphyromonas asaccharolytica (n = 2), Fusobacterium varium (n = 3), F. necrogenes (n = 2), F. nucleatum (n = 1), and F. mortiferum (n = 1).ePeptostreptococcus anaerobius (n = 9) and P. micros (n = 4).fAcitnomyces odontolyticus (n = 3), A. israelii (n = 2), A. meyeri (n = 1), A. naeslundii (n = 1), Bifidobacterium adolescentis (n = 3), Bifidobacterium sp. (n = 1), Eubacterium lentum (n = 1), and Eubacterium sp. (n = 1).gNA, not available/not applicable.Overall, Prevotella, Porphyromonas, and Fusobacterium isolates are more susceptible than B. fragilis group organisms (7). Among these organisms, β-lactamase producers were resistant to penicillin and ampicillin (3, 7). A recent study showed that 94% of the Prevotella isolates tested were β-lactamase producers, which correlated well with susceptibility to penicillin (11). In the present study, β-lactamase production was detected in 26 Prevotella isolates (94%) and 1 Fusobacterium isolate (14%). While 50% of the non-P. intermedia Prevotella isolates were resistant to clindamycin, all of the P. intermedia isolates were susceptible to clindamycin. Other studies indicated that 17% and 36% of the P. intermedia isolates were resistant to clindamycin (8, 16).Anaerobic Gram-positive cocci account for approximately one-quarter of all isolates from anaerobic infections. They may cause various infections, including skin infections, necrotizing pneumonia, and bacteremia (18). Several species previously placed in the genus Peptostreptococcus have been reclassified into new genera, including Anaerococcus, Finegoldia, Micrococcus, and Peptoniphilus (7). These organisms exhibited various rates of resistance to penicillin, clindamycin, and metronidazole (7). A European surveillance study showed that the majority of the isolates found to be resistant to clindamycin and penicillin were identified as F. magna (2). In our study, the rates of resistance of Gram-positive cocci to clindamycin and moxifloxacin varied according to species. The highest clindamycin resistance observed was 40% of A. prevotii isolates, followed by 33% of P. asaccharolyticus isolates. These rates were much higher than those reported in Europe (4%) and the United States (8%) (1, 2) but similar to the 25.9% observed in 1994 in South Korea (10). The rates of resistance to moxifloxacin varied from 27% among F. magna isolates to 0% among P. asaccharolyticus isolates. The difference in resistance rates among anaerobic Gram-positive cocci may be of importance. The resistance patterns of these organisms could help in the selection of appropriate antimicrobial treatment options, although susceptibility testing of individual patient isolates is not performed.C. perfringens is generally very susceptible to most antibiotics (7). The present study showed that all of the antimicrobial agents tested had high activity against this organism. C. difficile has highly variable resistance to β-lactams, including penicillin, cephalosporins, imipenem, clindamycin, and moxifloxacin (6, 7). In our study, the rates of resistance to cefoxitin, clindamycin, and moxifloxacin were 100%, 85%, and 77%, respectively. The C. difficile NAP1/027 epidemic isolates were known to be resistant to moxifloxacin (12). A high rate of resistance to moxifloxacin was observed in this study, although none of the isolates were NAP1/027 strains. Other Gram-positive bacilli, such as Actinomyces, Bifidobacterium, and Eubacterium species, are generally susceptible to β-lactams, including penicillin. Metronidazole-resistant isolates were common among these organisms (3). In our study, 46% of these organisms were resistant to metronidazole.In conclusion, piperacillin-tazobactam, cefoxitin, imipenem, meropenem, metronidazole, and chloramphenicol remain active against most anaerobic isolates. The rates of resistance of Gram-positive cocci to clindamycin and moxifloxacin are variable according to species. The rates of resistance to moxifloxacin are as follows: C. difficile, 75%; anaerobic Gram-positive cocci, 0 to 27%; B. fragilis group organisms, 11 to 18%. Continuous monitoring is necessary to detect pattern changes at regional centers.  相似文献   

11.
Streptococcus agalactiae isolates (n = 189) from patients with invasive infections were analyzed for capsular type by PCR, for antimicrobial susceptibility, and for the presence of resistance genes. In contrast to the predominance of capsular type III in children, types Ib and V were most common among adults. All 45 levofloxacin-resistant strains had two amino acid substitutions, Ser81Leu in the gyrA gene and Ser79Phe in the parC gene, and showed similar pulsed-field gel electrophoresis patterns.Streptococcus agalactiae (a group B streptococcus [GBS]) is the main microorganism causing meningitis and sepsis in infants and also sepsis in nonpregnant adults (12, 14).GBS infection in infants is classified as early onset, occurring in newborns within the first week of life, or late onset, developing in infants more than 1 week old, with most infections arising in the first 3 months and only extremely rarely in older infants (18). In the 1970s, morbidity and mortality from these GBS infections were high (3, 4, 9). In 1996, however, recommendations for the prevention of perinatal GBS infection were issued by the American College of Obstetricians and Gynecologists (2), the Centers for Disease Control and Prevention (7), and later also the American Academy of Pediatrics (1). As a result, preventive efforts increased and the incidence of early-onset disease decreased substantially (6, 23). A more detailed revised guideline, based on prenatal bacterial cultures and epidemiologic studies, was recommended in 2002 (17).Recently, Phares et al. (15) reported on a 7-year epidemiologic survey of invasive GBS disease in the United States that demonstrated a significant decline in the incidence of early-onset disease in infants, contrasting with an increase in GBS disease among adults ≥65 years old.In the present paper, we describe details concerning patient age, disease, and underlying diseases associated with invasive GBS infection, as well as the capsular types, antimicrobial susceptibilities, and resistance genes of isolates in Japan.Between August 2006 and July 2007, our laboratory received 189 GBS strains from the bacteriologic laboratories of 97 medical institutions participating in the Invasive Streptococcal Disease Working Group at the 19th Annual Meeting of the Japanese Society for Clinical Microbiology. All isolates were from sterile sites: blood (n = 124), cerebrospinal fluid (n = 54), pustule fluid (n = 7), joint fluid (n = 3), and tissue (n = 1).To identify the capsular type of GBS by PCR, we used nine sets of primers from types Ia to VIII as reported by Poyart et al. (16). We also applied our newly designed dltS primers for the identification of GBS (Table (Table11).

TABLE 1.

Primers for PCR and sequencing for FQ resistance in S. agalactiae
Gene and primerSequence (5′-3′)Length (mer)Amplicon size (bp)
dltS
    dlts-FCTGTAAGTCTTTATCTTTCTCG22199
    dlts-RTCCATTCGCTTAGTCTCC18
gyrA
    gyrA-FGGTTTAAAACCTGTTCATCGTCGT24407
    gyrA-RGCAATACCAGTTGCACCATTGACT24
gyrB
    gyrB-FCGAAGCTTTCAATCGATTCCTATT24495
    gyrB-RGGTCGCATAAAACGATAAATCAGAG25
parC
    parC-FCCGGATATTCGTGATGGCTT20403
    parC-RTGACTAAAAGATTGGGAAAGGC22
parE
    parE-FGCAAAGCAACTTCGATATGAAATTC25368
    parE-RCGGAGCTATTTACAGACAACGTTTT25
Open in a separate windowOne colony was picked up from each agar plate and placed in 30 μl of lysis solution containing 1 U of mutanolysin. The lytic reaction was carried out for 20 min at 60°C, followed by 5 min at 94°C. The lysate was added to each of five tubes containing PCR mixtures for individual capsular types: types Ia and Ib in tube A, types II and III in tube B, types IV and dltS in tube C, types V and VII in tube D, and types VI and VIII in tube E. The reaction mixture (25 μl) consisted of 20 pmol of each primer, 0.625 U of AmpliTaq Gold polymerase (Applied Biosystems, Tokyo, Japan), 2.5 μl of 10× PCR Gold buffer, 2.5 μl of 25 mM MgCl2, 2 μl of a 2 mM deoxynucleotide triphosphate mixture, and 16.875 μl of DNase- and RNase-free distilled water. DNA amplification was carried out with 40 cycles of 94°C for 1 min, 53°C for 2 min, and 72°C for 2 min.We measured the antimicrobial susceptibilities of GBS strains to 14 antibiotics including oral and parenteral agents by agar plate dilution methods using blood agar.Three genes for macrolide (ML) resistance, erm(A), erm(B), and mef(A), were identified with the three sets of primers and PCR conditions described previously (21).To identify fluoroquinolone (FQ) resistance, four sets of primers were designed based on the sequences of the gyrA, gyrB, parC, and parE genes (Table (Table1).1). The PCR mixture (50 μl) consisted of 20 pmol of each primer, 0.625 U of TaKaRa Ex Taq polymerase (Takara Bio, Kyoto, Japan), 5 μl of 10× Ex Taq buffer, 4 μl of the 2.5 mM deoxynucleotide triphosphate mixture, and 38.25 μl of DNase- and RNase-free distilled water. Amplified and purified DNA samples were sequenced with a BigDye Terminator cycle sequencing kit (version 3.1; Applied Biosystems, Foster City, CA). The pbp2x gene encoding the PBP2X enzyme, which mediates septum formation during cell wall synthesis, was also sequenced with primers reported previously (11).We performed pulsed-field gel electrophoresis (PFGE) on the 45 GBS strains determined to have FQ resistance according to mutations in the gyrA and parC genes. Plug-embedded GBS cells were lysed with lysozyme (5,000 U/3 ml) and mutanolysin (20 U/ml) at 50°C for 3 h by a modification described previously (5, 8). Chromosomal DNA was digested at 37°C for 18 h with ApaI (100 U/ml). PFGE was performed with 1% agarose and 0.5× TBE buffer (1× TBE is 90 mM Tris base, 88 mM boric acid, and 2 mM EDTA) at pulse times of 2.91 to 17.33 s, at an angle of 120°, at 6.0 V/cm, and at 14°C for 20 h with the CHEF Mapper (Bio-Rad Laboratories, Hercules, CA).Table Table22 shows relationships between capsular types of GBS pathogens and diagnoses, separately considering children ≤17 years old (n = 65) and adults (n = 124). Diseases were classified into meningitis, sepsis, and other infection groups. In children including newborns (10.8%) with early-onset disease and neonates (70.8%) with late-onset disease, capsular type III predominated at 67.7%, with small numbers of other types. Among adults, those at least ≥50 years old accounted for 83.1% of the cases; capsular type Ib predominated at 31.5%, followed by V (18.5%), II (12.1%), and III (12.1%). In addition to sepsis (75.0%), a variety of diseases were noted: cellulitis, arthritis, necrotizing fasciitis, meningitis, and bacterial endocarditis. Importantly, 88.7% of the affected adults had underlying disease such as diabetes, liver dysfunction, or immune compromise. Instances of death and neurologic sequelae included one of each among children, and eight (6.4%) and two (1.6%) among adults, respectively.

TABLE 2.

Correlation of capsular types of strains with 189 invasive GBS infections
Patient group and infectionCapsular type (no. of cases)
Total
IaIbIIIIIIVVVIVIIVIII
Children
    Meningitis3539350 (76.9)a
    Sepsis522514 (21.5)
    Other11 (1.5)
        Subtotal8 (12.3)8 (12.3)2 (3.1)44 (67.6)3 (4.6)65 (100)
Adults
    Meningitis1124 (0.8)
    Sepsis9311262061893 (75.0)
    Other272733327 (21.8)
        Subtotal11 (8.9)39 (31.5)15 (12.1)15 (12.1)23 (18.5)9 (7.3)1 (0.8)11 (8.9)124 (100)
Open in a separate windowaValues in parentheses are percentages.Table Table33 shows the MIC ranges and MICs for 50 and 90% of the strains tested (MIC50, and MIC90, respectively) of oral and intravenous antibiotics for GBS strains. The MIC range of β-lactam agents was narrow, and penicillin-resistant strains were not recognized. Notably, in a strain where cefotiam susceptibility was reduced to 2 μg/ml, four amino acid substitutions, Gly398 to Ala, Gln412 to Leu, His438 to Tyr, and Ile600 to Val, were identified in the pbp2x gene.

TABLE 3.

Susceptibilities of 189 S. agalactiae isolates to 14 antimicrobial agents
Delivery route and antibioticMIC rangeaMIC50aMIC90a
Oral
    Penicillin G0.016-0.1250.0630.063
    Ampicillin0.031-0.250.1250.125
    Amoxicillin0.031-0.250.0630.125
    Cefdinir0.016-0.1250.0310.063
    Cefditoren0.016-0.0630.0310.031
    Erythromycin0.016-≥640.032≥64
    Clarithromycin0.031-≥640.125≥64
    Clindamycin0.031-≥640.063≥64
    Levofloxacin0.5-≥642≥64
Intravenous
    Cefazolin0.063-0.50.1250.25
    Cefotiam0.125-20.50.5
    Cefotaxime0.016-0.1250.0310.063
    Panipenem0.008-0.0310.0160.031
    Meropenem0.031-0.1250.0630.063
Open in a separate windowaValues are in micrograms per milliliter.Table Table44 shows relationships between ML and FQ resistance and capsular type, separately considering children and adults. Of 23 strains showing ML resistance (12.2%), 3 possessed the erm(A) gene and 20 possessed the erm(B) gene. The M type was not recognized. ML-resistant strains detected in both children and adults were mostly type III, but a few strains showed other capsular types.

TABLE 4.

Correlation of capsular types with FQ and ML resistance
Patient group and resistance patternNo. of strains of serotype:
Total no. (%)
IaIbIIIIIIVVVIVIIVIII
Children
    FQr66 (9.2)
    MLr [erm(A)]22 (3.1)
    MLr [erm(B)]167 (10.8)
    Susceptible72236350 (76.9)
        Subtotal882440030065
Adults
    FQr3211135 (28.2)
    FQr MLr [erm(A)]11 (0.8)
    FQr MLr [erm(B)]21a3 (2.4)
    MLr [erm(B)]144110 (8.1)
    Susceptible11513819801175 (60.4)
        Subtotal113915150239111124
Open in a separate windowaThis strain showed three amino acid substitutions in PBP2X. The MICs of ampicillin and cefotiam for the strain were 0.25 and 2.0 μg/ml, respectively.In 45 strains showing high levofloxacin resistance (23.8%), two amino acid substitutions, Ser81 to Leu encoded by the gyrA gene and Ser79 to Phe encoded by the parC gene, were identified simultaneously. The capsular type of these strains, including six isolated from children, was predominately Ib, which was observed in 34 strains; other types (II, III, and VI) were each seen in a few strains.The PFGE patterns of 45 FQ-resistant strains are shown in Fig. Fig.1.1. These strains included 40 strains of type Ib and 5 strains representing other types. All type Ib strains showed highly homologous restriction patterns that differed clearly from those of type II or III strains.Open in a separate windowFIG. 1.PFGE patterns of levofloxacin-resistant S. agalactiae isolates. Each DNA sample was digested with the ApaI restriction enzyme. Lanes M, lambda ladder.In Japan, the proportion of the elderly population with underlying diseases has increased rapidly. As a consequence, invasive infections caused not only by GBS, but also S. dysgalactiae subsp. equisimilis and S. pneumoniae, are expected to increase gradually and to become serious problems (19, 20).The capsular type in isolates from newborns was mostly type III, in agreement with previous results. In most cases involving adults at least 50 years old, however, type Ib was predominant, followed by type V. These findings differ from previous epidemiologic data from the United States; the reason for this disparity is not known.The percentage of ML resistance was not particularly high compared with that in other countries. Much attention has been drawn to the emergence of GBS with reduced susceptibility to penicillin and cephalosporin antibiotics arising from mutations in the pbp2x gene (11). One of our collected strains had mutations of the pbp2x gene; this was a type III strain with multiple-antibiotic resistance to ML and FQ. FQ-resistant strains have been reported previously (10, 13, 22) but at extremely low rates. In our results, however, strains resistant only to FQ accounted for 23.8% of the isolates, and most of these were type Ib. FQ-resistant GBS from newborns, who had not been exposed to the agent, showed a PFGE pattern very similar to type Ib from adults. The observations suggest that a single clone acquired FQ resistance and spread rapidly throughout Japan.Antimicrobial use in Japan favors oral cephalosporins as the drugs of first choice for children, while oral FQ and ML, as well as cephalosporins, are often prescribed for adults. Notably, the size of individual doses of antimicrobials typically is small in Japan compared with that in other countries. These factors will expand the mutant selection window for many pathogens, including GBS, and thus may cause an increase in resistant microorganisms.To control the emergence of resistant organisms, continuous molecular epidemiologic surveillance for pathogens is needed.  相似文献   

12.
This study assessed the in vitro activities of ceftaroline and five comparator agents against a collection of Staphylococcus aureus isolates. Ceftaroline demonstrated potent activity against community-associated methicillin-resistant S. aureus (CA-MRSA) isolates and showed bactericidal activity against vancomycin-intermediate S. aureus (VISA), vancomycin-resistant S. aureus (VRSA), heteroresistant VISA (hVISA), and daptomycin-nonsusceptible S. aureus (DNSSA) isolates. Ceftaroline may represent a bactericidal treatment option for infections caused by these pathogens.The increasing prevalence of resistant Staphylococcus aureus strains, including methicillin-resistant S. aureus (MRSA), community-associated MRSA (CA-MRSA), and S. aureus strains with reduced susceptibility to vancomycin, emphasizes the need for innovative antimicrobials with activity against these pathogens (19, 24). Although the prevalence of S. aureus strains with reduced vancomycin susceptibility remains low, such strains have been associated with vancomycin treatment failure, limiting the treatment options for patients with such infections (3, 8). Ceftaroline is a novel, parenteral, broad-spectrum cephalosporin exhibiting bactericidal activity against Gram-positive organisms, including MRSA and multidrug-resistant Streptococcus pneumoniae (MDRSP), as well as common Gram-negative pathogens (6, 9, 21). Ceftaroline is currently in phase 3 development for the treatment of complicated skin and skin structure infections and community-acquired bacterial pneumonia. The study described here evaluated the in vitro activities of ceftaroline and five comparator agents against CA-MRSA, vancomycin-intermediate S. aureus (VISA), vancomycin-resistant S. aureus (VRSA), heteroresistant VISA (hVISA), and daptomycin-nonsusceptible S. aureus (DNSSA) isolates.(A preliminary report of these results was presented at the 48th Interscience Conference on Antimicrobial Agents and Chemotherapy-Infectious Diseases Society of America Annual Meeting, Washington, DC, 25 to 28 October 2008.)A total of 132 MRSA strains were selected for evaluation. CA-MRSA strains (n = 92) were isolated from patients admitted to St. John Hospital and Medical Center in Detroit, MI. These patients had positive MRSA cultures within 48 h of admission, in accordance with the definition of CA-MRSA described by the Centers for Disease Control and Prevention (2). DNSSA strains (n = 7) and hVISA strains (n = 3) were obtained from blood collected from patients at the same hospital. The hVISA isolates were identified by a modified population analysis profile method (12). VISA isolates (n = 20) and VRSA isolates (n = 10) were obtained via the Network on Antimicrobial Resistance in Staphylococcus aureus (NARSA) program, supported under NIAID/NIH contract HHSN272200700055C.All CA-MRSA isolates were typed by pulsed-field gel electrophoresis (PFGE) with the restriction enzyme SmaI, followed by visual interpretation and categorization. PCR methods were used to determine the staphylococcal cassette chromosome mec (SCCmec) type, the presence of Panton-Valentine leukocidin (PVL) genes lukS-PV and lukF-PV, and the presence of the arginine catabolic mobile element (ACME) via detection of the arcA locus.In vitro susceptibility tests were performed with the antimicrobials ceftaroline (lot CI 170-07; Forest Laboratories, Inc., New York, NY), vancomycin-HCl (Sigma), daptomycin (Cubist), clindamycin-HCl (Sigma), linezolid (Pfizer), and trimethoprim-sulfamethoxazole (Sigma), which were obtained individually and reconstituted to a 1:19 ratio. All antibiotics were received in powder form and were reconstituted according to guidelines of the Clinical and Laboratory Standards Institute (CLSI). MICs and minimum bactericidal concentrations (MBCs) were determined according to the guidelines of the CLSI (4, 5). Microdilution tests with cation-adjusted Mueller-Hinton broth were used to identify the MICs of all antimicrobial agents tested. The percentages of susceptible isolates were determined by using the CLSI breakpoints. For the testing of daptomycin, additional calcium was added to the broth for a final concentration of 50 mg/liter. The MICs were read visually and corresponded to the concentration in the well with the lowest drug concentration with no visible bacterial growth. The MBC was defined as the antibiotic concentration that reduced the number of viable cells by ≥99.9%. This was based on colony counts from the growth control well and rejection values determined by tables provided by Pearson (18).S. aureus ATCC 29213 and Enterococcus faecalis ATCC 29212 were used as the control strains for the MIC determinations, and S. aureus ATCC 25923 was used as the control strain for the MBC determinations.Approximately 58% of the CA-MRSA isolates were SCCmec type IV/IVa, whereas the majority of the other isolates were SCCmec type II (Table (Table1).1). All SCCmec type IVa isolates were positive for the PVL and ACME genes, whereas isolates of all other SCCmec types were negative for these genes. Of all the other S. aureus strains, only two DNSSA isolates were SCCmec type IVa and positive for the PVL and ACME genes.

TABLE 1.

SCC mec elements, PVL and ACME genes, and ceftaroline MIC90s and MBC90s of CA-MRSA isolatesa
SCCmec typePVLACMENo. of isolatesCeftaroline MIC90 (μg/ml)Ceftaroline MBC90 (μg/ml)PFGE strain(s)
IVa++400.50.5USA 300 (n = 40)
IV130.250.5USA 100 (n = 4), 9 genetically unrelated isolates
II3811USA 100 (n = 15), USA 600 (n = 2), 21 genetically unrelated isolates
Not typeable1NAbNAGenetically unrelated to the other isolates
Open in a separate windowaA total of 92 CA-MRSA isolates were tested: 26 from blood, 33 from the respiratory tract, and 33 from wounds or tissues.bNA, not applicable.The highest ceftaroline MIC observed for the CA-MRSA, VISA, VRSA, and DNSSA isolates was 1 μg/ml (Table (Table2).2). The MICs of ceftaroline were not influenced by traits conferring resistance to other classes of antimicrobials. All isolates susceptible to the drugs ceftaroline, vancomycin, daptomycin, and trimethoprim-sulfamethoxazole demonstrated bactericidal activity, with the MIC90/MBC90 ratio being less than or equal to 1:2. The MBC90 of 1 μg/ml for ceftaroline, vancomycin, and daptomycin against CA-MRSA isolates was equal to the MIC90, indicative of bactericidal activity, which was not observed for the bacteriostatic agents linezolid and clindamycin.

TABLE 2.

MIC50/MIC90 and MBC50/MIC90 values for all antimicrobials tested for their activities against CA-MRSA, VISA, hVISA, VRSA, and DNSSA isolatesa
Isolate and antimicrobialMIC (μg/ml)
% S% RMBC (μg/ml)
Range50%90%bRange50%90%b
CA-MRSA (n = 92)
    CPT0.25-10.51NANA0.25-10.51
    VAN0.5-21110000.5-211
    DAP0.25-10.511000.25-20.51
    LZD1-4221004->8>8>8
    CLI0.06->640.12>6464361->648>64
    SXT1.2/0.06->76/42.4/0.129.5/0.59821.2/0.06->76/42.4/0.1219/1
VISA (n = 20) and hVISA (n = 3)
    CPT0.25-10.51NANA0.5-211
    VAN1-8481301-848
    DAP0.5-824350.5-824
    LZD0.5-2221001->84>8
    CLI0.06->64>64>6417830.12->64>64>64
    SXT1.2/0.06->76/44.8/0.25>76/478222.4/0.12->76/49.5/0.5>76/4
VRSA (n = 10)
    CPT0.12-10.50.5NANA0.12-10.51
    VAN32->64>64>64010064->64>64>64
    DAP0.5-10.511000.5-111
    LZD1-4221008->8>8>8
    CLI>64>64>640100>64>64>64
    SXT1.2/0.06->76/42.4/0.1238/290102.4/0.12->76/42.4/0.12>76/4
DNSSA (n = 7)
    CPT0.25-10.50.55NANA0.25-110.74
    VAN1-221.641000222
    DAP44404-885.94
    LZD1-221.481002->8>810.7
    CLI<0.03->64>643514862->64>6470.66
    SXT1.2/0.06->76/42.4/0.125.6/0.4071292.4/0.12->76/44.8/0.2511.65/0.60
Open in a separate windowaAbbreviations: CPT, ceftaroline; VAN, vancomycin; DAP, daptomycin; LZD, linezolid; CLI, clindamycin; SXT, trimethoprim-sulfamethoxazole; NA, not available; S, susceptible; R, resistant.bFor DNSSA, geometric mean MICs and MBCs are used in place of MIC90s and MBC90s, respectively, because less than 10 isolates were studied.Previous studies have associated CA-MRSA isolates containing SCCmec type I through III genes with multidrug-resistant nosocomial infections, whereas the SCCmec type IV gene is typically found in CA-MRSA strains susceptible to various antibiotics (7, 13, 16, 22). In our study, approximately 58% of the CA-MRSA isolates were characterized as SCCmec type IV or IVa, and 41% were characterized as SCCmec type II. The MICs of ceftaroline against CA-MRSA observed in this study correlate well with the activity of ceftaroline previously reported against this pathogen (20). Ceftaroline demonstrated in vitro activity against all CA-MRSA isolates, regardless of their SCCmec type, with MICs ranging from 0.25 to 1 μg/ml and with the MIC90 being 1 μg/ml. These findings are supported by those of earlier studies, in which the ceftaroline MICs ranged from 0.25 to 1 μg/ml for CA-MRSA and from 0.12 to 2 μg/ml for MRSA (1, 6, 10, 14, 15, 20, 21).Daptomycin nonsusceptibility among VISA isolates has been reported (11, 17), potentially limiting the treatment options for patients infected with these pathogens. In the present study, ceftaroline demonstrated activity against all VISA isolates, including those not susceptible to daptomycin (Table (Table3).3). Recent work by Vidaillac et al. has evaluated the in vitro activity of ceftaroline against MRSA and hVISA strains by using an experimental pharmacokinetic/pharmacodynamic model and has shown that ceftaroline not only demonstrated activity equal to or greater than that of vancomycin but also had a lower potential to select for resistant mutants (23).

TABLE 3.

SCCmec elements, PVL and ACME genes, and MICs of VISA and VRSA isolatesa
IsolateSCCmec typeMIC (μg/ml)
CPTVANDAPLZDSXTCLI
VISA
    NRS-3II184219/1>64
    NRS-4II0.54122.4/0.12>64
    NRS-17II0.58412.4/0.120.06
    NRS-18II0.54122.4/0.12>64
    NRS-19II0.54219.5/0.5>64
    NRS-22II0.5421>76/4>64
    NRS-23II0.5422>76/4>64
    NRS-24II0.54229.5/0.5>64
    NRS-26II0.25441>76/4>64
    NRS-27II0.5411>76/4>64
    NRS-51II14122.4/0.12>64
    NRS-68II0.54114.8/0.25>64
    NRS-73IVd0.5421>76/40.06
    NRS-74II0.254422.4/0.12>64
    NRS-76II0.254222.4/0.120.06
    NRS-118I184119/1>64
    NRS-126II14222.4/0.12>64
    NRS-402II0.5880.51.2/0.06>64
    NRS-403II0.54212.4/0.12>64
    NRS-404II0.58229.5/0.50.06
VRSA
    VRS-1II1>641238/2>64
    VRS-2II0.5320.5219/1>64
    VRS-3IV0.5320.521.2/0.06>64
    VRS-4IV0.5>640.522.4/0.12>64
    VRS-5II0.5>640.522.4/0.12>64
    VRS-6II0.5>64141.2/0.06>64
    VRS-7NT0.12>64111.2/0.06>64
    VRS-8NT0.12>640.511.2/0.06>64
    VRS-9II0.25>640.51>76/4>64
    VRS-10II0.5>640.5238/2>64
Open in a separate windowaAbbreviations: NT, not typeable; CPT, ceftaroline; VAN, vancomycin; DAP, daptomycin; LZD, linezolid; SXT, trimethoprim-sulfamethoxazole; CLI, clindamycin.This study assessed the in vitro activities of ceftaroline and five comparator agents against a collection of S. aureus isolates characterized into USA types by PFGE. The SCCmec elements were characterized, and the presence of PVL and ACME genes was determined. Ceftaroline demonstrated bactericidal activity against CA-MRSA, VISA/hVISA, VRSA, and DNSSA isolates. Ceftaroline represents a bactericidal option for the treatment of MRSA infections, including those caused by isolates with reduced susceptibilities to vancomycin and daptomycin, and should undergo further clinical studies.  相似文献   

13.
In vitro antistaphylococcal activities of panduratin A, a natural chalcone compound isolated from Kaempferia pandurata Roxb, were compared to those of commonly used antimicrobials against clinical staphylococcal isolates. Panduratin A had a MIC at which 90% of bacteria were inhibited of 1 μg/ml for clinical staphylococcal isolates and generally was more potent than commonly used antimicrobials.Staphylococci are frequently refractory to many new and commonly used antimicrobial agents and have become a problem in recent years (8, 12, 17). Methicillin (meticillin)-resistant Staphylococcus aureus (MRSA) infections have emerged as a worldwide problem, and clinical strains of MRSA exhibit reduced susceptibility to antimicrobial agents (18). Moreover, coagulase-negative staphylococci are well established due to nosocomial bacteremia and indwelling medical device-associated infection, showing increased multidrug resistance (1, 14). Thus, the identification of novel agents effective in inhibiting these strains has gained renewed urgency (7). In addition, there is renewed interest in plants with antimicrobial properties as a consequence of current problems associated with the use of antibiotics (4, 9). Panduratin A, a natural chalcone compound isolated from the rhizome of fingerroot (Kaempferia pandurata Roxb.), has been reported to possess antibacterial activity against Prevotella intermedia, Prevotella loescheii, Porphyromonas gingivalis, Propionibacterium acnes, and Streptococcus mutans, as well as antibiofilm activity against multispecies oral biofilms in vitro (6, 13, 15, 16). However, antimicrobial activities of panduratin A against other pathogenic bacteria, such as staphylococci, have not yet been investigated.In this study, we compared the in vitro activities of panduratin A against MRSA, methicillin-susceptible S. aureus (MSSA), methicillin-resistant coagulase-negative staphylococci (MRCNS), and methicillin-susceptible coagulase-negative staphylococci (MSCNS) with those of treatments with available antimicrobial agents, such as ampicillin, erythromycin, gentamicin, levofloxacin, linezolid, oxacillin, tetracycline, thymol, and vancomycin.Clinical MRSA (n = 27), MSSA (n = 27), MRCNS (n = 28), and MSCNS (n = 26) were obtained from the Research Institute of Bacterial Resistance, College of Medicine, Yonsei University, South Korea. The clinical Staphylococcus strains were collected in 2008 from patients at a Korean tertiary-care hospital. The strains were isolated from body fluids, blood, genital secretions, pus, or sputum, of the patient. The species were identified by conventional methods (2) or by using the Vitek system (bioMerieux SA, Marcy l''Etoile, France) according to the manufacturer''s instructions. Reference strains S. aureus ATCC 29213 and Staphylococcus epidermidis ATCC 12228 from the American Type Culture Collection (Rockville, MD) were included as controls.Panduratin A (FIG. (FIG.1)1) was isolated in pure form from an ethanol extract of Kaempferia pandurata Roxb. according to the method of Park et al. (13). Panduratin A was dissolved in 10% dimethyl sulfoxide (DMSO) to obtain a 1,024-μg/ml stock solution. Ampicillin, erythromycin, gentamicin, tetracycline, thymol, and vancomycin were purchased from Sigma-Aldrich. Co. (St. Louis, MO). Levofloxacin and oxacillin were purchased from Sigma-Fluka Co. (Steinheim, Germany), and linezolid was provided by Dong-A Pharmaceutical Co. (Seoul, South Korea). Stock solutions of commercial antimicrobial agents were prepared according to the manufacturer''s instructions.Open in a separate windowFIG. 1.Structure of panduratin A.In vitro susceptibility tests were performed in a 96-well microtiter plate to determine MICs of panduratin A and other antimicrobial agents against 108 isolates of clinical staphylococci using standard broth microdilution methods with an inoculum of 5 × 105 CFU/ml, according to the guidelines of CLSI standard M7-A6 (3). A twofold dilution of panduratin A stock solution or other antimicrobial agent preparation was mixed with the test organisms (5 × 105 CFU/ml) in Mueller-Hinton broth (MHB) medium (Difco Becton Dickinson, Sparks, MD). Column 12 of the microtiter plate contained the highest concentrations of panduratin A or other antimicrobial agents, and column three contained the lowest concentrations of panduratin A or other antimicrobials agents. Column 2 served as the positive control for all samples (only medium and inoculum or antimicrobial agent-free wells), and column 1 was the negative control (only medium, no inoculum, and no antimicrobial agent). Microtiter plates were incubated aerobically at 37°C for 24 h. The MIC was defined as the lowest concentration of antimicrobial agent that resulted in the complete inhibition of visible growth.Panduratin A was diluted in 10% DMSO, followed by twofold dilutions in the test wells; thus, the final concentration of DMSO would be serially decreased. We examined the effect of DMSO on the growth and viability of all staphylococci tested. DMSO at ≤10% was found not to affect growth or viability of the staphylococci tested. These results suggest that DMSO had no effect on activity and that all the antimicrobial activity was due to panduratin A.Minimal bactericidal concentrations (MBCs) were determined for each antimicrobial agent per Staphylococcus strain as outlined for MICs (5). Briefly, medium (approximately 100 μl) from each well showing no visible growth was spread onto MHA (MHB supplemented with 1.5% bacterial agar) plates. Wells in column 2, the positive controls (antimicrobial agent-free wells), and wells in column 1, growth-negative controls, were included for the MBC test. Plates were incubated at 37°C for 24 h or until growth was seen in the growth-positive control plates. MBC was defined as the lowest concentration of antimicrobial agent at which all bacteria in the culture are killed or the lowest concentration at which no growth occurs on MHA plates (5, 10).Table Table11 shows the MICs and MBCs of panduratin A in comparison to those of ampicillin, erythromycin, gentamicin, levofloxacin, linezolid, oxacillin, tetracycline, thymol, and vancomycin for clinical staphylococci isolates. In this study, all isolates were susceptible to panduratin A, with MICs of ≤2 μg/ml. In our previous report (13), the MIC of panduratin A against P. gingivalis, P. loescheii, and S. mutans was 4 μg/ml while that of panduratin A against P. intermedia and P. acnes was 2 μg/ml (6, 13, 15). These results show that panduratin A has activities against clinical staphylococci stronger than those against P. gingivalis, P. loescheii, and S. mutans and comparable or equal to those against P. intermedia and P. acnes. Moreover, panduratin A has the capability of preventing the biofilm formation of primary multispecies oral bacteria (Actinomyces viscosus, S. mutans, and Streptococcus sanguis) in vitro (16). This report suggests that panduratin A might also have the ability to inhibit staphylococcal biofilm formation. Hence, future research is necessary to determine the inhibition activity of panduratin A against staphylococcal biofilm formation.

TABLE 1.

Comparative in vitro activities of panduratin A and other antimicrobial agents against clinical staphylococcal isolates
Staphylococcal group (na) or antimicrobial agentMIC (μg/ml)
Susceptibility (%)b
MBC (μg/ml)
Range50%90%SIRRange50%90%
MRSA (27)
    Ampicillin16-1281664010032-51264256
    Erythromycin16-6432640010032-256128256
    Gentamicin0.5-6432641522632-25664128
    Levofloxacin0.5-2568128180820.5-51216512
    Linezolid0.5-2121008-16816
    Oxacillin32-6464640100256-512256512
    Tetracycline0.5-641632300702-51264512
    Thymol64-128128128256-512512512
    Vancomycin0.25-10.50.5100000.5-812
    Panduratin A0.5-10.512-824
MSSA (27)
    Ampicillin0.5-64323201001-25664128
    Erythromycin0.5-650.532564401-2564128
    Gentamicin0.125-64464520480.5-2568128
    Levofloxacin0.125-640.1258814150.125-1618
    Linezolid0.25-814962-32832
    Oxacillin0.125-320.125189-111-1281664
    Tetracycline0.5-32487019111-64832
    Thymol64-128128128128-512256512
    Vancomycin0.125-20.250.5100000.25-812
    Panduratin A0.5-20.511-824
MRCNS (28)
    Ampicillin0.5-128166401002->51232256
    Erythromycin0.5-3216321115741-12832128
    Gentamicin0.125-64864457480.25-1288128
    Levofloxacin0.25-80.2542615590.5-32416
    Linezolid0.125-40.2521000.5-824
    Oxacillin0.125-64646426741-256128256
    Tetracycline0.25-64164187750.5-1284128
    Thymol32-5126412864-512256512
    Vancomycin0.125-20.1252100000.5-824
    Panduratin A0.125-20.2511-824
MSCNS (26)
    Ampicillin0.5-12821601001-2564128
    Erythromycin0.5-1280.5165023271-2562128
    Gentamicin0.125-640.12532730270.25-64164
    Levofloxacin0.25-80.250.596040.5-3214
    Linezolid0.125-160.254960.25-3224
    Oxacillin0.125-320.1250.59640.25-12828
    Tetracycline0.5-1280.5326919121-2568128
    Thymol4-128646464-512256512
    Vancomycin0.063-10.251100000.125-414
    Panduratin A0.063-20.510.125-414
Open in a separate windowan, no. of isolates tested.bS, susceptible; I, intermediate; R, resistant; —, CLSI breakpoint is not available.In this study, all isolates of MRSA, MSSA, MRCNS, and MSCNS were resistant to ampicillin. However, all isolates of MRSA, MSSA, MRCNS, and MSCNS were inhibited by ≤2 μg/ml of panduratin A. MICs of panduratin A against all isolates tested were much lower than those of thymol (≤512 μg/ml), which has been reported to possess antistaphylococcal activity (11). Moreover, most isolates of MRSA were resistant to erythromycin, gentamicin, levofloxacin, oxacillin, and tetracycline. Although all isolates of MRSA were inhibited by ≤2 μg/ml of linezolid, these MICs of ≤2 μg/ml were still higher than that of panduratin A or vancomycin, which inhibited the growth of all isolates of MRSA with MICs of ≤1 μg/ml.The majority of MSSA isolates were susceptible to erythromycin (MIC at which 90% of bacteria were inhibited [MIC90] = 32 μg/ml), gentamicin (MIC90 = 64 μg/ml), levofloxacin (MIC90 = 8 μg/ml), linezolid (MIC90 = 4 μg/ml), oxacillin (MIC90 = 1 μg/ml), tetracycline (MIC90 = 8 μg/ml), and vancomycin (MIC90 = 0.5 μg/ml). However, the MIC90 of panduratin A was 1 μg/ml. These results indicate that panduratin A has stronger antistaphylococcal activity against MSSA isolates than erythromycin, gentamicin, levofloxacin, linezolid, or tetracycline.The MRCNS isolates were also resistant to most of the antimicrobial agents tested. All MRCNS isolates were inhibited by ≤2 μg/ml of vancomycin and ≤4 μg/ml of linezolid. The MIC range and MIC90 of panduratin A for MRCNS isolates was 0.125 to 2 μg/ml and 1 μg/ml, respectively. These results indicate that antistaphylococcal activity of panduratin A against MRCNS is equal to that of vancomycin and stronger than that of linezolid.Finally, the MIC range of panduratin A against MSCNS isolates (0.063 to 2 μg/ml) was narrower than those of erythromycin, gentamicin, levofloxacin, linezolid, oxacillin, and tetracycline. Vancomycin had the narrowest range of MIC (0.063 to 1 μg/ml) against MSCNS isolates. Interestingly, the MIC90 of vancomycin against MSCNS isolates was the same as the MIC90 of panduratin A against MSCNS isolates.The range of MICs of panduratin A for MRSA and MSSA were very narrow at 0.5 to 1 μg/ml and 0.5 to 2 μg/ml, respectively. In contrast, the ranges of panduratin A MICs for MRCNS and MSCNS were large at 0.125 to 2 μg/ml and 0.063 to 2 μg/ml. These results could be interpreted to mean that MRSA and MSSA are composed of only one species of Staphylococcus, S. aureus, whereas MRCNS and MSCNS are composed of different species of Staphylococcus: S. hominis, S. epidermidis, S. haemolyticus, S. simulans, and S. sciuri. Yong et al. (17) reported that the ranges of MICs for DA-7867, a novel oxazolidinone, for MRSA and MSSA were broader than those for MRCNS and MSCNS. Moreover, the ranges of MICs for CG400549, a novel FaI inhibitor, for MRSA and MSSA were very narrow at 0.12 to 0.5 μg/ml and 0.12 to 1 μg/ml, respectively. In contrast, the ranges of CG400549 MICs for MRCNS and MSCNS were broad at 0.12 to 16 μg/ml and 0.5 to 8 μg/ml (18). Thus, the MICs of panduratin A for MRSA, MSSA, MRCNS, and MSCNS were in agreement with other reports. In addition, the MIC of panduratin A against P. intermedia was 2 μg/ml, whereas that against P. loescheii was 4 μg/ml. They belong to the same genus, Prevotella, but are different species (13). Thus, coagulase-negative staphylococci (MRCNS and MSCNS) have a wider MIC dispersion with panduratin A than that of coagulase-positive staphylococci (MRSA and MSSA).The in vitro MBCs of panduratin A with an endpoint after 24 h demonstrated that panduratin A was able to kill staphylococcus strains with MBCs of ≤8 μg/ml for MRSA, MSSA, and MRCNS. On the other hand, panduratin A can kill MSCNS with MBCs of ≤4 μg/ml. These results were similar to the MBCs of vancomycin against the clinical staphylococcal strains (Table (Table1).1). These panduratin A MBC results suggest that panduratin A may be as bactericidal as vancomycin. In addition, the MBC of panduratin A against a P. gingivalis, P. loescheii, and S. mutans was 8 μg/ml, and the MBC of panduratin A against P. intermedia and P. acnes was 4 μg/ml (6, 13, 15). Panduratin A has been reported to have the ability to reduce the biofilm of multispecies oral bacteria in vitro (16). It would be interesting to evaluate the antibiofilm activity of panduratin A in reducing staphylococcal biofilms. Further work toward these objectives may resolve these issues.In conclusion, panduratin A is an antimicrobial agent with high in vitro activities against clinical MRSA, MSSA, MRCNS, and MSCNS, including organisms resistant to other antimicrobials. These results suggest that panduratin A should undergo further testing to assess its potential for the treatment of diseases caused by staphylococci. Obviously, toxicity studies, animal model studies, and human clinical trials will determine whether in vitro microbiological results translate into a useful drug for treating human infections.  相似文献   

14.
A collection of 2,278 isolates belonging to 86 different fungal species was tested with micafungin and eight other drugs using the EUCAST procedures. Micafungin was active against species of Candida and Aspergillus (even azole-resistant species) as well as Penicillium spp., Scedosporium apiospermum, and Acremonium spp. It was inactive for species of Basidiomycota and Mucorales and for multiresistant species such as those of Fusarium.Micafungin is a new drug that belongs to the echinocandin class of antifungal agents. Its mechanism of action is by means of the inhibition of 1,3-β-d-glucan synthesis in the fungal cell wall (10).Micafungin has been recently approved in Europe and the United States for the treatment of candidemia, acute disseminated candidiasis, Candida peritonitis and abscesses, esophageal candidiasis, and recently for the prophylaxis of Candida infections in patients undergoing hematopoietic stem cell transplantation.The in vitro activity of micafungin against most common species of Candida is well known (4, 11-13). However, information is limited for uncommon species of yeasts as well as for molds.The aim of this study is to analyze the in vitro activity of micafungin and eight other antifungal agents against a collection of clinical isolates of yeasts and molds from human beings using the methods approved by AFST-EUCAST.The strains were recovered from 115 Spanish hospitals through a period of 3 years, from 2005 to 2007. A total of 2,278 clinical isolates were included in the analysis. Isolates were identified by morphological and biochemical methods and sequencing of DNA targets if necessary. They belonged to 86 different species of common and emerging fungal pathogens. The isolates were obtained from blood (559; 24.5%), biopsies and other deep sites (217; 9.5%), respiratory tract specimens (751; 33%), skin samples (180; 7.9%), and other locations (707; 25.1%).The following drugs were used: amphotericin B (range, 16.0 to 0.03 μg/ml; Sigma-Aldrich Quimica S.A., Madrid, Spain), flucytosine (64.0 to 0.12 μg/ml; Sigma-Aldrich), fluconazole (64.0 to 0.12 μg/ml; Pfizer S.A. Madrid, Spain), itraconazole (8.0 to 0.015 μg/ml; Janssen S.A., Madrid, Spain), voriconazole (8.0 to 0.015 μg/ml; Pfizer S.A.), posaconazole (8.0 to 0.015 μg/ml; Schering-Plough, Kenilworth, NJ), caspofungin (16.0 to 0.03 μg/ml; Merck & Co., Inc., Rahway, NJ), micafungin (16.0 to 0.03 μg/ml; Astellas Pharma Inc., Tokyo, Japan), and anidulafungin (16.0 to 0.03 μg/ml; Pfizer S.A.).Susceptibility testing was performed by using broth microdilution. For Candida species, MICs were determined using the reference procedure for testing fermentative yeasts described by AFST-EUCAST (7, 17). For Cryptococcus neoformans and other species of nonfermentative yeasts, such as Trichosporon and Rhodotorula spp., susceptibility testing strictly followed the recommendations by the EUCAST with a minor modification in order to improve the growth of microorganisms (3). For filamentous fungi, broth microdilution testing was performed following the AFST-EUCAST reference method (18). For testing echinocandins against molds, the MIC was defined as the lowest drug concentration resulting in aberrant hyphal growth by examination with an inverted microscope, that is, the minimum effective concentration (MEC) (2).Tables Tables11 and and22 display the susceptibility results obtained when the collection of clinical isolates was tested.

TABLE 1.

Summary of susceptibility results of antifungal agents tested in vitro against yeast speciesa
SpeciesnAmphotericin B
Flucytosine
Fluconazole
Itraconazole
Voriconazole
Posaconazole
Caspofungin
Anidulafungin
Micafungin
MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90
Candida species
    Candida albicans3930.060.120.120.500.120.250.020.030.020.020.020.020.060.120.030.030.030.03
    Candida parapsilosis2250.120.250.120.250.501.00.020.030.020.020.020.031.01.01.02.00.51.0
    Candida tropicalis1050.120.250.120.250.252.00.020.060.030.120.020.030.120.250.030.030.030.03
    Candida glabrata1820.120.250.120.254.032.00.251.00.121.00.251.00.120.250.030.030.030.03
    Candida krusei530.500.504.08.032.0>64.00.250.250.250.500.120.250.250.500.030.060.030.06
    Candida guilliermondii200.060.500.120.254.0>64.00.25>8.00.12>8.00.12>8.00.501.01.08.00.121.0
    Candida lusitaniae210.060.120.1216.00.251.00.020.030.020.020.020.020.501.00.030.120.030.03
    Candida kefyr150.250.500.504.00.250.500.020.060.020.020.030.060.060.120.030.030.030.03
    Other Candida spp.b290.121.00.2516.08.0>64.00.12>8.00.12>8.00.03>8.00.060.500.060.250.030.12
Other Ascomycota yeasts
    Dipodascus capitatus300.500.500.121.016.032.00.250.500.501.00.500.50>16.0>16.02.04.0>16.0>16.0
    Saccharomyces cerevisiae250.120.250.120.504.0>64.00.502.00.060.500.250.500.501.00.120.250.030.06
    Yarrowia lipolytica100.500.50>64.0>64.08.0>64.00.50>8.00.502.00.252.00.501.00.060.120.031.0
    Galactomyces geotrichum100.250.500.120.2564.0>64.00.502.00.501.00.500.50>16.0>16.0>16.0>16.0>16.0>16.0
Basidiomycota yeasts
    C. neoformans var. neoformans350.122.016.032.016.0>64.00.120.500.120.500.120.50>16.0>16.0>16.0>16.0>16.0>16.0
    Trichosporon asahii138.0>16.032.0>64.016.0>64.00.50>8.00.25>8.00.25>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Rhodotorula mucilaginosa110.120.500.251.0>64.0>64.08.0>8.04.0>8.02.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Trichosporon inkin100.121.064.0>64.04.04.00.120.250.060.120.060.12>16.0>16.0>16.0>16.0>16.0>16.0
    Other Basidiomycota yeastsc350.508.016.0>64.016.0>64.00.50>8.00.50>8.00.25>8.0>16.0>16.0>16.0>16.0>16.0>16.0
Open in a separate windowaMIC50 and MIC90 values (μg/ml) were calculated for those species with 10 or more isolates.b“Other Candida species” includes the following species with less than 10 isolates: Candida rugosa (4), Candida famata (4), Candida pelliculosa (4), Candida colliculosa (3), Candida norvegensis (2), Candida intermedia (2), Candida inconspicua (2), Candida metapsilosis (2), Candida orthopsilosis (2), Candida pintolopesii (2), and Candida zeylanoides (2).c“Other Basidiomycota yeasts” includes the following species with less than 10 isolates: Trichosporon ovoides (5), Trichosporon dermatis (5), Trichosporon jirovecii (4), Cryptococcus albidus (3), Ustilago spp. (3), Rhodotorula glutinis (3), Cryptococcus laurentii (2), Trichosporon domesticum (2), Trichosporon mycotoxinivorans (2), and Trichosporon spp. (6).

TABLE 2.

Summary of susceptibility results of antifungal agents tested in vitro against mold speciesa
SpeciesnAmphotericin B
Itraconazole
Voriconazole
Posaconazole
Caspofungin
Anidulafungin
Micafungin
MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90MIC50MIC90
Aspergillus spp.
    Aspergillus fumigatus2990.250.500.250.500.501.00.060.250.250.500.030.030.030.03
    Aspergillus terreus1551.04.00.250.501.02.00.060.121.02.00.030.030.030.03
    Aspergillus flavus811.02.00.250.501.02.00.120.25>16.0>16.0>16.0>16.0>16.0>16.0
    Aspergillus niger830.250.251.0>8.01.02.00.250.500.250.500.030.030.030.06
    Aspergillus nidulans291.04.00.250.500.251.00.060.250.50>16.00.030.250.03>16.0
    Aspergillus sydowii281.02.00.50>8.01.02.00.250.500.251.00.030.030.030.03
    Aspergillus versicolor121.02.00.501.01.02.00.250.500.252.00.030.030.030.06
    Other Aspergillus spp.b180.50>16.00.50>8.00.25>8.00.12>8.00.120.500.060.250.060.25
Other hyaline fungi
    Penicillium spp.720.502.00.50>8.02.0>8.00.254.00.258.00.030.030.030.06
    Fusarium solani321.02.0>8.0>8.0>8.0>8.0>8.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Fusarium proliferatum192.04.0>8.0>8.08.0>8.0>8.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Fusarium oxysporum171.02.0>8.0>8.04.0>8.04.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Fusarium verticillioides112.04.0>8.0>8.08.0>8.02.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Paecilomyces variotii170.030.500.060.251.08.00.030.250.504.00.030.030.030.03
    Paecilomyces lilacinus10>16.0>16.0>8.0>8.00.504.00.250.50>16.0>16.0>16.0>16.0>16.0>16.0
    Acremonium spp.104.0>16.0>8.0>8.04.08.08.0>16.00.501.00.030.030.120.25
    Other hyaline fungic260.50>16.0>8.0>8.04.0>8.04.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
Black fungi
    Scedosporium apiospermum364.0>16.01.0>8.00.502.01.08.02.0>16.00.504.00.03>16.0
    Scedosporium prolificans17>16.0>16.0>8.0>8.0>8.0>8.0>8.0>8.08>16.04.0>16.08.0>16.0
    Alternaria alternata110.50>16.00.50>8.04.0>8.01.0>8.04.0>16.00.06>16.00.25>16.0
    Alternaria infectoria100.06>16.02.0>8.02.0>8.00.50>8.0>16.0>16.02.0>16.0>16.0>16.0
    Other black fungid140.251.00.12>8.00.06>8.00.06>8.00.251.00.030.060.030.06
Mucorales
    Mycocladus corymbiferus160.120.251>8.0>8.0>8.00.50>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Rhizopus oryzae110.502.04.0>8.0>8.0>8.02.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Mucor spp.110.121.0>8.0>8.0>8.0>8.04.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
    Other Mucorales speciese112.0>16.0>8.0>8.0>8.0>8.04.0>8.0>16.0>16.0>16.0>16.0>16.0>16.0
Open in a separate windowaMIC50 and MIC90 values (μg/ml) were calculated for those species with 10 or more isolates.b“Other Aspergillus spp.” includes the following species with less than 10 isolates: Aspergillus ochraceus (5), Aspergillus ustus (4), Aspergillus niveus (3), Aspergillus sclerotiorum (2), Aspergillus glaucus (2), and Aspergillus spp. (2).c“Other hyaline fungi” includes the following species with less than 10 isolates: Scopulariopsis brevicaulis (6), Trichoderma spp. (5), Phialemonium curvatum (5), Hormographiella aspergillata (3), Fusarium equiseti (2), Fusarium reticulatum (2), Hormographiella verticillata (1), Chrysonilia sitophila (1), and Beauveria bassiana (1).d“Other black fungi” includes the following species with less than 10 isolates: Exophiala dermatitidis (4), Exophiala jeanselmei (3), Aureobasidium pullulans (2), Hortae werneckii (1), Cladosporium spp. (1), Cladophialophora bantiana (1), Scytalidium hyalinum (1), and Lecythophora hoffmannii (1).e“Other Mucorales species” includes the following species with less than 10 isolates: Cunninghamella bertholletiae (6), Rhizopus microsporus (2), Rhizomucor spp. (2), and Saksenaea vasiformis (1).Micafungin exhibited a potent activity in vitro against Candida spp. That activity was somehow better than the in vitro activity of caspofungin and similar to that of anidulafungin. AFST-EUCAST has not yet proposed breakpoints to read the susceptibility testing of echinocandins. CLSI and others have published that Candida isolates exhibiting MICs of echinocandins of >2 μg/ml can be considered nonsusceptible in vitro (14). Following that criterion, only one isolate out of 20 Candida guilliermondii clinical isolates tested (1/20; 5%) had a MIC of micafungin of >2 μg/ml. That isolate exhibited cross-resistance in vitro to both caspofungin and anidulafungin.A total of 15.3% of the Candida isolates analyzed (160/1,043; 61 Candida glabrata, 53 Candida krusei, 25 Candida albicans, 8 Candida tropicalis, 7 C. guilliermondii, 3 Candida parapsilosis, 1 Candida lusitaniae, 1 Candida norvegensis, and 1 Candida rugosa) were resistant in vitro to fluconazole according to AFST-EUCAST criteria (fluconazole MIC, >4 μg/ml) (19). In addition, 7.9% of the isolates (82/1,043; 36 C. glabrata, 18 C. krusei, 16 C. albicans, 7 C. tropicalis, 4 C. guilliermondii, and 1 C. parapsilosis) were also resistant in vitro to voriconazole (MIC, >0.12 μg/ml by EUCAST criteria) (20). All azole-resistant Candida strains exhibited low MICs of micafungin and other echinocandins.Micafungin and the other echinocandins were inactive in vitro against Dipodascus capitatus and Galactomyces geotrichum and against species of all genera belonging to Basidiomycota, such as Cryptococcus, Trichosporon, and Rhodotorula. The echinocandins were active in vitro against some non-Candida Ascomycota species, such as Saccharomyces cerevisiae and Yarrowia lipolytica.Regarding Aspergillus spp., micafungin exhibited a good activity in vitro against most of the Aspergillus isolates. Notably, the echinocandin seemed to be inactive in vitro against Aspergillus flavus and a number of Aspergillus nidulans clinical isolates as others have reported before (1, 9). Micafungin and the other echinocandins were active in vitro (MEC, ≤ 2μg/ml) against 11 strains of Aspergillus spp. (1.6%; 11/705; 7 Aspergillus niger, 2 Aspergillus fumigatus, and 2 Aspergillus ustus) that had MICs of itraconazole of ≥8 μg/ml.Micafungin also inhibited in vitro some other species of hyaline fungi, such as Penicillium spp., Paecilomyces variotii, and Acremonium spp. In addition, it exhibited activity against some isolates of black fungi as the echinocandin had low MEC values (MEC, ≤ 2μg/ml) for 32 out of 36 (88.9%) isolates of Scedosporium apiospermum, 7/11 (63.6%) isolates of Alternaria alternata, and for most of the Exophiala strains tested.On the contrary, micafungin and the other two echinocandins were inactive in vitro against some species of molds, such Fusarium spp., Paecilomyces lilacinus, Scopulariopsis spp., and Trichoderma spp. which are characterized by their resistance to other antifungal families. The echinocandins were inactive against Mucorales species as well.These results of in vitro activity match published data by other authors for Candida and Aspergillus species (5, 6, 8, 9, 21), including the good activities of the three echinocandins against azole-resistant isolates (15, 16). There again, this study collects susceptibility data on species of fungal pathogens that have not been published before, such as non-Candida Ascomycota yeasts and some species of hyaline and black molds. Notably, micafungin showed activity in vitro against most of the strains of S. apiospermum tested. Zeng et al. reported that that species is resistant in vitro to echinocandins when the MIC is defined as total inhibition of growth (22). However, as a criterion of the MIC definition, we use the lowest drug concentration resulting in aberrant hyphal growth by examination with an inverted microscope (MEC). Following that, micafungin exhibited activity in vitro against 89% of the clinical isolates analyzed.In summary, micafungin is a broad-spectrum antifungal agent with a good profile of activity in vitro which is comparable to those of caspofungin and anidulafungin. It exhibits activity against Candida and Aspergillus spp. except for A. flavus. The echinocandin also inhibits Candida isolates with resistance to both fluconazole and voriconazole and Aspergillus strains with resistance in vitro to itraconazole. It must be noted that micafungin and the other echinocandins are inactive in vitro against Basidiomycota spp., Mucorales spp., and some species of multiresistant fungi, such as Fusarium spp.  相似文献   

15.
A phylogenetic analysis was performed for 34 Aspergillus strains belonging to section Nigri. Molecular methods allowed for the correct classification into three different clades (A. niger, A. tubingensis, and A. foetidus). Correlation with in vitro itraconazole susceptibility distinguished the following three profiles: susceptible, resistant, and showing a paradoxical effect. A number of different species whose morphological features resemble those of A. niger showed unusual MICs to itraconazole that have never been described for the Aspergillus genus.Black aspergilli are widely distributed in nature (16); they are common food spoilers but are also well used for industrial purposes (15). Among Aspergillus species of the Nigri group, A. niger constitutes the most frequent etiological agent of otomycosis (13) and is considered the third cause of pulmonary aspergillosis (10). Nevertheless, the clinical implications of other species are rarely reported, and they are generally identified as A. niger (14, 22).Clinically, identification of unknown Aspergillus clinical isolates to the species level may be important given that different species have dissimilar susceptibilities to antifungal drugs. Thus, the knowledge of the species identity may influence the choice of appropriate antifungal therapy (2, 4). Furthermore, since the antifungal susceptibility patterns for most of the species within section Nigri have been poorly investigated, their identification and antifungal susceptibility profiles appear to be of clinical interest for further research.Black aspergilli belong to one of the most difficult groups concerning classification and identification (18), and so a number of different techniques have been developed in order to solve this issue. Among them, molecular tools are the gold standard (1, 18), as the sequencing of the β-tubulin or calmodulin gene is suitable, and enough, to discriminate between species within section Nigri (3, 18, 21).Thirty-four Aspergillus section Nigri strains belonging to the Mold Collection of the Centro Nacional de Microbiologia and collected since 2004 were analyzed. Thirty-three strains were independent clinical isolates, and 1 had an environmental origin. All strains were identified as A. niger using conventional methods of morphology at the macroscopic as well as microscopic levels (9). Species identification analysis was addressed using sequences of the β-tubulin gene from all the strains included in this study together with the sequences of different Aspergillus section Nigri type strains and others that were available at GenBank as follows: A. tubingensis AY820007T, AY820009, and AY585527; A. foetidus AY585533T, AY585534, and DQ768454; and A. niger FJ629288T, EF422213, and AY585537. Partial sequences of the β-tubulin gene were amplified using the primer set βtubAniger1 and βtubANiger2 (11) and were carried out according to standard PCR guidelines (Applied Biosystems). Sequences were assembled and edited using the SeqMan II and EditSeq software packages (Lasergene 8.0; DNAStar, Inc., Madison, WI).All phylogenetic analyses were conducted with InfoQuest FP software, version 4.50 (Bio-Rad). The methodology used was maximum parsimony clustering. Phylogram stability was assessed by using parsimony bootstrapping with 2,000 simulations and by using the Aspergillus clavatus AY214441T sequence as the out-group.The phylogenetic tree grouped the 34 clinical isolates into three different clades consisting of 13 A. niger isolates, 18 A. tubingensis isolates, and 3 A. foetidus isolates (Fig. (Fig.1).1). Table Table11 shows β-tubulin gene identification, as well as the origin and susceptibility profiles of the isolates.Open in a separate windowFIG. 1.Phylogenetic tree using maximum parsimony phylogenetic analysis and 2,000 bootstrap simulations based on β-tubulin gene sequences from all the Aspergillus section Nigri strains included in the study. Percentages indicate the bootstrap support for each group of sequences. (T), type strain.

TABLE 1.

Source, molecular identification, MICs, and MECs for species of Aspergillus section Nigria
IsolateSourceMolecular identification (β-tubulin gene)MIC (mg/liter)b
MEC (mg/liter)c
AMBITCVCZRVCPOSTRBCASMICA
Isolates of Aspergillus section Nigri showing low ITC MICs
    CM-3236RespiratoryA. niger0.190.50.51.00.121.00.250.03
    CM-3257RespiratoryA. niger0.251.01.01.0.0.251.00.250.03
    CM-3506RespiratoryA. niger0.190.50.751.00.120.310.150.03
    CM-3507RespiratoryA. tubingensis0.190.51.01.50.150.620.060.03
    CM-3585EnvironmentalA. tubingensis0.190.51.01.670.120.420.370.03
    CM-3586CatheterA. niger0.250.51.02.00.120.121.00.03
    CM-3636RespiratoryA. niger0.250.50.51.00.191.00.250.03
    CM-3641RespiratoryA. niger0.250.51.01.00.1250.250.50.03
    CM-3672CutaneousA. niger0.120.51.01.50.190.070.150.03
    CM-4004UnknownA. niger0.251.01.01.670.250.130.100.03
    CM-4213RespiratoryA. niger0.330.140.330.580.030.220.390.03
    CM-4264Blood cultureA. tubingensis0.120.51.01.50.120.500.030.03
    CM-4296RespiratoryA. tubingensis0.120.751.01.50.190.620.250.03
    CM-4316RespiratoryA. niger0.190.50.51.00.1250.50.250.03
    CM-5094RespiratoryA. tubingensis0.120.50.752.00.060.620.190.03
    CM-5095RespiratoryA. niger0.190.50.751.50.120.620.250.03
GM for group0.200.560.821.310.140.500.280.03
Isolates of Aspergillus section Nigri showing much higher ITC MICs
    CM-3123RespiratoryA. tubingensis0.25111.672.670.251.170.250.03
    CM-3810RespiratoryA. tubingensis0.254.02.02.00.121.00.50.03
    CM-4003UnknownA. tubingensis0.12162.04.00.251.00.250.03
    CM-4005UnknownA. tubingensis0.12162.04.00.50.250.50.03
    CM-4688RespiratoryA. tubingensis0.213.672.03.330.251.500.180.03
    CM-5264RespiratoryA. foetidus0.12162.08.00.50.50.060.03
GM for group0.1811.111.954.00.310.900.290.03
Isolates of Aspergillus section Nigri showing paradoxical effect against ITC
    CM-3125RespiratoryA. tubingensis0.120.511.670.120.50.050.03
    CM-3177RespiratoryA. tubingensis0.16123.330.250.670.050.03
    CM-3551RespiratoryA. niger0.54.75120.120.250.030.03
    CM-3654Blood cultureA. tubingensis0.19122.500.250.630.140.03
    CM-4000UnknownA. tubingensis0.161220.250.330.100.03
    CM-4001UnknownA. tubingensis0.1911.752.500.250.560.110.03
    CM-4002UnknownA. foetidus0.25122.670.120.330.140.03
    CM-4262OphthalmicA. niger0.251220.250.290.130.03
    CM-4352RespiratoryA. tubingensis0.2810.8820.250.310.150.03
    CM-4897Blood cultureA. tubingensis0.161220.250.420.100.03
    CM-4899RespiratoryA. tubingensis0.16122.670.250.330.100.05
    CM-4995ProsthesisA. foetidus0.211220.160.330.140.03
GM for group0.21.31.722.280.210.410.100.03
Open in a separate windowaGM, geometric means of MICs and MECs for the strains within each group.bMIC geometric mean of amphotericin B (AMB), itraconazole (ITC), voriconazole (VCZ), ravuconazole (RVC), posaconazole (POS), and terbinafine (TRB).cMEC geometric mean of caspofungin (CAS) and micafungin (MICA).Antifungal susceptibility testing (AST) was performed following the EUCAST Definitive Document E.DEF 9.1 method for the determination of broth dilution MICs of antifungal agents for conidium-forming molds (17). Antifungal ranges used in the microdilution assays have been described previously (2). Endpoints were determined at 48 h. The endpoint for MEC determination was the minimal antifungal concentration that produced morphological alterations of hyphal growth at 48 h. The paradoxical effect to itraconazole was defined as an increase in growth occurring at least 2 drug dilutions above the MIC. AST was repeated at least twice on different days.Three different antifungal patterns were clearly distinguishable based on the itraconazole MIC values (Table (Table1):1): low and high itraconazole MICs and a third group (12 strains) showing an uncommon paradoxical effect of this antifungal (5). Either those strains classified as paradoxical strains or those showing much higher itraconazole MICs also had higher MIC values to voriconazole and ravuconazole.Posaconazole showed better activity in vitro. Moreover, all strains were susceptible to the rest of the following antifungals tested: amphotericin B, terbinafine, and echinocandins.In summary, A. niger MICs for itraconazole, voriconazole, and ravuconazole were slightly higher than A. fumigatus MICs and even more so for A. tubingensis and A. foetidus MICs. Identification of clinical isolates belonging to Aspergillus section Nigri and involved in proven or probable infections should be to the species level because it is the only way to monitor the development of secondary resistances of these molds (7, 8).The paradoxical effect or “Eagle effect” (12) has been previously described for yeasts or A. fumigatus but always in relation to echinocandins (5, 6, 19, 20). This is the first report showing the paradoxical effect of azole drugs against Aspergillus spp. The link between the paradoxical effect against itraconazole and a molecular mechanism responsible for it is yet to be determined, as is the clinical impact of those findings. Therefore, further studies including experimental models of aspergillosis to address any in vitro/in vivo correlations are warranted.  相似文献   

16.
Antimicrobial susceptibilities were determined for 1,586 isolates of Stenotrophomonas maltophilia from globally diverse medical centers using the Clinical Laboratory Standards Institute broth microdilution method. The combination trimethoprim-sulfamethoxazole (96.0% of isolates susceptible at ≤2 μg/ml trimethoprim and 38 μg/ml sulfamethoxazole) and tigecycline (95.5% of isolates sussceptible at ≤2 μg/ml) were the only antimicrobials tested with >94% susceptibility in all regions. Susceptibility rates for other commonly used were lower than expected and varied geographically. This in vitro data supports tigecycline as a potential candidate for clinical investigations into S. maltophilia infections.Stenotrophomonas maltophilia is a Gram-negative bacillus, inherently multidrug resistant (MDR) and frequently recovered from environmental sources. It has been associated with severe nosocomially acquired bacteremia and pneumonia, usually among immunocompromised patients, as well as meningitis, endocarditis, and urinary tract, skin/soft tissue, and ocular infections. S. maltophilia infections are associated with high morbidity and mortality, with estimated crude mortality rates ranging from 20 to 70% and with the risk of mortality highest among patients receiving inappropriate initial antimicrobial therapy (5). Treatment of S. maltophilia infections represents a significant challenge because of the organism''s high levels of intrinsic resistance to many antimicrobial agents, difficulties in susceptibility testing, the development of resistance during therapy, and the paucity of clinical trials to determine optimal therapy (8, 12).The combination trimethoprim-sulfamethoxazole (TMP/SMX) is the recognized antimicrobial of choice for the treatment of infections caused by S. maltophilia with ceftazidime, ticarcillin-clavulanate, minocycline, tigecycline, fluoroquinolones, and the polymyxins being described as alternative therapies. It is important to note that all recommended therapy options have been based on in vitro studies and anecdotal experience rather than appropriately structured clinical trials (11). Resistance to TMP/SMX has been described and varies geographically, being shown by as many as 10% of isolates in Europe (7). In addition, allergic reactions to the combination TMP/SMX are common and can be severe, which further compromises its application (1). Clearly, therapeutic alternatives are needed to treat infections caused by S. maltophilia.Tigecycline is a 9-t-butylglycylamido derivative of minocycline and is the first glycylcycline licensed for clinical use. Tigecycline binds to the 30S ribosomal subunit, resulting in inhibition of protein synthesis (13). It exhibits a wide range of activity against Gram-positive and -negative organisms, including MDR strains. Tigecycline is approved by the United States Food and Drug Administration (USFDA) for the treatment of complicated skin and skin structure infections (cSSSI), intra-abdominal infections, and, more recently, community-acquired bacterial pneumonia. Tigecycline has demonstrated good in vitro activity against S. maltophilia in several studies (6, 9, 14). The aim of this study was to assess antimicrobial resistance in S. maltophilia against commonly used agents by using the largest and most geographically diverse collection of contemporary isolates available, with the rationale being the paucity of such information in the face of a clear need for clinical and research options.From January 2003 to December 2008, a total of 1,586 unique clinical S. maltophilia strains were recovered and identified from 119 medical centers located across Asia and the Pacific (Asian-Pacific), Europe, Latin America, and North America. Bacterial identification was confirmed by the central monitoring site (JMI Laboratories, North Liberty, IA) using standard algorithms (microscopy, culture characteristics, and oxidase reaction) followed by an automated system (Vitek 2; bioMerieux, Hazelwood, MO). MIC values were determined for all isolates based on the Clinical Laboratory Standards Institute (CLSI) broth microdilution method using commercially prepared and validated panels (TREK Diagnostic Systems, Cleveland, OH) in fresh cation-adjusted Mueller-Hinton broth (2). Tigecycline breakpoints established by the USFDA for Enterobacteriaceae (≤2 μg/ml for susceptibility and ≥8 μg/ml for resistance) as well as the polymyxin B breakpoints established by the CLSI for P. aeruginosa (≤2 μg/ml for susceptibility and ≥8 μg/ml for resistance), were applied for comparison only (Tygacil; Wyeth Pharmaceuticals, Philadelphia, PA). CLSI quality control ranges and interpretive criteria were used for comparator compounds (3).Clinical sites of infection for S. maltophilia were primarily bloodstream (51%) and respiratory tract (37%). Tigecycline activities were similar across the four geographic regions (94.5 to 96.5% of isolates inhibited at ≤2 μg/ml) and were most similar to those of TMP/SMX (90.8 to 98.9% of isolates susceptible) (Tables (Tables11 and and2).2). When tested against S. maltophilia isolates from North America and Europe, TMP/SMX was the most active compound (MIC50, ≤0.5 μg/ml and MIC90, 1 μg/ml; 97.6 to 98.9% of isolates susceptible), followed by tigecycline (MIC50, 1 μg/ml, and MIC90, 2 μg/ml; 94.5 to 95.3% o isolates susceptible) and levofloxacin (MIC50, 1 μg/ml, and MIC90, 4 μg/ml; 82.5 to 83.7% of isolates susceptible) (Table (Table2).2). Tigecycline was the most active compound tested against S. maltophilia isolates from the Asian-Pacific and Latin American regions (MIC50, 0.5 μg/ml, and MIC90, 2 μg/ml; 96.1 to 96.5% of isolates susceptible), followed by TMP/SMX (MIC50, ≤0.5 μg/ml, and MIC90, 1 μg/ml; 90.8 to 95.5% of isolates susceptible) (Table (Table2).2). Levofloxacin exhibited good in vitro activity against S. maltophilia isolates from Latin America (91.3% susceptible), but its activity was more restricted when tested against isolates from other geographic regions (78.0 to 83.7% of isolates susceptible) (Table (Table2).2). In general, ceftazidime (32.6 to 51.0% of isolates susceptible), ticarcillin-clavulanate (27.0 to 46.1% of isolates susceptible), and polymyxin B (33.4 to 76.4% of isolates susceptible) showed the most limited in vitro activities against S. maltophilia.

TABLE 1.

Regional MIC distributions for tigecycline tested against 1,586 S. maltophilia strains, stratified by geographic region
Region (no. of strains tested)Cumulative % inhibited at tigecycline MIC (μg/ml) of:
≤0.120.250.512a4>4
North America (491)2.216.549.779.894.598.4100.0
Europe (447)1.813.748.183.595.399.3100.0
Asian-Pacific (359)1.412.557.987.596.199.2100.0
Latin America (289)1.715.252.387.596.5100.0
All regions (1,586)1.814.651.684.095.599.1100.0
Open in a separate windowaSusceptibility breakpoint established by the CLSI for Enterobacteriaceae (3).

TABLE 2.

Antimicrobial activity of tigecycline and comparator agents tested against S. maltophilia isolates from four geographic regions
Region (no. of strains tested) and antimicrobial agentMIC (μg/ml)a
% of isolates
50%90%SusceptibleResistant
North America (491)
    Tigecycline1294.5b1.6b
    Ceftazidime8>1651.034.9
    Levofloxacin1482.58.4
    Polymyxin B≤1>473.2c17.4c
    Ticarcillin-clavulanate3212846.117.6
    TMP/SMXd≤0.5197.62.4
Europe (447)
    Tigecycline1295.3b0.7b
    Ceftazidime16>1645.243.6
    Levofloxacin1483.78.5
    Polymyxin B≤1>472.6c16.2c
    Ticarcillin-clavulanate32>12842.716.2
    TMP/SMX≤0.5198.91.1
Asian-Pacific (359)
    Tigecycline0.5296.1b0.8b
    Ceftazidime>16>1632.653.5
    Levofloxacin1>478.011.7
    Polymyxin B>4>433.4c57.7c
    Ticarcillin-clavulanate64>12827.035.1
    TMP/SMX≤0.5190.89.2
Latin America (289)
    Tigecycline0.5296.5b0.0b
    Ceftazidime16>1648.838.4
    Levofloxacin1291.33.8
    Polymyxin B≤1>476.4c14.9c
    Ticarcillin-clavulanate3212836.722.5
    TMP/SMX≤0.5195.54.5
All regions (1,586)
    Tigecycline0.5295.5b0.9b
    Ceftazidime16>164.842.2
    Levofloxacin1483.483
    Polymyxin B≤1>464.6c25.7c
    Ticarcillin-clavulanate32>12839.124.2
    TMP/SMX≤0.5196.04.0
Open in a separate windowa50% and 90%, MIC50 and MIC90, respectively.bTigecycline breakpoints established by the USFDA (Tygacil; Wyeth Pharmaceuticals, Philadelphia, PA) for Enterobacteriaceae (≤2 μg/ml for susceptibility and ≥8 μg/ml for resistance) were applied for comparison only.cPolymyxin B breakpoints established by the CLSI (3) (Tygacil; Wyeth Pharmaceuticals, Philadelphia, PA) for P. aeruginosa (≤2 μg/ml for susceptibility and ≥8 μg/ml for resistance) were applied for comparison only.dTMP/SMX, trimethoprim-sulfamethoxazole.Tigecycline exhibited similar potencies across all geographic regions, and its antimicrobial activity was similar to that of TMP/SMX. Overall, tigecycline showed a greater potency against S. maltophilia than levofloxacin, ceftazidime, and ticarcillin-clavulanate. Tigecycline and TMP/SMX were the only antimicrobial agents tested with susceptibility rates of >90% in all regions and overall. Prevalence of resistance to alternative therapies varied geographically and was higher than expected or previously reported for these antimicrobials in some geographic regions. There is some evidence to suggest that resistance to alternative drugs could be increasing. Ticarcillin-clavulanate susceptibility was reported as 59.1% in Brazil in 70 clinical isolates collected between 2000 and 2002 (10), compared to our data which show susceptibility at 39.1% for ticarcillin-clavulanate in several Latin American nations, including Brazil. This data highlights the need for continued antimicrobial resistance surveillance at the local level, especially for these alternative agents.Few treatment options are available to treat S. maltophilia infections, and this study demonstrates that antimicrobial resistance to alternate antimicrobial agents is higher than projected and geographically varied. Infections caused by S. maltophilia are life threatening and have a high mortality, and the lack of evidence-based therapeutic options often forces clinicians to make difficult decisions regarding antimicrobial therapy. The role of tigecycline in the treatment of S. maltophilia infections warrants further investigation due to its high in vitro activity and potency. Synergies between tigecycline and TMP/SMX and also amikacin have been reported, and hence combination therapy would be a potential approach for clinical investigations and experimental therapy trials (4).  相似文献   

17.
Planktonic and sessile susceptibilities to micafungin were determined for 30 clinical isolates of Candida albicans obtained from blood or other sterile sites. Planktonic and sessile MIC90s for micafungin were 0.125 and 1.0 μg/ml, respectively.Candida albicans device-related infections are associated with growth of organisms in a biofilm state (3, 6). Device removal is often considered necessary for cure (10), since antimicrobial agents have been considered to have poor activity against microbial biofilms. If, however, antimicrobial agents were active against microbial biofilms, device removal might be avoidable.Cell walls are integral to C. albicans biofilms; therefore, antifungal agents that target cell wall synthesis may be active against fungal biofilms (1). We previously showed that caspofungin and anidulafungin had MIC90s of 2 and ≤0.03 μg/ml, respectively, against 30 C. albicans isolates in biofilms (7, 12). We also demonstrated that caspofungin was active in vivo in an experimental intravascular catheter infection model (13).Herein, we evaluated the activity of micafungin against planktonic and sessile forms of the 30 clinical isolates of C. albicans against which we had previously studied caspofungin, anidulafungin, amphotericin B deoxycholate, and voriconazole (7, 12). One isolate per patient was included; isolates were included only if ≤3 types of organisms were cultured from the specimen from which C. albicans was isolated. Isolates were from blood cultures (n = 10), peritoneal fluid (n = 6), abscess fluid (n = 5), soft tissue (n = 5), bone (n = 2), pleural fluid (n = 1), and urine (n = 1). C. albicans GDH 2346 was used as a positive control.Planktonic MICs were determined using broth microdilution (5). Isolates were grown on Sabouraud dextrose agar for 24 h at 37°C. C. albicans was titrated to 76.6% transmittance at 530 nm in sterile saline and then diluted 1/1,000 in RPMI. Serial twofold micafungin dilutions ranging from 16 to 0.03 μg/ml were assayed. Drug dilution and titrated organism (100 μl each) were placed into corresponding wells of a 96-well, round-bottomed microtiter plate and incubated at 37°C. Forty-eight hours later, MICs were read using a reading mirror and scored according to CLSI guidelines. The lowest concentration associated with a ≥50% reduction in turbidity compared with that for the positive-control well was reported as the MIC. Planktonic MICs for micafungin ranged from ≤0.03 to 0.25 μg/ml (Table (Table1).1). The MIC50 and MIC90 were 0.125 μg/ml. The GDH 2346 MIC was 0.06 μg/ml.

TABLE 1.

Comparison of planktonic and sessile susceptibilities of 30 C. albicans isolatesa
Antimicrobial susceptibilityNo. of isolates with MIC (μg/ml) of:
≤0.030.060.130.25≤0.50.5124816>163264>256
Planktonic
    Anidulafungin (n = 30)178311
    Caspofungin (n = 30)11613
    Micafungin (n = 30)29163
    Voriconazole (n = 30)264
    Amphotericin B (n = 30)11712
Sessile
    Anidulafungin (n = 30)282
    Caspofungin (n = 29)54265412
    Micafungin (n = 30)1246134
    Voriconazole (n = 28)113311126
    Amphotericin B (n = 29)14771
Open in a separate windowaSusceptibilities of anidulafungin are from reference 7; susceptibilities of caspofungin, voriconazole, and amphotericin B are from reference 12; and susceptibilities of micafungin were determined in this study.Sessile MICs (SMICs) were determined with biofilms formed in 96-well, flat-bottomed microtiter plates, as previously described (12). Organisms were inoculated into 7 ml of yeast nitrogen base medium. After 24 h, they were centrifuged and rinsed twice with phosphate-buffered saline (PBS). After being standardized to 1 × 107 CFU/ml in RPMI, 100 μl of each suspension was placed in the wells of a 96-well, flat-bottomed microtiter plate and incubated at 37°C. Approximately 24 h later, the suspensions were discarded, and the wells were rinsed three times with sterile PBS and filled with 100 μl of micafungin in RPMI. Serial twofold micafungin dilutions ranging from 16 to 0.03 μg/ml were studied. Negative-control wells received 100 μl RPMI alone. Microtiter plates were incubated at 37°C for an additional 48 h. Then, media were discarded and wells rinsed three times with sterile PBS. A mixture (100 μl) of 1:10 menadione (1 mM solution in acetone; Sigma, St. Louis, MO) and 2,3-bis(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide inner salt (1 mg/ml in phosphate-buffer saline; sigma) was then placed into each well. Plates were incubated at 37°C for 2 h. A microtiter plate reader was used to measure each well''s absorbance at 492 nm. The lowest concentration associated with a 50% reduction in absorption compared with the level for the control well was reported as the SMIC. SMICs for micafungin were ≤0.03 to 1.0 μg/ml for the 30 clinical isolates (Table (Table1).1). The SMIC50 and SMIC90 were 0.5 to 1.0 μg/ml, respectively. The GDH 2346 SMIC was 0.5 μg/ml.We showed that micafungin is active against C. albicans biofilms; its activity cannot necessarily be predicted based on the activity of other echinocandins (Table (Table1).1). The seven isolates with caspofungin SMIC values of ≥2 μg/ml had anidulafungin SMIC values of <0.03 μg/ml and micafungin SMIC values of <0.5 μg/ml (7). Overall, anidulafungin was the most potent agent against C. albicans biofilms; the anidulafungin SMIC was previously determined to be ≤0.03 μg/ml for 28/30 isolates (7). However, the remaining two isolates had anidulafungin SMIC values of >16 μg/ml, one having the highest planktonic anidulafungin MIC (2 μg/ml) observed (7). The two isolates with anidulafungin SMICs of >16 μg/ml had caspofungin SMICs of ≤0.25 μg/ml and micafungin SMICs of 0.25 μg/ml (7, 12). Together, these data suggest that there may be a need to determine individual echinocandin SMIC values if results are to be translated to the clinical setting.Choi et al. reported micafungin SMIC values of ≤0.5 μg/ml for 12 C. albicans isolates (4). Cateau et al. recently published a study comparing echinocandin treatments of two strains of C. albicans in biofilms on sections of silicone catheters in microtiter plates (2). Exposure to 2 μg/ml of caspofungin or 5 μg/ml of micafungin for 12 h significantly reduced the metabolic activity of 12-h- and 5-day-old C. albicans biofilms, an effect that was maintained, even 48 h later (2). Finally, Kuhn et al. studied the activity of micafungin against two isolates of C. albicans (including GDH 2346, studied herein) (8). Planktonic MICs were 0.001 μg/ml for both isolates; the SMIC for GDH 2346 was identical to ours, and the SMIC of the second isolate was 0.25 μg/ml (8).Our planktonic MIC findings for micafungin are in accordance with previously published results. A recently published study of 2,869 C. albicans isolates showed that the MIC90s were 0.06, 0.06, and 0.03 μg/ml for anidulafungin, caspofungin, and micafungin, respectively (11). In the same study, the highest MICs for anidulafungin, caspofungin, and micafungin were 2, 0.5, and 1 μg/ml, respectively (11). There were 12 isolates with anidulafungin MICs of 2 μg/ml, which, although considered susceptible based on CLSI breakpoints, is high, given that the modal MIC for this species is 0.3 μg/ml; the 12 isolates had micafungin MICs of 0.5 to 1 μg/ml (the modal MIC of micafungin was 0.015 μg/ml) and caspofungin MICs of 0.12 to 0.25 μg/ml (the modal MIC of caspofungin was 0.03 μg/ml) (11). Isolates with such high echinocandin MICs have been associated with echinocandin treatment failure (9). The highest micafungin MIC noted in our study, however, was only 0.25 μg/ml (n = 3); these three isolates had anidulafungin MICs of ≤0.03 (n = 2) or 0.06 μg/ml and caspofungin MICs of 0.25 (n = 2) or 0.5 μg/ml.Our in vitro studies show that micafungin is active against C. albicans in biofilms.  相似文献   

18.
Anidulafungin Etest and CLSI MICs were compared for 143 Candida sp. isolates to assess essential (within 2 log2 dilutions) and categorical agreements (according to three susceptibility breakpoints). Based on agreement percentages, our data indicated that Etest is not suitable to test anidulafungin against Candida parapsilosis and C. guilliermondii (54.4 to 82.4% essential and categorical agreements) but is more suitable for C. albicans, C. glabrata, C. krusei, and C. tropicalis (87.9 to 100% categorical agreement).The echinocandins are available for intravenous treatment of Candida infections, especially for patients with recent azole exposure (17, 26, 27). The Clinical and Laboratory Standards Institute (CLSI) has established guidelines and an interpretive susceptibility breakpoint (≤2 μg/ml) for testing echinocandins against Candida spp. (11, 12). We evaluated the suitability (essential and categorical agreements) of anidulafungin Etest MICs for 143 Candida sp. bloodstream isolates from the Hospital La Fe, Valencia, Spain. Categorical agreement was evaluated according to CLSI (11), Garcia-Effron et al. (21), and Desnos-Ollivier et al. (13, 14) susceptibility microdilution breakpoints (≤2, ≤0.5, and ≤0.25 μg/ml, respectively).The 143 isolates included caspofungin-resistant (six heterozygous and homozygous C. albicans mutants, one C. krusei isolate), caspofungin-susceptible (Table (Table1),1), and quality control (QC; C. parapsilosis ATCC 22019 and C. krusei ATCC 6258) isolates (8, 15, 23); anidulafungin MICs were within the established QC limits (12).

TABLE 1.

Echinocandin MICs for eight reference isolates of C. albicans and one of C. krusei determined by the CLSI M27-A3 broth microdilution and Etest methodsa
StrainCLSI MICb of:
Etest MICc of AND
CASANDMCA
CAI4R1d40.5 (0.25)0.252
T258112
T2681 (0.12)10.5
NR2d40.12 (0.5)0.251
NR3e81132
NR4d20.250.250.5
T320.250.032≤0.032≤0.032
CAI4≤0.032≤0.032≤0.032≤0.032
CY-11840.51≤0.032
Open in a separate windowaM27-A3 MICs obtained in this study were comparable to those reported in references 8 and 23. The reference isolates included seven caspofungin-resistant laboratory mutants (the first seven isolates listed), one wild-type C. albicans isolate (CA14) (15), and one caspofungin-resistant C. krusei isolate (CY-118).bMICs in parentheses indicate discrepancies between this study and those reported in reference 23. CAS, caspofungin; AND, anidulafungin; MCA, micafungin.cAnidulafungin Etest MICs obtained in this study.dHeterozygous fks mutant (15).eHomozygous fks mutant (15).Anidulafungin (Pfizer, Madrid, Spain) MICs were determined for the 143 isolates (Table (Table2)2) by both the CLSI M27-A3 and Etest methods after 24 h at 35°C (11). Reference microdilution trays containing serial drug dilutions (0.016 to 8 μg/ml) in RPMI 1640 medium (0.2% glucose; Sigma-Aldrich, Madrid, Spain) were inoculated with a 1 × 103- to 5 × 103-CFU/ml inoculum. MICs were the lowest drug dilutions that showed ≥50% inhibition (11). Etest MICs were determined according to the manufacturer''s instructions (AB BIODISK, Solna, Sweden) using RPMI agar (2% glucose), an approximately 1 × 106- to 5 × 106-CFU/ml inoculum, and Etest strips (0.002 to 32 μg/ml). MICs were the lowest drug concentrations at which the border of the elliptical inhibition intercepted the strip scale, ignoring trailing growth.

TABLE 2.

Susceptibilities of 143 isolates of Candida spp. to anidulafungin determined by the CLSI broth microdilution (M27-A3) and Etest methods
Species (no. of values) and methodaMICb rangeMIC50bMIC90b% Essential agreementc
C. albicans (33)
    BMD≤0.016-10.0160.2591
    Etest≤0.016-320.0161
C. parapsilosis (57)
    BMD0.03-40.5473.7
    Etest≤0.016-3228
C. tropicalis (15)
    BMD≤0.016-0.060.160.03100
    Etest≤0.016-0.30.160.03
C. glabrata (13)
    BMD≤0.016-0.250.030.1269.2
    Etest≤0.0160.0160.016
C. krusei (12)
    BMD≤0.016-0.50.030.1275
    Etest≤0.016-80.030.5
C. guilliermondii (9)
    BMD≤0.016-11177.8
    Etest≤0.016-818
Other (4)d
    BMD0.03-0.50.03NDe50
    Etest0.06-20.06ND
Total (143)
    BMD≤0.016-40.125279.7
    Etest≤0.016-320.064
Open in a separate windowaBMD, CLSI M27-A3 broth microdilution MICs (50% inhibition). MICs were determined by both tests after 24 h of incubation.bMICs are given in micrograms per milliliter.cAgreement between BMD and Etest MICs.dIncluding C. famata (three isolates) and C. lusitaniae (one isolate).eND, not determined.Anidulafungin MICs were in essential agreement when the discrepancies between the two methods were within 2 dilutions. Categorical errors were calculated according to each of the three breakpoints as follows: (i) very major errors when the reference MIC indicated resistance while Etest indicated susceptibility and (ii) major errors when the Etest categorized the isolate as resistant and the reference as susceptible. For the correlation between the methods, a linear regression analysis using the least-squares method (Pearson''s correlation coefficient; MS Excel software) was performed by plotting Etest versus reference MICs.Echinocandin resistance in Candida spp. has been associated with high MICs, mutations in the FKS1 gene, and therapeutic failure (2, 4, 8, 13, 14, 20, 25). MICs higher than those for other species are consistently observed for C. parapsilosis and C. guilliermondii (28), along with reduced glucan synthase sensitivity (19) and a lack of killing activity for C. guilliermondii (5, 6, 7). Based on both reproducibility and the ability to discriminate between wild strains and caspofungin-resistant mutants (Table (Table1),1), the CLSI established standard conditions for testing echinocandins against Candida spp. (11, 12); we followed this methodology. The evaluation of a new assay requires both essential and categorical agreements; the latter was accomplished using CLSI (11) and two other nonsusceptible microdilution breakpoints (>2, >0.5, and ≥0.5 μg/ml) (13, 14, 21).Our CLSI MIC data for most species were similar to those previously published (28), as demonstrated by our MIC90s (MICs for 90% of the isolates tested), except for C. albicans. However, most of the Etest MIC90s were higher than the CLSI results in this and another study (28), which impacted both agreements (Table (Table2).2). Although the overall essential agreement was 79.7% (R, 0.82; Fig. Fig.1),1), it was >90% for two of the six species (Table (Table2)2) and similar to prior Etest and CLSI comparisons for caspofungin and C. albicans (91 versus 89%), higher for C. tropicalis (100 versus 88%), and lower for the other four species (69.2 to 77.8% versus 90 to 100%) (1, 30). Caspofungin Etest MICs usually were lower than the reference results for yeasts (1, 9, 30) and Aspergillus spp. (16), but our anidulafungin Etest MICs were mostly higher. Although lack of prior evaluations precluded comparisons, the acceptable essential agreement is ≥90% (10). It is unfortunate that C. parapsilosis and C. guilliermondii were among the species with unsuitably low essential agreement, because little information has been gathered in either efficacy clinical trials or molecular studies (5, 22, 24, 31). High MICs (>0.5 μg/ml) were not observed for the clinical isolates of the other species where the essential agreement was low, and therefore those results did not affect the categorical agreement (Tables (Tables22 and and33).Open in a separate windowFIG. 1.Comparison of anidulafungin broth microdilution and Etest MICs for 143 Candida sp. isolates. Interpretive susceptibility MIC breakpoints (≤2 μg/ml and ≤0.5 μg/ml) are indicated by the horizontal and vertical lines, respectively (11, 21).

TABLE 3.

Categorical agreement between anidulafungin CLSI broth microdilution and Etest MIC pairs (n = 143) of Candida spp.
Species (no. of values) and methodaBreakpointb% of MICs by categoryc
% Errors
% Categorical agreementd
SRMajorVery major
C. albicans (33)
    BMD>21000
    Etest>29733097
    BMD>0.593.96.1
    Etest>0.587.912.16.1093.9
    BMD≥0.593.96.1
    Etest≥0.581.818.212.1087.9
C. parapsilosis (57)
    BMD>284.215.8
    Etest>2861412.21473.8
    BMD>0.55347
    Etest>0.5257536.88.854.4
    BMD≥0.51288
    Etest≥0.512888.88.882.4
C. tropicalis (15)
    BMD>21000
    Etest>2100000100
    BMD>0.51000
    Etest>0.5100000100
    BMD≥0.51000
    Etest≥0.5100000100
C. glabrata (13)
    BMD>21000
    Etest>2100000100
    BMD>0.510000
    Etest>0.5100000100
    BMD>0.510000
    Etest>0.5100000100
C. krusei (12)
    BMD>21000
    Etest>291.78.38.3091.7
    BMD>0.51000
    Etest>0.591.78.38.3091.7
    BMD≥0.591.78.3
    Etest≥0.591.78.38.3091.7
C. guilliermondii (9)
    BMD>21000
    Etest>266.733.333.3066.7
    BMD>0.544.455.6
    Etest>0.511.188.933.3066.7
    BMD≥0.533.366.7
    Etest≥0.511.188.922.2077.8
Other (4)
    BMD>21000
    Etest>2752525075
    BMD>0.51000
    Etest>0.5752525075
    BMD≥0.57525
    Etest≥0.5752500100
Total (143)
    BMD>293.76.3
    Etest>290.29.89.15.685.3
    BMD>0.576.223.8
    Etest>0.560.139.919.63.576.9
    BMD≥0.55842
    Etest≥0.553.146.98.43.588.1
Open in a separate windowaBMD, CLSI M27-A3 broth microdilution MICs (50% inhibition). MICs were determined by both tests after 24 h of incubation.bBreakpoints by microdilution methods: (i) CLSI susceptible MICs of ≤2 μg/ml and nonsusceptible MICs of >2 μg/ml (11), (ii) susceptible MICs of ≤0.5 μg/ml and nonsuceptible MICs of >0.5 μg/ml (encompassed >95% of all clinical C. albicans fks1 mutants) (21), (iii) susceptible MICs of ≤0.25/ml and nonsusceptible MICs of ≥0.5 μg/ml (13, 14).cPercentages of BMD and Etest MICs that were within each of the three breakpoint ranges evaluated. S, susceptible; R, nonsusceptible.dPercentages of BMD and Etest MIC pairs that were in agreement regarding each breakpoint category.The categorical agreement was suitable (87.9 to 100%) for four of the six species evaluated, breakpoint dependent (11, 13, 14, 21) (Table (Table3).3). Again, the lowest percentages were for C. parapsilosis and C. guilliermondii (54.4 to 82.4% according to breakpoint) due to major errors (8.8 to 36.8% false resistance) and very major errors (8.8 to 14% false susceptibility for C. parapsilosis only). The best categorical agreement for C. albicans was according to the CLSI breakpoint (11). The FDA target for major errors is ≤3% and ≤1.5% for very major errors (18). Therefore, Etest could be considered unsuitable for testing of C. parapsilosis and C. guilliermondii with anidulafungin but suitable for the other four species (Table (Table3).3). Categorical agreement was not assessed during prior Etest caspofungin evaluations (1, 9, 30), but the agreement was >99% for echinocandin YeastOne MICs for Candida spp. (29).Etest has detected echinocandin resistance (fks1 gene mutations) among Candida and Aspergillus species (2, 3, 4, 14), but similar MICs were obtained by reference methodology for Candida spp. (2, 4). While these results confirmed the lower susceptibility breakpoint (≤0.5 μg/ml) for micafungin and anidulafungin versus C. albicans (21), it is uncertain if this endpoint is applicable for either C. parapsilosis or C. guilliermondii. The CLSI susceptibility breakpoint (≤2 μg/ml) was based on clinical trial data, global susceptibility surveillance, resistance mechanisms, and pharmacokinetic and pharmacodynamic parameters from model systems (11, 28). The response to therapy has been comparable for Candida species, but few isolates of C. parapsilosis (9 to 10%) and C. guilliermondii (none) were included in anidulafungin clinical trials (24, 31). More information is needed for these species; the response of most C. parapsilosis infections to echinocandin therapy, regardless of the reduced susceptibility of these two species, could be due to their lower virulence.In conclusion, our preliminary data indicated unsuitable percentages of both essential and categorical agreements for C. parapsilosis and C. guilliermondii. To our knowledge, Etest has not been evaluated in multicenter studies to assess its reliability and ability to identify echinocandin resistance. Such studies with large numbers of isolates, including well-documented resistant isolates, are essential before using Etest routinely.  相似文献   

19.
All of the carbapenem-resistant Escherichia coli (CREC) isolates identified in our hospital from 2005 to 2008 (n = 10) were studied. CREC isolates were multidrug resistant, all carried blaKPC-2, and six of them were also extended-spectrum beta-lactamase producers. Pulsed-field gel electrophoresis indicated six genetic clones; within the same clone, similar transferable blaKPC-2-containing plasmids were found whereas plasmids differed between clones. Tn4401 elements were identified in all of these plasmids.Carbapenem resistance in Escherichia coli is usually attributed to the acquisition of β-lactamases such as AmpC (14, 23, 24, 27, 31), metallo β-lactamases (4, 17, 25, 33), or KPC-type carbapenemases (2, 3, 26, 32). In 2005, KPC-2-mediated carbapenem-resistant E. coli (CREC) clinical strains were first identified in our hospital (21).The increasing prevalence of carbapenem-resistant Enterobacteriaceae in Israel (30), along with concerns regarding the emergence of highly epidemic clones, led to the study of carbapenem resistance in E. coli in our hospital. We determined CREC prevalence, elucidated the molecular mechanisms contributing to carbapenem resistance, and explored the molecular epidemiology and plasmids associated with this resistance.All of the CREC isolates identified in our hospital from February 2005 to October 2008 were included in this study. Strains were identified as resistant to at least one carbapenem using the Vitek-2 and agar dilution (MIC of imipenem or meropenem, >4 μg/ml; MIC of ertapenem, >2 μg/ml). Antibiotic susceptibilities were determined by Vitek-2 (bioMérieux Inc., Marcy l''Etoile, France), and MICs of carbapenems were determined by agar dilution according to the Clinical and Laboratory Standards Institute (CLSI) protocols (8). MICs of tigecycline and colistin and MICs of imipenem, meropenem, and ertapenem lower than 0.5 μg/ml were determined by Etest (AB Biodisk, Solna, Sweden). The criteria used for the interpretation of carbapenem MICs were based on the CLSI 2010 guidelines (9). The interpretive criterion used for tigecycline was based on FDA breakpoint values for Enterobacteriaceae that define a MIC of ≤2 as susceptible.β-Lactamases were analyzed by analytical isoelectric focusing (IEF) (16) on crude enzyme preparations from sonicated cell cultures as described elsewhere (21). The following β-lactamases were used as controls: TEM-1, pI = 5.4; TEM-26, pI = 5.6; K1, pI = 6.5; SHV-1, pI = 7.6; P99, pI = 7.8; ACT-1, pI = 9.The genetic relatedness between isolates was determined using pulsed-field gel electrophoresis (PFGE) as previously described (7). DNA macrorestriction patterns were compared according to the Dice similarity index (1.5% tolerance interval) (9a) using GelCompar II version 2.5 (Applied Maths, Kortrijk, Belgium). A PFGE clone was defined as a group of strains showing >85% banding pattern similarity (19).Multilocus sequence typing (MLST) was performed on two representative E. coli clones according to the protocol at the E. coli MLST website (http://www.pasteur.fr/recherche/genopole/PF8/mlst/EColi.html).Epidemiological links and potential contact between patients were analyzed using data on room location, consulting physicians, and other procedures performed during their hospitalization.PCR molecular screening of β-lactamase genes and Tn4401 elements (18) was performed using the primers listed in Table Table1.1. PCR products were sized on an agarose gel and sequenced using an ABI PRISM 3100 genetic analyzer (PE Biosystems). Nucleotide and deduced protein sequences were identified using the BLAST algorithm (www.ncbi.nlm.nih.gov/).

TABLE 1.

Primers used in this study
Screened gene and primer typeSequenceaReference
blaKPC
    ForwardATGTCACTGTATCGCCGTCT5
    ReverseTTTTCAGAGCCTTACTGCCC
blaSHV group
    ForwardTTTATCGGCCYTCACTCAAGG5
    ReverseGCTGCGGGCCGGATAACG
blaTEM group
    ForwardKACAATAACCCTGRTAAATGC5
    ReverseAGTATATATGAGTAAACTTGG
blaCTX-M-2 group
    ForwardATGATGACTCAGAGCATTCG5
    ReverseTTATTGCATCAGAAACCGTG
blaCTX-M-3 group
    ForwardGTTGTTGTTATTTCGTATCTTCC5
    ReverseCGATAAACAAAAACGGAATG
blaCTX-M-9 group
    ForwardGTGACAAAGAGAGTGCAACGG5
    ReverseATGATTCTCGCCGCTGAAGCC
blaCTX-M-25 group
    ForwardCACACGAATTGAATGTTCAG5
    ReverseTCACTCCACATGGTGAGT
blaCMY-1 group
    ForwardCAACAACGACAATCCATCCTGTGThis paper
    ReverseCAACCGGCCAACTGCGCCAGGA
blaCMY-2 group
    ForwardATGAAAAAATCGTTATGCTGCGCTCTGThis paper
    ReverseATTGCAGCTTTTCAAGAATGCGCC
blaOXA-9
    ForwardGCGGACTCGCGCGGCTTTATThis paper
    ReverseGCGAGATCACCAAGGTAGTCGGC
blaOXA-40
    ForwardGCAAATAMAGAATATGTSCC10
    ReverseCTCMACCCARCCRGTCAACC
blaOXA-58
    ForwardCGATCAGAATGTTCAAGCGC28
    ReverseACGATTCTCCCCTCTGCGC
ISKpn6
    ForwardGAAGATGCCAAGGTCAATGC19
    ReverseGGCACGGCAAATGACTA
ISKpn7
    ForwardGCAGGATGATTTCGTGGTCT13
    ReverseAGGAAGTCGGTGAAGCTGAA
tnpA
    ForwardCACCTACACCACGACGAACC19
    ReverseGCGACCGGTCAGTTCCTTCT
tnpR
    ForwardACTGTGACGCATCCAATGAG13
    ReverseACCGAGGGAGAATGGCTACT
Open in a separate windowaK is G or T, M is A or C, R is A or G, S is G or C, and Y is C or T.Plasmids were purified as described previously (12) and transformed into E. coli DH10B by electroporation (Electroporator 2510; Eppendorf, Hamburg, Germany). Transformants were selected on LB agar plates containing 100 μg/ml ampicillin, and selected colonies were screened by PCR for the presence of blaKPC. Plasmid size estimation was performed by digestion of plasmid DNA prepared as described previously (7, 22), followed by S1 nuclease (190 U; Promega, Madison, WI) (1) and PFGE. Electrophoresis was carried out as described previously (7), using the Lambda ladder marker (New England Biolabs, Boston, MA).Comparison of KPC-encoding plasmids was performed using restriction length polymorphism (RFLP) following digestion with the BglII, EcoRV, SmaI, and KpnI endonucleases (New England Biolabs, Boston, MA). Southern analysis was performed as described previously (21), using a radioactively labeled blaKPC-2 probe (892 bp) obtained with blaKPC primers (5).Ten CREC isolates were studied. They originated from various isolation sites of 10 patients with no apparent epidemiological connection. The overall prevalence of carbapenem resistance in E. coli during this study period was 0.063% (10 cases out of 15,918 E. coli isolates). All 10 isolates were multidrug resistant (Table (Table2).2). The MIC50s and MIC90s of imipenem and meropenem were 4 and 8 μg/ml, those of ertapenem were 16 and 32 μg/ml, and those of doripenem were 1 and 4 μg/ml, respectively. All of the isolates were susceptible to tigecycline (MIC50 and MIC90 of 0.19 and 0.75 μg/ml) and to colistin (MIC50 and MIC90 of 0.125 and 0.19 μg/ml) (Table (Table22).

TABLE 2.

Antibiotic susceptibility testing results of clinical CREC strains isolated at the Tel Aviv Sourasky Medical Center from 2005 to 2008 and their transformants
E. coli isolateDate of isolationIsolation sitePFGE clusterβ-Lactamase gene(s)MIC (μg/ml)a
CROCAZFEPATMTZPAMKGENCIPLVXIPMbMEMbETPbDPbTGCcCSTc
1572/2005UrineIIIKPC-2 CTX-M-15 TEM>64>64>64>64>1288<1>4>884820.380.125
157TdKPC-2>64162>64>128<2<1<0.25<0.2521410.190.047
3299/2005BloodIIKPC-2 CTX-M-2 TEM>64>6416>64>12832>16>4>8883240.190.19
329TKPC-216162>64>128<2<1<0.25<0.2521410.1250.047
3399/2005Peritoneal fluidVKPC-2 TEM>6442>64644>16>4>8221610.190.125
339TKPC-28162>6464<2<1<0.25<0.252140.50.190.047
36010/2005UrineIKPC-2 CTX-M-15 OXA-9 TEM>64>64>64>64>128322>4>844810.750.19
360TKPC-2 OXA-98162>6464>642<0.25<0.25111<0.50.1250.094
38610/2005WoundIKPC-2 CTX-M-15 OXA-9 TEM>64>6416>64>128>644>4>811810.750.125
386TKPC-2 OXA-98162>6464>644<0.25<0.2510.251<0.50.190.047
5406/2006Synovial fluidIVKPC-2 TEM>6442>646416>16>4>8441610.190.125
5435/2006Synovial fluidIVKPC-2 CTX-M-15 TEM>641616>64>128>64>16>4>8841620.190.125
543TKPC-216162>64>12816<1<0.25<0.2520.52<0.50.1250.047
5445/2006AbscessIVKPC-2 TEM32162>64>1288>16>4>8441610.380.19
5475/2006BloodIVKPC-2 TEM>6442>646416>16>4>8448<0.50.1250.19
547TKPC-28162>646416<1<0.25<0.2510.50.25<0.50.1250.047
167910/2007Peritoneal fluidVIKPC-2 SHV-12 TEM32>648>64>128<2>16>4>88163280.1250.38
1679TKPC-2 TEM8>644>646416<1<0.25<0.2510.510.50.380.047
DH10B recipient<0.25<0.250.125<1<1<1<1<0.25<0.250.250.250.023<0.50.1250.047
Open in a separate windowaCRO, ceftriaxone; CAZ, ceftazidime; FEP, cefepime; ATM, aztreonam; TZP, piperacillin-tazobactam; AMK, amikacin; GEN, gentamicin; CIP, ciprofloxacin; LVX, levofloxacin; IPM, imipenem; MEM, meropenem; ETP, ertapenem; DP, doripenem; TGC, tigecycline; CST, colistin. Unless otherwise noted, MICs were determined by Vitek-2.bMIC determined by agar dilution; carbapenem MICs lower than 0.5 (except for DP) were determined by Etest.cMIC determined by Etest.dT, transformant.PFGE of the 10 CREC isolates revealed six distinct genetic clones (Fig. (Fig.1):1): four clones from 2005 (21), a new clone consisting of four isolates in 2006, and a different clone in 2007. Isolates belonging to the 2006 clone, although genetically identical, originated from four patients hospitalized in different wards during a 1-month period with no apparent epidemiological relatedness. MLST of CREC isolate 386 from 2005 identified this strain as being of sequence type 471 (ST471) reported before in France (11). CREC isolate 547, which belonged to the 2006 clone (Fig. (Fig.1),1), possessed a novel sequence type, ST39.Open in a separate windowFIG. 1.PFGE of clinical CREC isolates. Shown are DNA restriction patterns and a dendrogram showing the level of similarity between SpeI-restricted patterns of CREC isolates. Isolates with asterisks were described previously (21). The scale indicates the degree of genetic relatedness between the strains. Isolates were placed into six different clusters based on GelCompar Dice algorithm coefficients, which range from 0 to 100%, as illustrated by the scale to the left of each dendrogram. The year of isolation of each isolate is shown at the right.IEF demonstrated the production of more than one β-lactamase by each isolate (data not shown). A β-lactamase with an apparent pI of 6.7 was observed in 9 out of 10 isolates, consistent with the pI of KPC-type carbapenemases. PCR screening for β-lactamase genes, followed by sequencing, revealed the presence of blaKPC-2 in all of the strains. Six of the 10 isolates were also extended-spectrum beta-lactamase (ESBL) producers (Table (Table2).The2).The strains carrying CTX-M enzymes showed higher MICs of ceftazidime and cefepime than the non-ESBL producers (Table (Table22).Transformation experiments were performed with eight CREC isolates (Table (Table2).2). Plasmid DNA derived from an E. coli 1679 transformant showed a plasmid size different from that of the donor, suggesting rearrangements of plasmid DNA within this strain; therefore, it was excluded from further analysis. Plasmid DNA analysis of transformants indicated that each has acquired a single plasmid (Fig. (Fig.2).2). PCR screening results of plasmid DNA confirmed the presence of blaKPC in all of them, while not all β-lactamases were transferred (Table (Table2).2). Acquisition of the blaKPC-2-containing plasmids usually elevated the MICs of cephalosporins, aztreonam, aminoglycosides, and carbapenems, yet none of the transformants presented the same level of carbapenem resistance as the respective donor strain (Table (Table22).Open in a separate windowFIG. 2.PFGE after S1 restriction of donor clinical strains and transformants (A) and Southern blotting using a blaKPC-2 probe (B). Plasmid profiles of six CREC isolates representing genetic clusters I to V and their transformants as determined by S1 nuclease treatment, followed by PFGE (A) and Southern blot analysis using blaKPC-2 probe (B). Lane M, Lambda Ladder PFG Marker (New England Biolabs, Boston, MA); lanes 1 to 12, E. coli clinical isolates (D) and their respective transformants (T).CREC isolates possessed two or three plasmids which differed in size (Fig. (Fig.2A,2A, lanes D). Isolates from the same year and belonging to the same clone carried highly similar-sized plasmids. Southern blot analysis of plasmid DNA from clinical isolates and their transformants showed that each clinical isolate carried a single plasmid encoding blaKPC-2 and that these plasmids varied in size, ranging from ∼45 kb (Fig. (Fig.2A,2A, lanes 4 and 6) to ∼100 kb (carried by E. coli strain 157 isolated in 2005) (lane 2). The plasmid DNA RFLP patterns of seven transformants, obtained by using several endonucleases, revealed different restriction patterns but a shared common region, especially between strains isolated in the same year. Southern analysis of the resulting fragments with a labeled-blaKPC-2 probe revealed the same hybridization signal, suggesting that these plasmids share a large fragment harboring blaKPC-2 (Fig. (Fig.3),3), similar to what we found previously in the 2005 isolates (21). Southern blot analysis following restriction with the SmaI endonuclease, which digests blaKPC-2 at nucleotide 790, led to two hybridization signals, suggesting the presence of a single copy of blaKPC-2 in all of the transferred plasmids (Fig. (Fig.3B3B).Open in a separate windowFIG. 3.Restriction analysis of blaKPC-2-harboring plasmids (A) and Southern blotting using a blaKPC-2 probe (B). Restriction analysis of blaKPC-2-containing plasmids derived from six CREC isolates after EcoRV (A) or SmaI (B) digestion, followed by Southern blotting using a blaKPC-2 probe. The two isolates belonging to cluster I (isolates 360 and 386) display the same restriction pattern; therefore, isolate 386 was chosen to depict the restriction pattern of both isolates. SmaI endonuclease cleaves blaKPC-2 at position 790 and the IstB and ISKpn6 open reading frames at positions 203 and 676, respectively, creating fragments of ∼1 and ∼1.5 kb. Lanes M, 1-kb DNA ladder (New England Biolabs, Boston, MA).PCR screening and sequencing of all Tn4401 elements (18) revealed the presence of tnpA transposase, tnpR, and the insertion sequences ISKpn6 and ISKpn7 in all of the isolates and transformants. Based on the sequence recognition of SmaI, blaKPC-2 should be digested, as well as the two genes surrounding it—IstB (part of ISKpn7) and ISKpn6—in a single site, resulting in two DNA fragments of ∼1 and 1.7 kb. These two fragments were visualized by Southern hybridization (Fig. (Fig.3B),3B), which may indicate that the structure of Tn4401 in the close vicinity surrounding blaKPC-2 in our strains is conserved.While carbapenem resistance in Enterobacteriaceae is increasing worldwide, CREC isolates are still rare. However, carbapenem resistance in E. coli is considered to be a great public health threat due to its potential to spread in hospital and community settings (29). This is the first study focusing on the molecular epidemiology and nature of carbapenem resistance in a collection of E. coli isolates within a hospital setting. This paper presents an extension of our previous study in which we first described CREC isolates residing outside the United States (21).Isolates showed a multidrug resistance phenotype, like all KPC-producing Enterobacteriaceae; however, they possessed lower carbapenem MICs (4- to 8-fold lower) compared to the MICs of carbapenem-resistant Klebsiella pneumoniae ST258 (12). Genotyping of the isolates revealed that resistance to carbapenems in E. coli from 2005 to 2008 was not clonally related, except for four cases in 2006 that were genetically identical, but epidemiological data did not prove an apparent linkage among them. Sixty percent of the KPC-2-producing strains were also ESBL producers but apparently belonged to clones different from those described before (7).Carbapenem resistance in E. coli during the years studied was rendered by a KPC-2 carbapenemase encoded on various-sized plasmids, which differed between clones but had regions in common. This is the first report describing the presence of Tn4401 elements in the vicinity of blaKPC-2 in E. coli previously described (18). The exact source of the blaKPC-2 gene from E. coli identified in our hospital is still uncertain. Originally, KPC-2 was detected from Enterobacter cloacae in our hospital in 2004 (6, 15), suggesting that they may have acted as a reservoir for the blaKPC-2 gene.In contrast to epidemic K. pneumoniae clone ST258 (20), carbapenem-resistant E. coli clones did not spread significantly during the last 4 years since their emergence in our hospital or worldwide. However, the potential transfer of blaKPC-2 genes into highly fit, rapidly spreading E. coli strains is disturbing. Strict infection control policies, together with joint efforts, will aid in limiting the further dissemination of blaKPC into E. coli, the most common clinical pathogen.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号