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 500 毫秒
1.
Of the 9 vancomycin-resistant Staphylococcus aureus (VRSA) cases reported to date in the literature, 7 occurred in Michigan. In 5 of the 7 Michigan VRSA cases, an Inc18-like vanA plasmid was identified in the VRSA isolate and/or an associated vancomycin-resistant Enterococcus (VRE) isolate from the same patient. This plasmid may play a critical role in the emergence of VRSA. We studied the geographical distribution of the plasmid by testing 1,641 VRE isolates from three separate collections by PCR for plasmid-specific genes traA, repR, and vanA. Isolates from one collection (phase 2) were recovered from surveillance cultures collected in 17 hospitals in 13 states. All VRE isolates from 2 Michigan institutions (n = 386) and between 60 and 70 VRE isolates (n = 883) from the other hospitals were tested. Fifteen VRE isolates (3.9%) from Michigan were positive for an Inc18-like vanA plasmid (9 E. faecalis [12.5%], 3 E. faecium [1.0%], 2 E. avium, and 1 E. raffinosus). Six VRE isolates (0.6%) from outside Michigan were positive (3 E. faecalis [2.7%] and 3 E. faecium [0.4%]). Of all E. faecalis isolates tested, 6.0% were positive for the plasmid, compared to 0.6% for E. faecium and 3.0% for other spp. Fourteen of the 15 plasmid-positive isolates from Michigan had the same Tn1546 insertion site location as the VRSA-associated Inc18-like plasmid, whereas 5 of 6 plasmid-positive isolates from outside Michigan differed in this characteristic. Most plasmid-positive E. faecalis isolates demonstrated diverse patterns by PFGE, with the exception of three pairs with indistinguishable patterns, suggesting that the plasmid is mobile in nature. Although VRE isolates with the VRSA-associated Inc18-like vanA plasmid were more common in Michigan, they remain rare. Periodic surveillance of VRE isolates for the plasmid may be useful in predicting the occurrence of VRSA.Currently, vancomycin-resistant Staphylococcus aureus (VRSA) infections are rare. Thus far, nine cases have been documented in the United States (9, 20), and two have been reported in other countries, including Iran (1) and India (22, 27). Despite this rarity, 7 of the 9 U.S. VRSA cases occurred in the metropolitan Detroit, MI, area. All U.S. VRSA isolates demonstrated either unique pulsed-field gel electrophoresis (PFGE) patterns or unique plasmid restriction patterns (20, 33). This suggests that each VRSA isolate, including the 7 from Michigan, acquired resistance independently and was not the result of transmission of a common VRSA strain between patients.VRSA isolates are thought to occur by in vivo transfer of a vanA plasmid from an Enterococcus isolate to an S. aureus isolate. For most of the VRSA cases, a vancomycin-resistant Enterococcus (VRE) isolate was either coisolated from the same body site as the VRSA isolate or was found to colonize the patient at another body site, such as the nares or rectum (Table (Table11 shows a summary of VRE isolates from VRSA patients). There is limited evidence that these VRE isolates are the vanA donors. In previous studies (29, 30, 33), we found that certain VRSA isolates carried vanA on a transposon, and in some cases, a plasmid was identified that was identical to the vanA plasmid or transposon in a VRE isolate from the same patient. In VRSA cases 3, 4, and 5, the VRE and VRSA isolates shared the same vanA plasmid. For all other VRSA cases, the VRSA isolates carried the same Tn1546-like element as their associated VRE isolate (10, 33). Hence, it was proposed that the VRSA phenotype occurred by conjugative transfer of a vanA plasmid from VRE to S. aureus (29, 33).

TABLE 1.

Inc18-like vanA plasmid characteristics of VRSA and VRSA-associated VRE isolates from nine VRSA cases in the United Statesc
CaseStateSpeciesSite of isolationInc18-like vanA plasmidTn1546 arrangementaTn1546 insertion sitebReference or source
1MIVRSAFoot woundNoWild-type sequence29, 33
E. faecalisFoot woundYesWild-type sequence+10
2PAVRSAFootNoInsertions and deletions4
E. faeciumSuspected contaminationNANA
3NYVRSAUrineNoInsertions and deletions30
E. faeciumRectumNoInsertions and deletionsOur unpublished data
4MIVRSAToe woundYesWild-type sequence+33
E. faecalisRectal swabYesWild-type sequence+33
5MIVRSAAbdominal woundYesWild-type sequence+33
E. faecalisAbdominal woundYesWild-type sequence+33
6MIVRSAWoundNoWild-type sequence33
E. faecalis/E. aviumRectalYesWild-type sequence+33
7MIVRSALeft elbowYesNA33
NANANANA
8MIVRSAFootNoNAOur unpublished data
NANANANA
9MIVRSAFootNoNAOur unpublished data
E. faecalisNANANANA
Open in a separate windowaThe wild-type arrangement of Tn1546 was considered to be the prototype (2).bThe Tn1546 insertion site test result was defined as positive (+) if the junction sites were the same as those of plasmids pWZ909, pWZ7140, or pWZ1668; if not, the test result was defined as negative (−).cNA, not available.Inc18 incompatibility plasmids are a family of broad-host-range conjugative plasmids that occur naturally in Enterococcus and Streptococcus spp. Plasmids pIP501 and pAMβ1 are two well-characterized examples in this group (15). These plasmids carry multiple antimicrobial-resistance genes, including genes that confer resistance to macrolides, lincosamides, and the streptogramin B (MLS) group, and they can be transferred to a wide variety of bacteria, including streptococci (26), lactococci (17), staphylococci (23), and enterococci (14). Specifically, Inc18-like vanA plasmids were identified in VRE isolates from 4 Michigan VRSA patients (cases 1, 4, 5, and 6) and in VRSA isolates from 3 patients (cases 4, 5, and 7) (10, 33). In all VRSA cases where an Enterococcus sp. with an Inc18-like vanA plasmid was found, the isolate was E. faecalis but each isolate demonstrated a different PFGE pattern, indicating that there was not a single enterococcus vanA donor in Michigan. These results suggest that Inc18-like vanA plasmids may be more likely than other vanA plasmids to transfer from an Enterococcus sp. to S. aureus. If VRE isolates with Inc18-like vanA plasmids are more common in Michigan than in other geographic areas, this may at least partially explain why VRSA isolates have occurred primarily in Michigan. In this study, we examined health care-associated VRE isolates from institutions in various geographical locations within the United States for the presence of an Inc18-like vanA plasmid in order to determine the occurrence of Inc18-like vanA plasmids in VRE and to examine the geographical distribution of these Inc18-positive VRE isolates.  相似文献   

2.
Eighteen hundred Neisseria gonorrhoeae isolates collected in Sydney, Australia, in 2007 and 2008 were examined for mosaic penA alleles that mediated cephalosporin resistance, and the genotypes of the isolates were evaluated. In 2008, there were substantial increases in numbers (from 15 to 85) and proportions (from 1.5 to 10.3%) of mosaic-containing gonococci and major shifts in genotypic patterns, with 10 new genotypes representing 74 of the 85 mosaic-containing isolates and genotypes detected between 2001 and 2005 having disappeared. Enhanced surveillance of gonococcal resistance to cephalosporins is necessary.Neisseria gonorrhoeae isolates collected following failed treatment for gonorrhea with oral extended-spectrum cephalosporins (ESC) possessed a mosaic penA gene that was associated with increased MICs of these antibiotics (1, 30). Penicillin-binding protein 2 (PBP2), encoded by penA, is the major target site for ESC in gonococci, and altered PBP2, when accompanied by polymorphisms in mtrR and porB1b, has a demonstrated role in gonococcal resistance to ESC (4, 7, 31). A small number of subtypes of N. gonorrhoeae with a mosaic PBP2 spread widely in Japan (4), but considerable sequence type (ST) heterogeneity in geographically and temporally diverse populations of mosaic PBP2-containing gonococci (mPBP2GC) was recorded following the dissemination of these subtypes in the Asia-Pacific region (20, 26). Small numbers of mPBP2GC with limited ST distribution were found in San Francisco (12), and “cefixime-resistant” gonococci from Taiwan (27) were of the same ST, 835, as mPBP2GC from Hong Kong responsible for treatment failures with ceftibuten (8).We examined gonococci from the Neisseria Reference Laboratory (NRL) in Sydney, Australia (6), isolated in 2007 and 2008 for the presence of a mosaic PBP2 (24) and compared the findings with our earlier data on mPBP2GC (20, 24, 26). MICs of ceftriaxone, penicillin, and ciprofloxacin (17) for these isolates, N. gonorrhoeae multiantigen STs (NG-MASTs) (10), and auxotypes (5, 6) of the isolates were determined. Control strains (22) included the mPBP2GC strain WHO K and a Hong Kong isolate of ST835 (8), kindly provided by Janice Lo.A sequential sample comprising the last 60 NRL isolates from 2007 and the first 60 isolates from 2008 contained no mPBP2GC from 2007 but six in the 2008 strains that were of NG-MASTs 3158 (n = 4) and 1407 (n = 2) (Table (Table1).1). The mPBP2GC were found exclusively among isolates with an extended phenotype defined by chromosomally mediated resistance to penicillin and ciprofloxacin (17), nonsusceptibility to ceftriaxone (defined here by an MIC of 0.016 μg/ml or more), and proline auxotrophy (CMRP/CipR/CefNS/Pro). mPBP2GC from earlier studies (20, 24) and both the control isolates also had this phenotype.

TABLE 1.

Comparison of mosaic PBP2-containing N. gonorrhoeae strains found in Sydney, Australia, in 2007 and 2008 with those detected in 2001 to 2005a
Yr of isolationNo. of isolates with mosaic PBP2/total no. of isolates (%) in:
Total no. (%) of isolates with mosaic PBP2NG-MASTNo. of isolates with indicated STCeftriaxone MIC (μg/ml)
Sequential sampleSelected sample
2001-200513 (NA)13 (NA)83540.06
141430.06
142420.06
245320.06
32610.06
167710.06
20070/60 (0)15 (1.5)15 (1.5)140790.016-0.06
315840.016-0.03
338010.03
350510.06
20086/60 (10)79 (73)85 (10.3)1407350.016-0.06
3149260.016-0.06
3158130.016-0.03
315920.03
316120.03
316820.03
329420.016
295510.03
338110.016
338010.03
Open in a separate windowaData for isolates from 2001 to 2005 are from reference 20. A sequential sample of 60 nonduplicate isolates collected in 2007 and 2008 and a selected sample of isolates with an extended phenotype collected in 2007 and 2008 were examined for the presence of mPBP2GC. STs were determined by NG-MAST genotyping (10). All isolates listed in the table had the Pro[minus] (proline-requiring) auxotype. NA, not applicable.All gonococci of the CMRP/CipR/CefNS/Pro phenotype identified among the N. gonorrhoeae isolates received by the NRL in 2007 (n = 978) and 2008 (n = 835) were then examined for the presence of a mosaic PBP2, and the NG-MASTs were determined. In 2007, there were 64 nonduplicate strains (6.5% of all isolates) with this phenotype, of which 15 (1.5% of all isolates) contained a mosaic PBP2 and were of STs 1407 (n = 9), 3158 (n = 4), 3380 (n = 1), and 3505 (n = 1) (Table (Table1).1). In 2008, a total of 114 gonococci (13.5%), including those examined in the initial sample, had the nominated phenotype, of which 85 nonduplicate strains (10.3%) contained a mosaic PBP2 allele. These strains comprised 10 STs: 1407 (n = 35), 3149 (n = 26), and 3158 (n = 13); 3159, 3161, 3168, and 3294 (n, 2 strains each); and 2955, 3380, and 3381 (n, 1 strain each). The STs of gonococci that had the extended phenotype but lacked a mosaic PBP2 were distinct from the STs of the mPBP2GC.These longitudinal data provided significant insights into the population dynamics of mPBP2GC. Substantial increases in the numbers and diversity of mPBP2GC isolated in Sydney emerged during the study period, consistent with earlier reports regarding the spread of mPBP2GC (4) and quinolone-resistant N. gonorrhoeae (15, 18). Changes in mPBP2GC STs in 2007 and 2008 that saw the loss of ST835 and the emergence of STs 1407, 3149, and 3158 (Table (Table1)1) were also typical of the dynamics of the spread of gonococcal subtypes (6, 9, 13, 15, 23). Rapid increases in resistant gonococci may also occur through their spread within sexual networks (2, 15, 18).Ceftriaxone MICs for mPBP2GC ranged between 0.016 and 0.06 μg/ml (Tables (Tables11 and and2).2). In 2007 and 2008, ceftriaxone MICs for non-mPBP2GC ranged between 0.004 and 0.12 μg/ml (data not shown). Almost all the mPBP2GC in the present study, as those in other studies, were resistant to penicillin and ciprofloxacin (7, 8, 19, 26). The highest ceftriaxone MICs associated with a mosaic PBP2 allele in N. gonorrhoeae were previously reported to occur in the presence of other gene polymorphisms, including some that are as yet undetermined (31). Ceftriaxone MICs may remain relatively unaltered in the presence of a mosaic PBP2 alone (4, 12), and other mosaic or “partial mosaic” PBP2 lesions confer little or no increase in ceftriaxone MICs (14, 25). Additionally, non-mosaic-based penA alterations, e.g., the widely distributed A501V mutation, also confer a wide range of increases in MICs of ceftriaxone (11, 26). Further, the in vitro effects of the mosaic PBP2 allele on susceptibilities to oral ESC and injectable ceftriaxone differ (31). The clinical relevance of increased MICs of ceftriaxone is at present unclear, and better correlates of clinical outcomes and in vitro susceptibility to ESC are therefore urgently required (19, 29). However, these considerations also need to account for the significant differences in ceftriaxone doses in current standard regimens (16) and for confounding effects on treatment outcomes of infections at different anatomical sites (21, 28).

TABLE 2.

Distribution of ceftriaxone MICs for 100 mosaic PBP2-containing N. gonorrhoeae isolates collected in Sydney, Australia, in 2007 and 2008
YrNo. of isolates for which ceftriaxone MIC (μg/ml) was:
Total
0.0160.030.06
200729415
20082061485
Total22708100
Open in a separate windowAltered susceptibility of gonococci to ESC is now recognized as a major concern which calls for enhanced surveillance of gonococcal resistance (3, 19, 28, 30). The findings reported here in regard to the local spread of one important genetic change associated with decreased ceftriaxone susceptibility reinforce these recommendations. However, in vitro detection of the phenomenon remains problematic (19). Molecular systems proposed for the direct detection of the mosaic PBP2 allele in clinical samples (12) are necessarily restricted to this nonexclusive mechanism and thus of limited clinical application. Careful ongoing appraisals of assay specificity and sensitivity would be required because penA mosaicisms are derived from commensal Neisseria spp. and continually change. The finding here that all mPBP2GC were of a particular phenotype is unlikely to be of practical assistance in their detection because of the ubiquity of the various phenotypic markers involved. About 25% of all NRL isolates in 2007 and 2008 were proline auxotrophs (data not shown). Thus, the current interim recommendation (19) that phenotypic examination of isolates for ESC resistance by MIC determination or some other form of validated susceptibility testing, e.g., disc screening, using recommended controls (22) would thus seem to be appropriate at present to ensure high surveillance standards and to enable reliable application of any findings to the optimization of treatment regimens for gonorrhea.  相似文献   

3.
The in vitro activities of eight antifungal drugs against clinical isolates of Fonsecaea pedrosoi (n = 21), Fonsecaea monophora (n = 25), and Fonsecaea nubica (n = 9) were tested. The resulting MIC90s for all strains (n = 55) were as follows, in increasing order: posaconazole, 0.063 μg/ml; itraconazole, 0.125 μg/ml; isavuconazole, 0.25 μg/ml; voriconazole, 0.5 μg/ml; amphotericin B, 2 μg/ml; caspofungin, 2 μg/ml; anidulafungin, 2 μg/ml; and fluconazole, 32 μg/ml.Fonsecaea spp., anamorph members of the order Chaetothyriales (black yeasts and other melanized fungi), are principal agents of human chromoblastomycosis (16), a chronic cutaneous and subcutaneous infection characterized by slowly expanding skin lesions, a granulomatous immune response, and the presence of meristematic melanized muriform fungal cells in tissue scrapings (4). The last characteristic is a crucial diagnostic indicator that tends to be similar irrespective of the fungal pathogen. Chromoblastomycosis occurs worldwide in tropical and subtropical climates. Fonsecaea spp. are recoverable from environmental sources, so the disease is considered to be of traumatic origin (8, 9). The taxonomy of the genus Fonsecaea has been reviewed recently (12), and on the basis of sequence data, the following three species are recognized: Fonsecaea pedrosoi, Fonsecaea monophora, and Fonsecaea nubica. These species are morphologically identical, but their clinical spectra differ slightly: F. pedrosoi and F. nubica appear to be associated strictly with chromoblastomycosis, whereas F. monophora has also been isolated from brain abscesses, cervical lymph nodes, and bile (4, 13, 18).Therapy for chromoblastomycosis is challenging because there is no consensus regarding the treatment of choice. Several treatment options have been applied, but these tend to result in protracted disease, low cure rates, and frequent relapses (5, 9, 10, 16, 18). The therapeutic outcomes are variable and are allegedly dependent on the site of infection, lesion size, the etiological agent, and the patient''s health status (4). The specific identification of the causative pathogen is important for epidemiological reasons. The vast majority of cases of chromoblastomycosis in which the pathogen has been identified are caused by F. pedrosoi; for example, F. pedrosoi was isolated from 94% (66/69 cases) of patients with chromoblastomycosis in Sri Lanka (2) and from 98% (77/78 cases) of patients with culture-positive chromoblastomycosis in Brazil (17).The present study aimed at determining the in vitro susceptibilities of clinical isolates of Fonsecaea spp. to seven marketed antifungal drugs and the experimental 1,2,4-triazole antimycotic isavuconazole (11).Fifty-five Fonsecaea strains were obtained from the Centraalbureau voor Schimmelcultures (Utrecht, The Netherlands) and comprised 21 F. pedrosoi strains, 25 F. monophora strains, and 9 F. nubica strains. Fifty isolates originated from patients with chromoblastomycosis, one isolate was recovered from a patient with a cerebral infection, two isolates were from diseased animals, and two isolates were clinical isolates from unknown sources. Seventeen strains came from southern China, 30 from South and Central America, and 8 from other countries (The Netherlands, Spain, Uruguay, Libya, France, United Kingdom). Strain identities were verified by sequencing the ribosomal internal transcribed spacer (ITS), tubulin (TUB1), and actin (ACT1) regions. In vitro susceptibility was determined as described in CLSI document M38-A2 (6). Briefly, the isolates were cultured on potato dextrose agar (35°C) for up to 7 days, and inocula were prepared by gently scraping the surface of the fungal colonies with a sterile cotton swab moistened with sterile physiological saline containing 0.05% Tween 40. Large particles in the cell suspensions were allowed to settle for 3 to 5 min at room temperature, and then the concentration of spores in the supernatant was adjusted spectrophotometrically (530 nm) to a percent transmission in the range 68 to 71, corresponding to 1.5 × 104 to 4 × 104 CFU/ml, as controlled by quantitative colony counts (6). Antifungal drugs were obtained as reagent-grade powders. The final concentrations of amphotericin B (AMB; Bristol-Myers Squibb, Woerden, The Netherlands), itraconazole (ITR; Janssen Research Foundation, Beerse, Belgium), voriconazole (VOR; Pfizer Central Research, Sandwich, United Kingdom), posaconazole (POS; Schering-Plough, Kenilworth, NJ), and caspofungin (CAS; Merck, Sharp & Dohme, Haarlem, The Netherlands) ranged from 0.016 to 16 μg/ml; the fluconazole (FLU; Pfizer) assay range was 0.063 to 64 μg/ml; and the isavuconazole (ISA; Basilea Pharmaceutica International AG, Basel, Switzerland) and anidulafungin (ANI; Pfizer) assay ranges were 0.008 to 8 μg/ml. After 72 h of incubation at 35°C, MICs and minimum effective concentrations (MECs) were determined visually by comparison of the growth in the wells containing the drug with the drug-free control. The MICs of AMB, ITR, VOR, POS, and ISA were defined as the lowest drug concentration that prevented any discernible growth (100% inhibition), whereas for FLU, the MIC was taken as the lowest concentration supporting ≥50% growth inhibition compared to the growth in the control wells. For CAS and ANI, MECs were determined microscopically as the lowest concentration of drug promoting the growth of small, round, compact hyphae relative to the appearance of the filamentous forms seen in the control wells. Quality control strains Paecilomyces variotii (ATCC 22319), Candida parapsilosis (ATCC 22019), and Candida krusei (ATCC 6258) were included in each assay run.The geometric mean MICs, MIC ranges, MIC50s, and MIC90s for the Fonsecaea isolates are presented in Table Table1.1. For each drug-species pair, the MIC50 and geometric mean MIC values differed by <1 log2 dilution step, indicating that in all cases the MIC50 obtained by inspection reasonably reflected the central tendency of the antifungal susceptibility of the population. All isolates had low MICs (MIC90s ≤ 0.5 μg/ml) for POS, ITR, ISA, and VOR; less active drugs (MIC90s ≥ 2 μg/ml) were AMB, CAS, ANI, and FLU. There were no significant differences in the activities of the surveyed drugs against F. pedrosoi, F. monophora, and F. nubica. The MICs obtained in this study were similar to those obtained in other studies of Fonsecaea isolates (1, 3, 7, 14-16, 21).

TABLE 1.

Geometric mean MICs, MIC ranges, MIC50s, and MIC90s obtained by susceptibility testing of antimycotic agents against Fonsecaea isolates
Strain (no. of strains) and drugMIC (μg/ml)
Geometric meanRange50%90%
All Fonsecaea strains (n = 55)
    Amphotericin B1.0130.5-212
    Fluconazole19.088-641632
    Itraconazole0.0820.031-0.250.0630.125
    Voriconazole0.290.125-10.250.5
    Posaconazole0.0410.016-0.0630.0310.063
    Isavuconazole0.1960.063-10.250.25
    Caspofungin2.151-422
    Anidulafungin3.431-842
Fonsecaea pedrosoi (n = 21)
    Amphotericin B0.9670.5-212
    Fluconazole22.258-323232
    Itraconazole0.08170.031-0.250.0630.125
    Voriconazole0.3360.125-0.50.50.5
    Posaconazole0.04970.031-0.0630.0630.063
    Isavuconazole0.2260.063-0.250.250.25
    Caspofungin2.432-444
    Anidulafungin3.52-888
Fonsecaea monophora (n = 25)
    Amphotericin B1.110.5-212
    Fluconazole19.918-641632
    Itraconazole0.07830.031-0.250.0630.125
    Voriconazole0.2570.125-10.0630.125
    Posaconazole0.03690.016-0.0630.0310.063
    Isavuconazole0.1840.063-10.1250.25
    Caspofungin1.941-422
    Anidulafungin3.781-848
Fonsecaea nubica (n = 9)
    Amphotericin B0.9250.5-212
    Fluconazole18.6616-321632
    Itraconazole0.0990.031-0.250.1250.25
    Voriconazole0.3140.25-0.50.250.5
    Posaconazole0.03620.031-0.0630.0310.063
    Isavuconazole0.170.063-0.50.1250.5
    Caspofungin2.162-424
    Anidulafungin2.512-828
Open in a separate windowTreatment of chromoblastomycosis is difficult. In cases caused by Cladophialophora carrionii and Phialophora verrucosa, patients generally respond well to relatively low doses of most antimycotics. The in vitro susceptibilities of C. carrionii strains to antifungal drugs (20) were similar to those of the Fonsecaea spp. In this study, using unique clinical isolates of Fonsecaea from patients with chromoblastomycosis, we demonstrated differences in the activities of the compounds. ITR has frequently been used to treat chromoblastomycosis attributed to Fonsecaea spp., although elevated ITR MICs have been encountered in sequential isolates during ITR treatment (1).POS is a new oral triazole that is used for the treatment of invasive fungal infections (19), including infections caused by the species associated with chromoblastomycosis (14). In the present study, POS had the lowest MICs among all the drugs examined, although the MIC90s for ITR and ISA were only 1 and 2 log2 dilution steps higher, respectively. The experimental drug ISA possesses potent, broad-spectrum activity against the yeasts and molds implicated in serious mycoses (11). POS, ITR, ISA, and VOR all seem to be potential candidates for use for the treatment of chromoblastomycosis, whereas echinocandins will probably have only a limited role in treatment for this indication due to their relatively high MICs and the lack of oral formulations. However, the in vitro results presented here need to be confirmed in studies with the appropriate animal models of chromoblastomycosis.  相似文献   

4.
We evaluated the in vitro activity of fosfomycin against urinary isolates in a region in Greece that exhibits considerable antimicrobial resistance by evaluating retrospectively relevant susceptibility data retrieved from the microbiological library of the University Hospital of Heraklion, Crete, Greece. We examined 578 urinary isolates. In total, 516 (89.2%) were susceptible to fosfomycin; 415 isolates were gram negative, and 101 isolates were gram positive. Fosfomycin appears to exhibit good levels of in vitro activity against the examined urinary isolates.Urinary tract infections (UTIs) are among the commonest types of bacterial infections (13). The antibiotic treatment for UTIs is associated with important medical and economic implications (12, 17).Antibiotic agents such as beta-lactams, trimethoprim, and cotrimoxazole have been used for the treatment of UTIs (1, 22). However, the emergence of resistant uropathogens led to a shift to fluoroquinolones (18, 21, 26, 27), shorter antibiotic regimens (24, 39), and early switch practices (38). Yet, resistance to fluoroquinolones has been reported (14, 23). Additionally, the emergence of uropathogens, mainly Escherichia coli, exhibiting high rates of resistance due to the production of extended-spectrum beta-lactamases (ESBLs) is worrisome (30, 40).Fosfomycin is an old broad-spectrum bactericidal antibiotic agent that inhibits the synthesis of the bacterial cell. Its pharmacokinetic profile encourages its use for UTIs; the mean peak urinary concentration of an oral single dose of 3 g fosfomycin tromethamine occurs within 4 h, while concentrations sufficient to inhibit the majority of the urinary pathogens are maintained for 1 to 2 days (28). Thus, fosfomycin tromethamine has been approved as an oral single-dose treatment for acute uncomplicated cystitis (34, 39). Data from studies evaluating the role of fosfomycin in infections other than UTIs are also encouraging (8-11). We aimed to evaluate the in vitro activity of fosfomycin against urinary isolates isolated from patients in a 700-bed university hospital in Heraklion, Crete, Greece.The data included in our study were retrieved from the database of the microbiological laboratory of the University Hospital of Heraklion, Crete, Greece. We retrospectively tested fosfomycin susceptibility in all clinical urinary isolates that were collected over a 12-month period (January to December 2008). No specific criteria for the selection of isolates to be subjected to fosfomycin susceptibility testing had been set. Only the first isolate of each species per patient could be included in our study.Quantitative urine cultures were performed by standard techniques using Columbia blood and MacConkey agar plates (bioMérieux, Marcy l''Étoile, France). Plates were incubated for 18 to 24 h at 36°C. The bacterial species identification was performed by using standard biochemical methods, the API system, or the Vitek 2 automated system (bioMérieux, Marcy l''Étoile, France) (36). Antimicrobial susceptibility testing was performed using the disk diffusion method according to the Clinical and Laboratory Standards Institute (CLSI) (5). Due to the lack of acknowledged fosfomycin breakpoints for bacteria other than Escherichia coli and Enterococcus faecalis (2, 5, 33), we used the fosfomycin breakpoints for E. coli proposed by the CLSI for the remaining evaluated gram-negative isolates, a practice that was also followed by authors of similar studies (6). Similarly, we used the CLSI breakpoint regarding E. faecalis for the other evaluated gram-positive bacteria. Identification of ESBL-producing strains was performed by phenotypic testing based on the synergy between clavulanic acid and extended-spectrum cephalosporins (5). Carbapenemase production was detected by the modified Hodge test (29).A total of 578 clinical urinary isolates were included. Two-hundred seven (35.8%) of these 578 isolates originated from adult outpatients, 167 (28.8%) from patients hospitalized in medical wards, 74 (12.8%) from adult patients hospitalized in surgical wards, 17 (2.9%) from intensive care unit adult patients, 9 (1.5%) from adult patients in renal replacement therapy clinics, and 14 (2.4%) from patients from areas other than the above-mentioned hospital units. Ninety (15.5%) of the 578 isolates originated from pediatric patients. Eighty-six (95.5%) of these 90 isolates originated from pediatric inpatients.The 578 tested urinary bacterial isolates represented 456 (78.8%) gram-negative isolates and 122 (21.1%) gram-positive isolates. The 456 gram-negative isolates represented 404 (88.5%) members of the Enterobacteriaceae, 266 (65.8%) of which were Escherichia coli, and 52 (11.5%) of which were gram-negative, nonfermentative bacilli. The 122 tested gram-positive isolates consisted of 74 (60.6%) Enterococcus faecalis isolates, 16 (13.1%) Staphylococcus saprophyticus isolates, 13 (10.6%) Staphylococcus aureus isolates, 12 (9.8%) Enterococcus faecium isolates, 6 (4.9%) Streptococcus agalactiae isolates, and a single (0.8%) Staphylococcus epidermidis isolate. In total, 516 (89.2%) of the 578 tested urinary isolates were found to be susceptible to fosfomycin; 415 were gram negative, and 101 were gram positive. Specific data are presented in Table Table1.1. The above-mentioned group of 516 fosfomycin-susceptible urinary isolates included specific categories of resistant isolates. Specific data are presented in Table Table22.

TABLE 1.

Susceptibility to fosfomycin of gram-negative and gram-positive urinary isolates tested
IsolateTotal no. of isolatesNo. of isolates (%) with:
SusceptibilityIntermediate susceptibility
Gram-negative bacteriaa
    Members of the Enterobacteriaceae
        Escherichia coli266266 (100.0)0 (0)
        Enterobacter species1612 (75.0)2 (12.5)
        Klebsiella pneumoniae6856 (82.3)5 (7.3)
        Proteus mirabilis3130 (96.7)0 (0)
        Other members of the Enterobacteriaceae2114 (66.6)2 (10.0)
    Nonfermentative gram-negative bacilli
        Acinetobacter baumannii111 (9.0)0 (0)
        Pseudomonas aeruginosa4034 (85.0)2 (5.0)
Gram-positive bacteriab
    Staphylococcus aureus1313 (100.0)0 (0)
    MRSAc44 (100.0)0 (0)
    S. saprophyticus1616 (100.0)0 (0)
    Enterococcus faecalis7468 (91.8)1 (1.3)
    Enterococcus faecium120 (0)0 (0)
    Streptococcus agalactiae63 (50.0)0 (0)
Open in a separate windowaThree single urinary isolates, namely, a Salmonella sp., Serratia marcescens, and Stenotrophomonas maltophilia, were also included in the total of 456 tested gram-negative isolates. Both the Salmonella and the Serratia isolates were susceptible to fosfomycin, whereas the Stenotrophomonas maltophilia isolate was resistant to fosfomycin.bA single isolate of Staphylococcus epidermidis was also included in the total of the 122 tested gram-positive urinary isolates. This isolate was susceptible to fosfomycin.cResistant to cefoxitin or oxacillin.

TABLE 2.

Susceptibility to fosfomycin in specific categories of resistant isolates tested
IsolateTotal no. of isolatesNo. of isolates (%) susceptible to fosfomycin
Gram-negative bacteria
    K. pneumoniae nonsusceptible to colistina127 (58.3)
    K. pneumoniae nonsusceptible to carbapenema2515 (60.0)
    Carbapenemase-producing K. pneumoniae2515 (60.0)
    P. aeruginosa nonsusceptible to carbapenema98 (88.8)
    Members of the Enterobacteriaceae nonsusceptible to cefotaxime, ceftriaxone, cefepime, ceftazidime, and aztreonama
        E. coli1414 (100.0)
        K. pneumoniae4030 (75.0)
        Enterobacter species55 (100.0)
    P. aeruginosa1716 (94.1)
    ESBL-producing bacteria
        E. coli1414 (100.0)
        K. pneumoniae1515 (100.0)
        Klebsiella oxytoca10 (0)
Gram-positive bacteria
    MRSA44 (100.0)
    Vancomycin-resistant E. faecium40 (0)
Open in a separate windowa“Nonsusceptible” includes resistant isolates and isolates with intermediate susceptibility.The main finding of our study is that fosfomycin exhibits considerably high antimicrobial activity against urinary clinical isolates with relatively high levels of antimicrobial resistance that were collected recently in a university hospital in Crete, Greece. Specifically, fosfomycin was active against all tested E. coli, S. aureus (including methicillin [meticillin]-resistant S. aureus [MRSA]), and S. saprophyticus isolates. However, the number of the above-mentioned gram-positive isolates was rather limited. Considerable rates of susceptibility to fosfomycin were found for Proteus mirabilis, Klebsiella pneumoniae, Pseudomonas aeruginosa (including the respective carbapenem-resistant isolates), and Enterobacter spp., as well as Enterococcus faecalis and E. faecium.In our study, fosfomycin susceptibility rates followed only those of colistin, imipenem, and meropenem for the majority of the tested gram-negative urinary isolates. Regarding E. coli, rates of susceptibility to fosfomycin were maximal and equaled those to colistin and to carbapenems. Increasing resistance of nosocomial or community-acquired E. coli strains to ampicillin, cotrimoxazole, or fluoroquinolones has been reported previously in various clinical settings (4, 18, 27, 35). In contrast, the reported rates of E. coli resistance to fosfomycin were lower (3, 16, 25). In our study, all the E. coli tested isolates were susceptible to fosfomycin. The emergence of ESBL-producing E. coli strains is also an evolving problem (40). There is some evidence that fosfomycin might be a promising solution for the treatment of such infections (19, 20, 30). It is also suggested that the antimicrobial activity of fosfomycin against ESBL-producing E. coli may be accompanied by an immunomodulating activity (37). Notably, in our study, fosfomycin was active against all the ESBL-producing E. coli and K. pneumoniae isolates.Since fosfomycin has been used extensively in many countries for the treatment of UTIs (7, 15, 32), the extrapolation of our findings should be dealt with cautiously. However, the alarming resistance rates observed in Greece necessitate the evaluation of alternative antibiotic agents (31). Moreover, genetic identification techniques were not implemented in our study. Thus, one may consider that a proportion of the tested urinary isolates might have been of the same clonal origin.In conclusion, our study indicates that fosfomycin is active in vitro against a considerable percentage of urinary isolates, which simultaneously exhibit high rates of antimicrobial drug resistance to the conventionally used antimicrobial agents for the treatment of UTIs. Consequently, fosfomycin may be considered a useful antibiotic agent in our armamentarium for the treatment of UTIs.  相似文献   

5.
One hundred fifty canine and feline Escherichia coli isolates associated with urinary tract infections were screened for the presence of extended-spectrum β-lactamase (ESBL) genes. Out of 60 isolates suspected to be ESBL positive based on antimicrobial susceptibility testing, 11 ESBLs were identified, including one SHV-12 gene, one CTX-M-14 gene, and nine CTX-M-15 genes. This study provides the first report of CTX-M- and SHV-type ESBLs in dogs and cats in the United States.The first detection of an extended-spectrum β-lactamase (ESBL) in an organism from an animal was reported in Japan in 1988 in an Escherichia coli isolate from a laboratory dog (13). Since that time, numerous reports of ESBL-positive isolates from dogs and cats, as well as from other animal species (5), have been made worldwide (4, 8, 11, 17, 25, 26). Only one study has identified the presence of an ESBL in isolates from animals in the United States, i.e., in Salmonella enterica serovar Newport from horses (21). We hypothesized that ESBL genes would be present in urinary E. coli isolates from companion animals in the United States. The purpose of this study was to screen a group of 150 E. coli isolates from canine and feline patients that had been diagnosed with a urinary tract infection (UTI) for the presence of ESBL genes.A convenience sample of 150 E. coli isolates collected from canine and feline patients at the Matthew J. Ryan Veterinary Hospital of the University of Pennsylvania with clinical signs or evidence on routine urinalysis of a UTI between 1 September 2004 and 31 December 2007 was used in this study. Isolates were frozen in Microbank tubes (ProLab Diagnostics, Austin, TX) and stored at −70°C prior to use.MICs were determined using a Negative Combo 31 panel on a MicroScan Walkaway 40 (Dade Behring, Siemens Healthcare Diagnostics, Deerfield, IL). Results were interpreted using Clinical and Laboratory Standards Institute (CLSI) breakpoints (7). Isolates collected from the same individual animal within a 45-day period were considered to be the same strain, and only the first isolate collected was analyzed (subsequent isolates were considered redundant). All isolates with a cefpodoxime MIC of ≥4 μg/ml and a ceftazidime MIC of ≥1 μg/ml were identified as an “ESBL alert” on the MicroScan Walkaway. ESBL confirmatory testing was performed via the Etest method using ceftazidime-ceftazidime-clavulanic acid and cefotaxime-cefotaxime-clavulanic acid strips in accordance with CLSI guidelines (7).Due to the high prevalence of cefoxitin resistance in this population, PCR was performed to detect the presence of a blaAmpC gene, the product of which can mask the effects of clavulanic acid on the ESBL confirmatory test (18, 27). Since the primers used in this study to identify the blaAmpC gene (18) have since been shown to amplify the plasmid-mediated blaCMY gene from Citrobacter freundii (19), it can be inferred that the genes detected were part of the blaCMY lineage. Salmonella Newport strain 0007-33 was used as the positive control (21). PCR was performed for the genes blaTEM, blaSHV, and blaCTX-M as published previously (9, 10, 27). Salmonella Newport strain 0007-33 was also used as the TEM and SHV positive control (21), and E. coli strain MISC 336 (CTX-M-1 positive) was used as the CTX-M positive control.The blaTEM, blaSHV, and blaCTX-M PCR products were sequenced using both strands of DNA for each PCR product. Protein sequences were aligned using Lasergene software (DNASTAR, Inc., Madison, WI) and included GenBank sequences (http://www.ncbi.nlm.nih.gov/GenBank/index.html) to confirm ESBL genotype. Mutations were evaluated with reference to the Lahey Clinic website (http://www.lahey.org/studies/). GenBank accession numbers used for alignment of protein sequences were AAR25033 for TEM-1 and ABF29674 for SHV-2. CTX-M accession numbers were derived from a list on the Lahey Clinic website. Specific primers for the CTX-M-1 group (M13U and M13L) were used to amplify the entire coding sequences of these blaCTX-M genes (23). Sequencing and analysis were carried out as described above to identify the specific CTX-M subtype.Seventy of the 150 E. coli isolates had an “ESBL alert” on the MicroScan Walkaway, and after removal of redundant isolates, 60 isolates were tested further. ESBL confirmatory testing was positive for six of these 60 isolates (Table (Table1,1, column 3).

TABLE 1.

Distribution of β-lactamase and extended-spectrum β-lactamase genes
IsolateSpeciesESBL testblaCMYblaSHVblaTEMblaCTX-M
1Canine+SHV-12TEM-1
3Canine+
6Canine+TEM-1
11Canine+TEM-1
17Canine+TEM-1
19Canine+
21Feline+CTX-M-15
26Canine+TEM-1
27Canine+TEM-1
31Canine+TEM-1
32Canine+TEM-1
33Canine+
41Canine+TEM-1
42Canine+
44Canine+TEM-1
53Canine+
57Feline+
62Canine+TEM-1
67Canine+
74Canine+TEM-1
75Feline+TEM-1CTX-M-15
82Feline+TEM-1
85Canine+TEM-1
86Canine+TEM-1
87Canine+
88Feline+
91Canine+
98Canine+TEM-1
102Canine+
104Canine+
112Canine+TEM-1
119Canine+
131CanineNot doneNot doneNot done
133Feline+
138Feline+TEM-1
147Canine+TEM-1
149Feline+TEM-1
157Canine+CTX-M-15
165Canine+TEM-1
166Canine+TEM-1
168Canine+
182Canine+
183Feline+TEM-1
190Feline+TEM-1
205Canine+TEM-1
209Canine+TEM-1
210Canine+
219Feline+TEM-1CTX-M-15
220Canine+TEM-1CTX-M-15
230Canine+TEM-1
234Canine+TEM-1
236Canine+TEM-1CTX-M-14
240Canine+TEM-1
242Canine+TEM-1CTX-M-15
246Feline+CTX-M-15
265Canine+TEM-1CTX-M-15
266Canine+TEM-1
267Canine+CTX-M-15
268Canine+
269Canine+
Open in a separate windowFifty-three of the 60 isolates were positive for the blaCMY gene (43 canine and 10 feline samples) (Table (Table1,1, column 4). Of the seven negative isolates, six were those previously found to be positive for ESBL production via ESBL testing. The remaining canine isolate (isolate 131, negative for both blaCMY PCR and ESBL testing) was not analyzed further.A total of 40 E. coli isolates were found to carry one or more β-lactamase genes. PCR detected a blaSHV gene in one canine isolate, a blaTEM gene in 29 canine and seven feline isolates, and a blaCTX-M gene in six canine and four feline isolates. The remaining 19 isolates were confirmed as negative for blaSHV, blaTEM, and blaCTX-M genes (Table (Table1,1, columns 5 to 7).Based on sequence analysis (Table (Table1,1, column 5), the one SHV-positive strain (isolate 1) was determined to carry SHV-12. All 36 strains positive for a TEM gene were identified as carrying TEM-1 (Table (Table1,1, column 6). Nine of 10 strains positive for a CTX-M gene carried genes of the CTX-M-1 group. The final strain (isolate 236) was concluded to carry CTX-M-14, given the presence of mutations in the amplified consensus sequence unique to CTX-M-14. DNA from each of the nine CTX-M-1 group strains was amplified using the CTX-M-1 group-specific primer set (23). Sequence analysis identified all nine strains as carrying CTX-M-15 (Table (Table1,1, column 7).The CTX-M-type ESBLs identified in this study provide evidence for the dissemination of these genes in the United States. The CTX-M-1 group has frequently been reported in animals in countries other than the United States (2, 4, 8, 11, 17). The CTX-M-15 gene has not been identified in any bacterial isolate from animals in the United States. Animal sources of this gene have been identified only in E. coli isolated from the cloacae of Belgian poultry and an E. coli isolate from the urine of a cow in France (14, 24). The presence of CTX-M-14 genes in the Enterobacteriaceae has also been documented across the globe (22), including in six E. coli strains isolated from the feces of dogs in Chile (17). The CTX-M-14 gene identified in this study is the first identified from an animal in the United States and the first linked to a clinical case of UTI in a dog. The SHV-12 gene has been detected in bacterial isolates from animals in several countries (1, 4, 6, 26), including, in 2005, a Salmonella enterica serovar Newport strain from a horse in the United States (21).Interestingly, of the 11 ESBL-positive isolates identified by sequence analysis, only six were positive by ESBL confirmatory testing, likely due to the concurrent presence of a blaCMY gene in the other five isolates. The product of this gene is known to mask the effects of clavulanic acid on the ESBL confirmatory test (28). By relying on ESBL confirmatory testing alone, it is likely that the prevalence of ESBLs is being underestimated, particularly in populations with a high frequency of blaAmpC, such as in the current study. During the study period, E. coli was isolated from samples submitted to our laboratory from 1,318 individual animals, and of these, 257 met the criteria for ESBL confirmatory testing. Of the 257 isolates tested, 14 (5%) were identified as being ESBL producers based on the Etest method, including the six reported in this study. The overall prevalence of E. coli isolates that were positive for an ESBL during this period was 1% (S. Rankin, unpublished observations). It is possible that this is an underestimation, based on the high frequency of blaCMY detected in this study.Though, historically, the most common ESBLs in the United States have been TEM and SHV types (16, 20), more recent studies have identified CTX-M genes, first reported in 2003 (15). CTX-M ESBLs now predominate in some U.S. health care systems (12). The findings from the current study are in agreement with current trends in the United States and other parts of the world (3). This study is the first report of E. coli strains that encode SHV-12, CTX-M-15, and CTX-M-14 ESBL genes in companion animals in the United States.  相似文献   

6.
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   

7.
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   

8.
MICs were determined for an investigational ketolide, CEM-101, and azithromycin, telithromycin, doxycycline, levofloxacin, clindamycin, and linezolid against 36 Mycoplasma pneumoniae, 5 Mycoplasma genitalium, 13 Mycoplasma hominis, 15 Mycoplasma fermentans, and 20 Ureaplasma isolates. All isolates, including two macrolide-resistant M. pneumoniae isolates, were inhibited by CEM-101 at ≤0.5 μg/ml, making CEM-101 the most potent compound tested.Mycoplasma pneumoniae, Mycoplasma hominis, Mycoplasma genitalium, Mycoplasma fermentans, and Ureaplasma spp. isolates are responsible for infections in the respiratory and urogenital tracts (17, 18). Macrolides have historically been the treatments of choice for M. pneumoniae respiratory infections of adults and children because they have the advantages of being safe and well tolerated in oral formulations and of possessing antiinflammatory properties independent of their antibacterial activity and activity against other microorganisms that may cause clinically similar illness. These properties have also made macrolides attractive for empirical treatment, since mycoplasmal infections are rarely confirmed by microbiological testing. Macrolides are also active against some other Mycoplasma spp., as well as Ureaplasma spp. M. fermentans and M. hominis, however, are resistant to some members of this class, such as erythromycin, but are susceptible to clindamycin (19).During the past several years, concerns have arisen over the impact of the widespread use of macrolides on antimicrobial resistance in respiratory pathogens, such as Streptococcus pneumoniae isolates, among which 30% or more of clinical isolates are no longer susceptible to macrolides and may not respond to treatment with these drugs (7). Recent publications from Japan have confirmed the emergence in 10 to 33% of M. pneumoniae isolates of macrolide resistance that may have implications for patient outcomes (8, 10, 11, 14). These isolates typically have mutations in domain V of the 23S rRNA gene and erythromycin MICs of 32 to >64 μg/ml. A recent report from Shanghai, China, documented that 39 of 50 (78%) M. pneumoniae strains isolated there were macrolide resistant (6). Macrolide-resistant M. pneumoniae has also been reported from France (12) and the United States (20, 21). The Centers for Disease Control and Prevention recently described 3 of 11 cases (27%) of M. pneumoniae infections from an outbreak in Rhode Island that were macrolide resistant (20). We have encountered two children in Birmingham, AL, with macrolide-resistant M. pneumoniae infections of the lower respiratory tract who did not respond initially to treatment with azithromycin and required several days of hospitalization (21). Fluoroquinolone resistance has been described in genital mycoplasmas (1, 3), and tetracycline resistance may now exceed 40% in some populations (17). Azithromycin resistance associated with clinical treatment failure has also been documented in M. genitalium (4). These findings clearly indicate the need for new drug classes or improvements in drugs of existing classes for treatment of mycoplasmal and ureaplasmal infections.CEM-101 is a new ketolide with activity against many bacteria that cause respiratory and/or urogenital infections, such macrolide-resistant S. pneumoniae, chlamydiae, Haemophilus influenzae, Moraxella catarrhalis, and Neisseria gonorrhoeae (2, 5, 9, 13). To investigate further the antimicrobial spectrum of CEM-101, we studied its vitro activities against human mycoplasmas and ureaplasmas in comparison to the activities of other antimicrobial agents (Table (Table11).

TABLE 1.

In vitro activities of CEM-101 and other antimicrobials against Mycoplasma and Ureaplasma species isolated from humans
Organism (no. of isolates) and antimicrobialMIC range (μg/ml)MIC50 (μg/ml)aMIC90 (μg/ml)a
M. pneumoniae (36)
    CEM-101≤0.000000063-0.50.0000320.000125
    Azithromycin≤0.000016 to ≥320.000250.0005
    Telithromycin0.000031 to ≥320.000250.001
    Doxycycline0.016-0.250.1250.25
    Levofloxacin0.125-10.50.5
    Linezolid32-12864128
M. genitalium (5)
    CEM-101≤0.000032NANA
    Azithromycin≤0.000032-0.005NANA
    Telithromycin≤0.00003-0.00025NANA
    Doxycycline≤0.008-0.031NANA
    Levofloxacin0.125-1NANA
    Linezolid4-128NANA
M. fermentans (15)
    CEM-101≤0.008≤0.008≤0.008
    Azithromycin0.125-10.50.5
    Telithromycin0.002-0.031≤0.0080.016
    Clindamycin≤0.008-0.0630.0160.031
    Doxycycline0.016-0.50.1250.5
    Levofloxacin≤0.008-0.250.0310.125
    Linezolid0.5-414
M. hominis (13)
    CEM-1010.002-0.0080.0040.008
    Azithromycin0.5-442
    Telithromycin0.125-0.50.250.5
    Clindamycin≤0.008-0.031≤0.0080.016
    Doxycycline≤0.008-0.0160.1258
    Levofloxacin0.125-0.50.250.5
    Linezolid1-824
U. parvum (10)
    CEM-1010.002-0.0310.0080.016
    Azithromycin0.5-424
    Telithromycin0.008-0.0630.0630.125
    Doxycycline0.031-16816
    Levofloxacin0.125-20.52
    Linezolid128 to >256>256>256
U. urealyticum (10)
    CEM-1010.004-0.0630.0080.031
    Azithromycin0.5-424
    Telithromycin0.016-0.250.0630.25
    Doxycycline0.031-32116
    Levofloxacin0.5-20.51
    Linezolid256 to >256>256>256
Open in a separate windowaNA, not applicable.Thirty-six M. pneumoniae isolates collected between 1992 and 2006 from the respiratory tracts of adults and children with pneumonia were tested. These included two macrolide-resistant isolates with azithromycin MICs of >32 μg/ml (21), both of which had been shown to have an A2063G mutation in domain V of the rRNA gene. The M. genitalium isolates included reference strains obtained from the urogenital tracts of patients in the United States (three isolates) and Denmark (two isolates). Fifteen M. fermentans isolates from the respiratory or urogenital tracts were obtained from the Mycoplasma Collection at the National Institutes of Health and patients in Birmingham, AL, between 1992 and 2004. Thirteen M. hominis isolates were obtained from clinical specimens from the urogenital tract or wounds between 1994 and 2007. Two isolates were resistant to doxycycline (MICs of 8 to 16 μg/ml). Ten Ureaplasma parvum isolates were obtained from urogenital specimens between 2002 and 2005. Seven were doxycycline resistant (MICs of 4 to 16 μg/ml). Ten U. urealyticum isolates were obtained from various urogenital tract, placenta, or neonatal respiratory secretion specimens between 1990 and 2005. Four were resistant to doxycycline (MICs of 4 to 32 μg/ml).Antimicrobial powders were dissolved as instructed by the manufacturers and frozen in 1-ml aliquots containing 2,048 μg/ml. The drugs tested included CEM-101, azithromycin, telithromycin, doxycycline, levofloxacin, and linezolid. A working dilution of each drug was prepared on the day of each assay based on the anticipated MIC ranges. Serial twofold antimicrobial dilutions were performed in 10B broth for Ureaplasma spp. and SP4 broth for Mycoplasma spp. in 96-well microtiter plates, and MICs were determined as previously described (16). The MIC was defined as the lowest concentration of a drug in which the metabolism of the organisms was inhibited, as evidenced by lack of color change at the time the drug-free control first showed a change in color. The inoculum of each isolate was verified by serial dilutions and plate counts. The quality control strains used to validate the accuracy of the MICs of the antimicrobial agents being compared included M. pneumoniae (UAB-834), M. hominis (UAB-5155), and U. urealyticum (UAB-4817), all of which are low-passage clinical isolates for which a three-dilution MIC range has been established. Nine M. pneumoniae isolates were tested to determine minimal bactericidal concentrations (MBCs) for CEM-101. Aliquots (30 μl) from each well that had not changed color at the time the MIC was read were added to 2.97 ml broth (1:100 dilution) to make certain the drug was diluted below the inhibitory concentration, to allow living organisms to grow to detectable levels. The growth control was subcultured to ensure the presence of viable organisms in the absence of the drug. Broths were incubated at 37°C. The MBC was defined as the concentration of the antimicrobial at which no growth was apparent, as shown by lack of color change in the broth after prolonged incubation.CEM-101 demonstrated the greatest overall potency against all species of human mycoplasmas and ureaplasmas tested when compared to azithromycin, telithromycin, doxycycline, levofloxacin, and linezolid. Excluding the two macrolide-resistant M. pneumoniae isolates, no isolate of any species tested had an MIC of >0.063 μg/ml for CEM-101. M. pneumoniae MICs for CEM-101 ranged from ≤0.000000063 to 0.5 μg/ml, with a MIC90 of 0.000125, making its activity fourfold higher than that of azithromycin and eightfold higher than that of telithromycin. CEM-101 MICs against two isolates with elevated MICs for azithromycin (MICs of >32 μg/ml) and telithromycin (MICs of 4 μg/ml) were 0.5 μg/ml. CEM-101 was equally active against doxycycline-susceptible and -resistant M. hominis and Ureaplasma spp. isolates. Linezolid was inactive against M. pneumoniae and Ureaplasma spp. isolates, but some M. fermentans and M. hominis isolates had linezolid MICs of ≤1 mg/ml. CEM-101 MBCs were ≥16-fold higher than the MICs for nine M. pneumoniae isolates, indicating that the drug is bacteriostatic against this organism, as are other agents in the macrolide and ketolide class (15).As a ketolide, CEM-101 is able to bind both domain II and V of rRNA, thus explaining why it maintains in vitro activity against macrolide-resistant M. pneumoniae isolates that have altered binding sites in domain V due to the A2063G mutation. As shown for other pathogens, such as the streptococci (9) and H. influenzae (5), CEM-101 has lower MICs than telithromycin for the human mycoplasmas. The side chain on the CEM-101 molecule differs from that of telithromycin, and it also has a fluorine atom at position 2 of the macrolide ring which could make the drug more lipophilic and facilitate more-efficient binding to the mycoplasma ribosome. CEM-101 maintained reasonably good in vitro activity against the two azithromycin-resistant M. pneumoniae isolates, with MICs of 0.5 μg/ml, while the 4-μg/ml telithromycin MICs exceeded the breakpoint of 1 μg/ml used to designate susceptibility for other bacterial species. The difference between the MIC50 for the M. pneumoniae group overall and the MICs for the two resistant isolates was the same at 14 twofold dilutions for each drug. This finding suggests that the in vitro activities of CEM-101 and telithromycin were affected in a similar manner by the rRNA mutation, but the lower MICs for CEM-101 could give it an advantage in treating infections caused by these organisms.Our study indicates that CEM-101 is active in vitro against six mycoplasmal and ureaplasmal species that are clinically important in humans, including azithromycin and telithromycin-resistant M. pneumoniae and doxycycline-resistant M. hominis and Ureaplasma spp. isolates. The results of other investigations documenting the in vitro activities of CEM-101 against numerous other gram-positive and gram-negative bacterial pathogens, including agents of community-acquired pneumonia and gonococcal and nongonococcal urethritis, suggest that this agent has great potential in future clinical studies.  相似文献   

9.
Sitafloxacin showed MICs of less than or equal to 0.5 μg/ml against 105 isolates of Helicobacter pylori, including 44 isolates with mutations in the gyrA gene. The highest MICs for garenoxacin and levofloxacin were 8 and 64 times, respectively, higher than the highest MICs observed for sitafloxacin.The guidelines for the management and treatment of Helicobacter pylori infections established by the European Helicobacter Study Group Third Masstricht Consensus Report recommend an eradication antimicrobial chemotherapy consisting of amoxicillin, clarithromycin, and a proton pump inhibitor alone or in combination with metronidazole and clarithromycin (10). On the other hand, a trend toward increased clarithromycin resistance in Japan has been reported (8); furthermore, high metronidazole resistance rates associated with H. pylori eradication failure have been seen in the United States, Europe, and Asia with the exception of Japan (11). In the search for alternative eradication treatment regimens, it has been recently reported in the United States and Europe that levofloxacin may be efficacious in H. pylori eradication therapy (7, 13). At the same time, the increased use of levofloxacin-based eradication regimens has led to increasing resistance to levofloxacin in H. pylori as a result of mutations in the quinolone resistance-determining region (QRDR) of the gyrA gene correlating with the decreased effectiveness of levofloxacin in eradication regimens (16). Sitafloxacin is a recently developed fluoroquinolone with wide-spectrum activity, ranging from gram-positive cocci to gram-negative bacilli (1, 15). We studied the effect of mutations in the gyrA gene and its impact on the antimicrobial activity of sitafloxacin in H. pylori.(This study was presented at the 48th Interscience Conference on Antimicrobial Agents and Chemotherapy, Washington, DC, 25 to 28 October 2008.)A total of 105 H. pylori isolates were recovered from the gastric mucosa of patients presenting with gastroduodenal diseases in health care facilities in Japan between 2004 and 2005. Of the 105 patients, 57 (54.3%) were males and 48 (45.7%) were females, and the average age was 57.9 years (range in age from 21 to 84). None of the patients had previously undergone eradication therapy. The spectrum of peptic ulcers included 39 (37.1%) cases of chronic gastritis, 21 (20.0%) gastric ulcers, 18 (17.1%) duodenal ulcers, 9 (8.6%) gastric cancers, 8 (7.6%) gastroduodenal ulcers, and 10 (9.5%) cases with other causes or an unspecified diagnosis. Only one isolate per patient was included among the 105 isolates.Susceptibilities to sitafloxacin (Daiichi Sankyo, Japan), garenoxacin [a novel des-fluoro(6)quinolone (Astellas, Japan) (6)], and levofloxacin (Daiichi Sankyo, Japan) were determined by agar dilution method according to CLSI guidelines by using drugs with known potency (4, 5). The agar dilution method was performed by serial twofold dilution on Mueller-Hinton agar (Becton Dickinson, MD) with 5% sheep blood using 1 to 3 μl of a McFarland 2.0-adjusted inoculum and incubation at 35 ± 2°C for 72 h under microaerophilic conditions. For quality control, H. pylori ATCC 43504 was tested with each run.PCR amplification and sequence analysis of the QRDR of gyrA was performed as previously described by Nishizawa et al. (12). The specific primers used for amplification were GYRA-F (5′-TTTRGCTTATTCMATGAGCGT-3′) and GYRA-R (5′-GCAGACGGCTTGGTARAATA-3′). For sequencing, the ABI Prism 3130xl genetic analyzer (Perkin-Elmer, ABI, CA) was used. Sequences were compared to that of the H. pylori wild-type strain (GenBank accession no. L29481).Forty-four of the 105 H. pylori isolates exhibited mutations in the gyrA gene (Table (Table1).1). Mutations at Asn87 were observed in 14 isolates, while mutations were seen at Asp91 in 25 isolates. The remaining five isolates had mutations in other regions. Table Table22 shows the effect of changes in the gyrA gene in the QRDR and its effect on the MICs of sitafloxacin, garenoxacin, and levofloxacin. Mutations involving Asn87 resulted in a shift to higher MIC levels of the drugs than mutations in other regions. Sitafloxacin demonstrated the narrowest MIC distribution with a MIC of ≤0.5 μg/ml against all isolates. In contrast, the highest MICs for garenoxacin and levofloxacin were 8 and 64 times, respectively, higher than the highest MIC observed for sitafloxacin. A scattergram depicting sitafloxacin MICs versus levofloxacin or garenoxacin MICs for 105 isolates with or without gyrA mutations is shown in Fig. Fig.1.1. From the scattergram, the increase in garenoxacin and levofloxacin MICs relative to the sitafloxacin MICs is observed. With respect to levofloxacin-resistant isolates with MICs ranging from 2 to 32 μg/ml, garenoxacin and sitafloxacin demonstrated MICs ranging from 0.125 to 4 μg/ml and ≥0.015 to 0.5 μg/ml, respectively.Open in a separate windowFIG. 1.Scattergram of sitafloxacin MICs versus levofloxacin or garenoxacin MICs for 105 Helicobacter pylori isolates. The red and black numbers indicate the frequency of isolates with and without gyrA mutations, respectively.

TABLE 1.

Genetic characteristics of Helicobacter pylori isolates in this study
Strain and no. of isolatesAmino acid substitution at position:
8791Other
Strains with wild-type gyrA
    61AsnAsp
Strains with gyrA mutations at Asn87 (n = 14)
    11LysAsp
    2LysAspAsp143Glu
    1IleAsp
Strains with gyrA mutations at Asp91 (n = 25)
    9AsnGly
    1AsnGlyAsp145Gly
    5AsnAsn
    2AsnAsnAsp143Glu
    6AsnTyr
    1AsnTyrAla97Val
    1AsnTyrAla129Val
Strains with gyrA mutations in other regions (n = 5)
    1AsnAspThr62Ile
    1AsnAspAsp99Val
    1AsnAspArg130Lys
    1AsnAspAsp143Glu
    1AsnAspLys158Arg
Open in a separate window

TABLE 2.

Correlation of quinolone MICs with gyrA mutations for 105 Helicobacter pylori isolates
AgentgyrA mutationNo. of isolatesMIC rangeMIC50MIC90
SitafloxacinAsn87140.03-0.50.250.5
Asp9125≤0.015-0.250.120.25
Other5≤0.015-0.12
None61≤0.015-0.060.030.03
GarenoxacinAsn87140.12-40.51
Asp91250.12-20.51
Other5≤0.015-0.5
None61≤0.015-0.250.030.06
LevofloxacinAsn87144-32816
Asp91252-848
Other50.12-4
None610.12-20.250.25
Open in a separate windowIn recent years, the increased resistance to clarithromycin and metronidazole of H. pylori has led to the inclusion of fluoroquinolones in eradication therapy in light of their minimal side effects and improved eradication rates (7, 13). However, resistance to a fluoroquinolone as a result of mutations in gyrA has been reported (12). Furthermore, a significant reduction in eradication effectiveness was observed when levofloxacin was used against H. pylori isolates exhibiting gyrA mutations compared to H. pylori isolates with no mutations in gyrA (16). Nishizawa et al. found elevated MICs to gatifloxacin along with mutations in gyrA in 47.9% of the isolates recovered from patients who had failed H. pylori eradication therapy (12).We observed that 41 (39%) of the H. pylori isolates recovered from patients prior to undergoing their first eradication regimen already showed resistance to levofloxacin based on a breakpoint of >1 μg/ml (2). The rate of resistance to levofloxacin we observed was considerably higher than the figure found in the American College of Gastroenterology Guideline (3). The prevalence of levofloxacin-resistant H. pylori may be associated with the increasing use of fluoroquinolones in clinical practice for many indications in Japan since the 1980s. There was little difference observed in the rate of resistance to clarithromycin between levofloxacin-resistant (14/41 [34%]) and -susceptible isolates (20/64 [31%]) in this study (data not shown). Compared to data generated in southeast Asia and Europe (9, 14), the prevalence of metronidazole-resistant isolates observed in our study was lower (5/105 [4.8%]). Two of the five isolates were resistant to levofloxacin.Bogaerts et al. has previously reported that two amino acid substitutions due to mutations in gyrA elevated MICs to levofloxacin, ciprofloxacin, and moxifloxacin (2). While we did not observe H. pylori isolates with mutations in both Asn87 and Asp91, there is a need for continued surveillance for the emergence of high resistance in Japan in light of the high use of fluoroquinolones.In this study, we observed the superior antibacterial activity of sitafloxacin against H. pylori compared to that of levofloxacin and garenoxacin even in the presence of mutations in gyrA. With respect to other organisms, sitafloxacin has been reported to have high affinity to DNA gyrase and topoisomerase IV along with superior antibacterial activity (1). As H. pylori lacks the topoisomerase IV enzyme, the high affinity of sitafloxacin to DNA gyrase may account for its lower MIC against H. pylori. Our report is the first to demonstrate the antibacterial activity of sitafloxacin against H. pylori isolates with gyrA mutations. Based on sitafloxacin''s superior antibacterial activity, clinical trials of second- or third-line eradication therapy including sitafloxacin are warranted.  相似文献   

10.
11.
12.
The in vitro activity of thimerosal versus those of amphotericin B and natamycin was assessed against 244 ocular fungal isolates. The activity of thimerosal against Fusarium spp., Aspergillus spp., and Alternaria alternata was 256 times, 512 times, and 128 times, respectively, greater than that of natamycin and 64 times, 32 times, and 32 times, respectively, greater than that of amphotericin B. Thimerosal''s antifungal activity was significantly superior to those of amphotericin B and natamycin against ocular pathogenic fungi in vitro.The problem of keratomycosis in developing countries like China is more acute because of its higher incidence and the unavailability of effective antifungals (18, 28, 30). To date, only fluconazole and natamycin are commercially available for ocular use in China. Fluconazole has high bioavailability against Candida spp., but Fusarium spp. and Aspergillus spp. are resistant to it (3, 27, 29). Fusarium spp. and Aspergillus spp. are more commonly associated with keratomycosis, while Candida spp. are rarely implicated as etiological agents of keratomycosis in China (23, 26). Natamycin is the only topical ophthalmic antifungal compound approved in the United States (14). Natamycin is poorly soluble in water. After topical application, natamycin penetrates the cornea and conjunctiva poorly and effective drug levels are not achieved in either the cornea or the aqueous humor (15); it is therefore useful only in the treatment of superficial keratomycosis. Due to the relative unavailability of effective antifungals, keratomycosis fails to resolve in many of the patients who receive antifungal treatment; some patients experience vision loss and eventually corneal perforation, ultimately require penetrating keratoplasty, or even enucleation or evisceration (20, 28). Therefore, it is very important and urgent to explore broad-spectrum antifungals to effectively suppress a wide variety of ocular fungal pathogens and to develop new antifungal eye drops to combat this vision-threatening infection.Thimerosal is a preservative commonly used in ophthalmic solutions, otic drops, topical medicine, and vaccines because of its bactericidal property. However, the efficacy of thimerosal against ocular pathogenic fungi has not been evaluated so far. The present study was performed to determine the antifungal activity of thimerosal versus those of amphotericin B and natamycin against ocular pathogenic fungi in vitro. To our knowledge, this is the first study to determine the antifungal activity of thimerosal against main ocular pathogenic fungi. Results obtained in this study may contribute to the development of new antifungal eye drops.Two hundred forty-four strains of fungi isolated from patients with keratomycosis from the Henan Eye Institute in Zhengzhou, China, were investigated. These isolates were identified based on morphology by standard methods (22, 23, 25). They included 136 Fusarium isolates, 98 Aspergillus isolates, and 10 Alternaria alternata isolates. Candida parapsilosis ATCC 22019 was used as quality control for each test.Thimerosal (Yili Pharmaceutical Co. Ltd., Beijing, China), amphotericin B (Bristol-Myers Squibb, Princeton, NJ), and natamycin (Yinxiang Biotechnology Co. Ltd., Zhejing, China) were studied. They were all dissolved in 100% dimethyl sulfoxide. The stock solutions were prepared at concentrations of 400 μg/ml for thimerosal and 1,600 μg/ml for amphotericin B and natamycin. Drug dilutions were made in RPMI 1640 medium buffered to pH 7.0 with 0.165 M morpholinepropanesulfonic acid. Final concentrations ranged from 0.0078 to 4 μg/ml for thimerosal and from 0.0313 to 16 μg/ml for amphotericin B and natamycin.A broth microdilution method was performed by following the Clinical and Laboratory Standards Institute M38-A2 document (13). The final inoculum was 0.4 × 104 to 5 × 104 CFU/ml. Following incubation at 35°C for 48 h, the MIC was determined as the lowest concentration of amphotericin B, natamycin, or thimerosal that prevented any discernible growth.The MIC range and mode, the MIC for 50% of the strains tested (MIC50), and the MIC90 were provided for the isolates with the SPSS statistical package (version 13.0). For calculation, any high off-scale MIC was converted to the next higher concentration.The in vitro activities of thimerosal, amphotericin B, and natamycin against the isolates are summarized in Tables Tables11 and and2.2. When comparing the MIC90s of thimerosal with those of natamycin and amphotericin B, the activity of thimerosal against Fusarium spp. is 256 times greater than that of natamycin and 64 times greater than that of amphotericin B, the activity of thimerosal against Aspergillus spp. is 512 times greater than that of natamycin and 32 times greater than that of amphotericin B, and the activity of thimerosal against Alternaria alternata is 128 times greater than that of natamycin and 32 times greater than that of amphotericin B. Therefore, thimerosal''s effect was significantly superior to those of amphotericin B and natamycin against main ocular pathogenic fungi in vitro.

TABLE 1.

In vitro susceptibilities of ocular Fusarium isolates to thimerosal, amphotericin B, and natamycin
Organism (no. of isolates) and antifungal agentMIC range (mode)bMIC50MIC90
Fusarium solani species complex (82)
    Thimerosal0.0078-0.0313 (0.0156)0.01560.0313
    Amphotericin B0.5-16 (1)12
    Natamycin4-32 (4)48
Fusarium moniliforme species complex (20)
    Thimerosal0.0156-0.0313 (0.0156)0.01560.0313
    Amphotericin B1-8 (2)22
    Natamycin4-8 (4)48
Fusarium avenaceum species complex (16)
    Thimerosal0.0156-0.0313 (0.0156)0.01560.0313
    Amphotericin B0.5-8 (2)24
    Natamycin4-32 (8)88
Other Fusarium isolates (18)a
    Thimerosal0.0078-0.0625 (0.0156)0.01560.0313
    Amphotericin B0.5-2 (1)12
    Natamycin4-8 (4)48
Fusarium spp. (136)
    Thimerosal0.0078-0.0625 (0.0156)0.01560.0313
    Amphotericin B0.5-16 (1)12
    Natamycin4-32 (4)48
Open in a separate windowaIncludes 9 strains of Fusarium oxysporum species complex, 5 strains of Fusarium poae species complex, and 4 strains of Fusarium lateritium species complex.bValues are in micrograms per milliliter.

TABLE 2.

In vitro susceptibilities of ocular Aspergillus and Alternaria alternata isolates to thimerosal, amphotericin B, and natamycin
Organism (no. of isolates) and antifungal agentMIC range (mode)bMIC50MIC90
Aspergillus flavus species complex (49)
    Thimerosal0.0313-0.0625 (0.0625)0.06250.0625
    Amphotericin B1-32 (2)22
    Natamycin8-32 (32)3232
Aspergillus fumigatus species complex (11)
    Thimerosal0.0156-0.0625 (0.0313)0.03130.0625
    Amphotericin B0.5-4 (1)12
    Natamycin4-32 (4)44
Aspergillus oryzae species complex (12)
    Thimerosal0.0156-0.0625 (0.0625)0.06250.0625
    Amphotericin B1-2 (1)12
    Natamycin4-32 (32)3232
Aspergillus versicolor species complex (12)
    Thimerosal0.0078-0.0625 (0.0078)0.01560.0625
    Amphotericin B0.5-2 (1)12
    Natamycin4-32 (8)832
Other Aspergillus isolates (14)a
    Thimerosal0.0078-0.0625 (0.0156)0.01560.0313
    Amphotericin B0.125-2 (1)12
    Natamycin0.25-32 (4)432
Aspergillus spp. (98)
    Thimerosal0.0078-0.0625 (0.0625)0.06250.0625
    Amphotericin B0.125-32 (1)12
    Natamycin0.25-32 (32)3232
Alternaria alternata (10)
    Thimerosal0.0078-0.0625 (0.0156)0.01560.0313
    Amphotericin B0.0625-2 (0.125)0.1251
    Natamycin2-8 (4)44
Open in a separate windowaIncludes 8 strains of Aspergillus niger species complex, 2 strains of Aspergillus candidus, 2 strains of Aspergillus nidulans, 1 strain of Aspergillus ochraceus, and 1 strain of Aspergillus wentii.bValues are in micrograms per milliliter.As shown in Tables Tables11 and and2,2, thimerosal has activity against different Aspergillus and Fusarium complexes. For each of these genera, this activity remains consistent and does not show significant interspecies variability. On the other hand, natamycin shows various activities against different Aspergillus spp. Most Aspergillus spp. are not susceptible, but Aspergillus fumigatus complex is susceptible to natamycin.A noteworthy finding is that thimerosal exhibits the greatest activity against Fusarium spp. in comparison to the effects of all of the antifungals studied in vitro to date. Some studies (10-12) of the in vitro efficacy of traditional and newer antifungals against keratitis and endophthalmitis fungal pathogens show that amphotericin B and voriconazole have the lowest MIC90s (2 to 4 μg/ml) against Fusarium spp., closely followed by terbinafine (8 μg/ml), natamycin (16 μg/ml), posaconazole (>8 μg/ml), itraconazole (>16 μg/ml), ketoconazole (>16 μg/ml), caspofungin (>16 μg/ml), 5-flucytosine (>64 μg/ml), and fluconazole (>256 μg/ml). When comparing the MIC90s of thimerosal with those of other antifungals, the activity of thimerosal against Fusarium spp. is 64 to >8,179 times greater than those of other antifungals. It is very important because Fusarium spp. remain the most frequently isolated ocular fungal pathogens in China, Portugal, Singapore, Australia, and the southern United States (2, 7, 12, 16, 18, 19, 23, 24, 26, 30) and the second most frequently isolated ocular fungal pathogens in India, Nepal, and Saudi Arabia (4, 8, 9, 17).Successful treatment of otomycosis with thimerosal has been reported by Tisner et al. (21). Recently, our primary work based on clinical trials addressed the suitability of thimerosal for the treatment of keratitis. At the Henan Eye Institute and the Anyang Eye Hospital, 21 patients with filamentous keratomycosis were treated with thimerosal because they were not improving after topical amphotericin B, ketoconazole, and natamycin treatment for 1 to 3 weeks. Twenty of the 21 infections responded well to thimerosal. The keratomycosis healed after topical thimerosal treatment for 14 to 45 days (unpublished data).Thimerosal is one of the main preservatives used worldwide in topical ophthalmic preparations, at concentrations ranging from 0.004 to 0.01%. Thimerosal has generally been accepted as a safe preservative agent in eye drops, and ocular side effects due to thimerosal are rare. No toxic effects have been observed following topical application of solutions containing thimerosal, even at concentrations 100 times higher than those required for a bactericidal effect (1, 5, 6). The findings from our study indicate that products formulated with thimerosal as both the main drug and a preservative can probably be used to treat keratomycosis successfully. We think that thimerosal has both antifungal and preservative effects without exceeding its value as a preservative and that the benefits of treating keratomycosis outweigh the potential risks for thimerosal.In conclusion, in this study, thimerosal exhibited potent in vitro activity against main ocular pathogenic fungi and was even more effective than amphotericin B and natamycin. The results suggest that thimerosal might play a role as a main drug in the treatment of keratomycosis and should be subjected to a prospective evaluation of efficacy and safety to further develop its clinical applications.  相似文献   

13.
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   

14.
A Pseudomonas aeruginosa isolate recovered in Belgium produced a novel extended-spectrum ß-lactamase, BEL-2, differing from BEL-1 by a single Leu162Phe substitution. That modification significantly altered the kinetic properties of the enzyme, increasing its affinity for expanded-spectrum cephalosporins. The blaBEL-2 gene was identified from a P. aeruginosa isolate clonally related to another blaBEL-1-positive isolate.Extended-spectrum ß-lactamases (ESBLs), such as TEM, SHV, PER, VEB, GES, and more recently, CTX-M variants, are reported increasingly to be found in Pseudomonas aeruginosa in various areas (1, 2, 7, 8, 10-12, 15, 17, 21, 23, 27, 28, 30). The BEL-1 ß-lactamase, distantly related to other ESBLs, was identified from a P. aeruginosa isolate from Roeselare, Belgium, which interestingly shows resistance to ticarcillin and ceftazidime but only reduced susceptibility to piperacillin, cefepime, cefpirome, and aztreonam (24). The blaBEL-1 determinant was found as a gene cassette in the chromosome-borne class 1 integron, In120, that includes other resistance genes (aacA4, aadA5, and smr2) and that was part of a Tn1404-type transposon structure (24). Very recently, Bogaerts et al. (5) reported on the diffusion of BEL-1-producing isolates in various hospital centers of Belgium and also found that BEL-1 could be associated with other relevant β-lactamases, such as the VIM-1 metallo-β-lactamase (5).P. aeruginosa isolate 531 (this study) was recovered from a urine sample of a patient hospitalized in Roeselare, Belgium, in February 2007 for pneumonia and was resistant to all β-lactams but imipenem (Table (Table1).1). A synergy between aztreonam or ceftazidime and clavulanic acid-containing disks suggested the synthesis of an ESBL (19). PCR followed by sequencing using ESBL gene-specific primers (24) identified a novel gene encoding BEL-2, which differs from BEL-1 by a single amino acid substitution (Leu to Phe at Ambler position 162) (3). Transfer of a ß-lactam resistance marker from P. aeruginosa 531 to Escherichia coli or to P. aeruginosa reference strains was unsuccessful by either conjugation or transformation (25). Plasmid extraction performed as described previously (14) did not identify any plasmid, suggesting a chromosomal location of the blaBEL-2-like gene in P. aeruginosa 531. A pulsed-field gel electrophoresis (PFGE) analysis (4) showed that isolates 531 (BEL-2 positive) and 51170 (BEL-1 positive), recovered from the same geographical area, were clonally related. A PCR mapping approach confirmed the presence of a class 1 integron whose structure was identical to that of In120 of P. aeruginosa 51170 (24) and identified an identical structure in P. aeruginosa 531 (data not shown). Overall, these data suggest that the blaBEL-2 sequence likely resulted from a mutational event that had occurred in In120-carrying P. aeruginosa strains.

TABLE 1.

MICs of β-lactamsa
β-Lactam(s)bMIC (μg/ml)
P. aeruginosa 531P. aeruginosa 51170E. coli TOP10 (pSB-2) (BEL-2)E. coli TOP10 (pSB-1) (BEL-1)E. coli TOP10
Amoxicillin>512>512>512>5124
Amoxicillin and CLA>512>51264644
Ticarcillin>512>512>512>5124
Ticarcillin and CLA>51212864644
Piperacillin1616321281
Piperacillin and TZB888321
Cephalothin>512>5122562562
Cefuroxime>512>512321282
Cefoxitin512512222
Ceftazidime25632128160.06
Cefotaxime256323240.12
Cefepime16410.250.06
Cefpirome64160.50.250.06
Aztreonam643216160.12
Imipenem110.120.120.12
Open in a separate windowaMICs of β-lactams for P. aeruginosa 531 and 51170 clinical isolates, producing ESBLs BEL-2 and BEL-1, respectively, E. coli TOP10 harboring recombinant plasmid pSB-2 expressing BEL-2, E. coli TOP10 harboring recombinant plasmid pSB-1 expressing BEL-1, and the E. coli TOP10 reference strain.bCLA, clavulanic acid at a fixed concentration of 4 μg/ml; TZB, tazobactam at a fixed concentration of 4 μg/ml.In order to compare the contributions of BEL-1 and BEL-2 to ß-lactam resistance, the corresponding genes (amplified using primers PreBEL-A [5′-AGACGTAAGCCTATAATCTC] and PreBEL-B [5′-GCGAATTGTTAGACGTATG]) were cloned in the pCR-BluntII-TOPO vector (Invitrogen, Cergy-Pontoise, France) and subsequently introduced into E. coli TOP10, giving rise to recombinant strains E. coli TOP10(pSB-1) and E. coli TOP10(pSB-2), producing BEL-1 and BEL-2, respectively. MICs of ß-lactams were determined by solid agar dilutions following the guidelines of the CLSI (9). E. coli TOP10(pSB-2) had MICs of piperacillin, cephalothin, and cefuroxime that were lower than those of E. coli TOP10(pSB-1), but its cefotaxime, ceftazidime, ceftriaxone, and cefepime MICs were higher than those of TOP10(pSB-1), while MICs of carbapenems were the same (Table (Table11).E. coli TOP10(pSB-2) produced a ß-lactamase with a pI value of 7.1 (identical to that of BEL-1) (18). Approximately 1.5 mg of BEL-2 was purified (>95% as estimated by SDS-PAGE analysis; data not shown) from an E. coli MCT236(pET-BEL-2) crude extract by using a two-step chromatography process (an anion exchange at pH 7.5 using a Q Sepharose Fast Flow column followed by a cation exchange at pH 6.2 using a 1-ml Resource S column). (The specific activity was 8,800 nmol/min·mg of protein with 100 μM of cephalothin as the substrate, purified 95-fold.) BEL-2 had a broad-spectrum hydrolysis profile, including penicillins and expanded-spectrum cephalosporins but not cephamycins and carbapenems (Table (Table2).2). BEL-2 overall showed higher catalytic efficiencies (kcat/Km) than BEL-1 for aztreonam and most oxyiminocephalosporins (cefotaxime, ceftazidime, ceftriaxone, and cefepime but not cefuroxime). This was due to a significant alteration of the Km values for these substrates with BEL-2, which were decreased relative to those of BEL-1 by 300-fold (ceftriaxone) to up to three orders of magnitude (ceftazidime) (Table (Table2).2). Interestingly, a decrease of the Km value was also observed with all the other substrates (though the variation was less important), likely reflecting a modification of the active site structure and thus substrate recognition. Overall, BEL-2 kcat values were also lower but to a lesser extent (Table (Table2).2). The values of catalytic efficiency toward expanded-spectrum cephalosporins for BEL-2 may explain the higher MICs observed for the BEL-2-producing recombinant E. coli strains and P. aeruginosa clinical isolate. Position 162 is located at the beginning of the Ω loop, which bears the functionally important Glu166 residue, which is conserved in class A enzymes, and where mutations conferring extended-spectrum properties have been extensively reported in natural TEM and SHV variants (13). The presence of a bulky Phe residue in BEL-2 might modify the orientation of the Ω loop and the overall geometry of the active site. The further extension of the substrate profile as a consequence of a single substitution in the Ω loop observed with the BEL-2 variant may parallel that of other enzymes, e.g., CTX-M-19 (CTX-M-14 Pro167Ser variant) (26) or GES-2 (GES-1 Gly170Asn variant) (28). Inhibition studies showed that BEL-2 and BEL-1 are similarly inhibited by clavulanic acid, tazobactam, and sulbactam (50% inhibitory concentrations of 0.1, 2, and 3 μM, respectively).

TABLE 2.

Kinetic parameters of purified ß-lactamase BEL-2, in comparison with previously reported values of BEL-1 (22)a
ß-LactamBEL-2
BEL-1
kcat (s−1)Km (μM)kcat/Km (mM−1· s−1)kcat (s−1)Km (μM)kcat/Km (mM−1· s−1)
Benzylpenicillin1.24300320150
Piperacillin1.10.33,700215130
Cephalothin4.322,200150280540
Cephaloridine1.2815030130230
Cefoxitin<0.01NDND<0.01NDND
Cefuroxime0.8812731040250
Ceftriaxone0.20.12,0002530830
Cefotaxime0.130.4529030250120
Ceftazidime0.030.6447>1.5>700ND
Cefepime0.0030.31011507
Aztreonam0.10.3628010100100
Imipenem<0.01NDND<0.01NDND
Open in a separate windowaStandard deviations were below 15%. ND, not determinable, due to the initial rate of hydrolysis being too low.  相似文献   

15.
16.
Six Bordetella pertussis strains isolated from children in Japan from 2004 to 2006 showed high-level resistance to nalidixic acid (NAL; MIC, >256 μg/ml) and decreased susceptibilities to fluoroquinolones. All of the NAL-resistant strains had the same D87G mutation in gyrA.Pertussis is an acute respiratory tract infection caused by Bordetella pertussis that is particularly serious for neonates and infants (2, 5, 19). Although the introduction of whole-cell and acellular vaccines caused a drastic decrease in the incidence of pertussis globally compared with that in the prevaccine era, developed countries have experienced a marked increase over the past 15 years (2). There has also been a recent shift in the age distribution of pertussis patients to adults and adolescents, an unrecognized but significant source of infection for neonates and infants (2, 5, 9, 19). Macrolides are widely used for antimicrobial treatment and postexposure prophylaxis of pertussis (2, 5). However, several erythromycin-resistant strains of B. pertussis have emerged in the United States (13, 14, 20) and alternative therapeutic agents are being sought to combat these strains. Several fluoroquinolones demonstrate excellent in vitro activity against B. pertussis (1, 3, 8, 11, 16, 22), and although contraindicated for children (1, 16), they might be candidate agents to treat adults with pertussis.However, as we found six nalidixic acid (NAL)-resistant B. pertussis strains isolated from 2004 to 2006 in Japan by disk diffusion test (no inhibitory zone detected around a 30-μg NAL disk on Bordet-Gengou agar), the MICs of erythromycin, NAL, norfloxacin, ciprofloxacin, sparfloxacin, levofloxacin, gatifloxacin, and chloramphenicol were determined by Etest (AB Biodisk, Solna, Sweden) on Bordet-Gengou agar (7, 10). The quality control strains used were B. pertussis CCUG 30837T (= ATCC 9797T) and Staphylococcus aureus ATCC 29213 (10). Strains for which the MICs of erythromycin and NAL were less than or equal to 0.12 and 32 μg/ml were considered susceptible (4, 10). Whole sequences of gyrA, gyrB, parC, and parE of the NAL-resistant strains were determined by PCR and direct sequencing. To prepare template DNA, one loopful colony of each strain was resuspended in 100 μl of 20 mM Tris-2 mM EDTA buffer (pH 7.5), incubated at 100°C for 15 min, and then centrifuged for 5 min at 23,200 × g. The 50-μl reaction mixture contained 2 μl of the supernatant fluid, 25 pmol of each primer, a 0.4 mM concentration of each deoxynucleoside triphosphate, 2.5 U of TaKaRa LA-Taq polymerase (TaKaRa, Shiga, Japan), and LA-Taq GC buffer I (TaKaRa). After an initial denaturation at 95°C for 5 min, amplification proceeded for 30 cycles of 95°C for 15 s, 60°C for 30 s, and 72°C for 2 min 30 s, with a final 10-min extension at 72°C. Following cleanup with a QIAquick PCR purification column (Qiagen, Valencia, CA), the PCR product was sequenced by using a Big Dye terminator (version 3.1; Applied Biosystems, Foster City, CA). The primers used for PCR and direct sequencing are listed in Table Table11.

TABLE 1.

Primers used in this study
Target and primerSequence (5′ to 3′)PositionaDirection
gyrA
    gyrA1bCTCGGGTTCATCCTTACATA−40 to −21Forward
    gyrA2GCATGGCCACCAACATTC530 to 547Forward
    gyrA3ACTTCATCGCCATCATCA1175 to 1192Forward
    gyrA4TGTACTGGCTGAAGGTGT1829 to 1846Forward
    gyrA5bCCGTGCCAGTCCAGCATTTC2776 to 2757Reverse
    gyrA6ACACCTTCAGCCAGTACA1846 to 1829Reverse
    gyrA7TGATGATGGCGATGAAGT1192 to 1175Reverse
    gyrA8GAATGTTGGTGGCCATGC547 to 530Reverse
gyrB
    gyrB1bATCTGATCCGCGACACAGAT−100 to −80Forward
    gyrB2ATCTTCACCAACATCGAG556 to 573Forward
    gyrB3GCAAGAGCGTGCTGGAAG1223 to 1240Forward
    gyrB4CGAACGCACCAAGGCATC1878 to 1896Forward
    gyrB5bCACGCGTATCGGCCAACGTC2568 to 2539Reverse
    gyrB6GATGCCTTGGTGCGTTCG1896 to 1878Reverse
    gyrB7CTTCCAGCACGCTCTTGC1240 to 1223Reverse
    gyrB8CGCGATGTTGGTGAAGAT573 to 556Reverse
parC
    parC1bGTGATCATTCCCAAGCGCG−128 to −110Forward
    parC2TGCCCGTGATGCTGCTCA518 to 535Forward
    parC3AAGAGTGGGTGGCGTTTC1148 to 1165Forward
    parC4CACCACCATGATCGATCT1806 to 1823Forward
    parC5bGAACACACCGATGTTGACGA2414 to 2395Reverse
    parC6AGATCGATCATGGTGGTG1823 to 1806Reverse
    parC7GAAACGCCACCCACTCTT1165 to 1148Reverse
    parC8TGAGCAGCATCACGGGCA535 to 518Reverse
parE
    parE1bACTTGTCCGTAAAATGTCGG−47 to −28Forward
    parE2GACGTGATCGAAGCACTG445 to 462Forward
    parE3CAAGGGCGTCAAGCTGCT936 to 953Forward
    parE4GATCCACGATATTTCGGT1410 to 1427Forward
    parE5bGTACGCATAGGCGGTGAATG2051to 2032Reverse
    parE6ACCGAAATATCGTGGATC1427 to 1410Reverse
    parE7AGCAGCTTGACGCCCTTG953 to 936Reverse
    parE8CAGTGCTTCGATCACGTC462 to 445Reverse
Open in a separate windowaThe position of each gene of B. pertussis Tohama I (GenBank accession number BX640422) is shown.bThe primers were used for PCR of each gene.The results are shown in Table Table2.2. These six strains were erythromycin sensitive but showed high-level resistance to NAL (MIC, >256 μg/ml) and decreased susceptibilities to fluoroquinolones. All of the NAL-resistant strains showed the same mutation in the quinolone resistance-determining regions (QRDRs; e.g., in Escherichia coli at positions 67 to 122) (18, 21) of gyrA from aspartic acid to glycine at position 87. gyrB, parC, and parE showed no mutations in the NAL-resistant strains (data not shown). Pulsed-field gel electrophoresis analysis (15) of XbaI-digested DNAs showed different patterns among the NAL-resistant strains tested (Fig. (Fig.11).Open in a separate windowFIG. 1.Pulsed-field gel electrophoresis patterns of XbaI-digested DNAs of NAL-resistant (*) and NAL-sensitive B. pertussis strains.

TABLE 2.

Quinolone susceptibilities and gyrA QRDR mutations of B. pertussis strains
StrainDate isolated (day/mo/yr)Patient
LocationMIC (μg/ml)a
NAL resistancedCodon 87e of gyrA (amino acid)
Age (yr)SexNALNAL with PAβNNORCIPSPXLVXGATCHL
CCUG 30837T20.250.1250.0320.0160.0160.0080.5SGAC(D); WTb
BP10126/07/20056MaleOita2NDc0.0320.0160.0080.0160.0080.25SGAC(D); WT
BP10613/10/20050MaleOita1ND0.0640.0080.0040.0080.0040.25SGAC(D); WT
BP10912/12/20050UnknownOita2ND0.0640.0080.0040.0040.0040.125SGAC(D); WT
BP11106/01/20064UnknownOita2ND0.0640.0160.0080.0080.0080.25SGAC(D); WT
BP11222/01/20060MaleOita1ND0.0640.0080.0040.0080.0040.25SGAC(D); WT
BP11308/02/20067MaleOita20.250.0640.0080.0040.0080.0040.5SGAC(D); WT
BP11517/02/20068FemaleOita1ND0.0640.0160.0040.0080.0080.5SGAC(D); WT
BP12805/04/20070MaleOsaka1ND0.0640.0160.0080.0160.0080.25SGAC(D); WT
BP5824/04/20040MaleOsaka>2566410.1250.1250.0640.0320.5RGGC(G)
BP9906/07/20050MaleOita>2563210.1250.0320.0640.0320.25RGGC(G)
BP11722/04/20060MaleOsaka>256640.50.1250.0640.0640.0320.25RGGC(G)
BP11804/05/20064MaleOita>256640.50.1250.0640.0640.0320.5RGGC(G)
BP12116/05/200610FemaleFukuoka>2563210.1250.0640.0640.0320.25RGGC(G)
BP12217/05/20065MaleFukuoka>2563210.1250.0640.0640.0320.25RGGC(G)
Open in a separate windowaNOR, norfloxacin; CIP, ciprofloxacin; SPX, sparfloxacin; LVX, levofloxacin; GAT, gatifloxacin; CHL, chloramphenicol.bWT, wild type.cND, not done.dS, susceptible; R, resistant.eIn each case, the amino acid changed is underlined and the amino acid substituted for it is in parentheses.To our knowledge, this is the first report of quinolone resistance in B. pertussis. All of the strains were isolated from children and were genetically and epidemiologically unrelated. Although fluoroquinolones are not usually prescribed for children, they are widely used to treat respiratory tract infections in adults (3). In the vaccination era, an increasing proportion of pertussis cases occur in adults and adolescents who have lost immunity to B. pertussis, and adult pertussis is a likely source of infant pertussis outbreaks (2, 5, 9, 19). Treatment of unrecognized adult pertussis with fluoroquinolones might, therefore, be important for selecting quinolone-resistant B. pertussis.There are three different mechanisms of quinolone resistance: mutations in drug targets such as DNA gyrase or topoisomerase IV, reduced accumulation of quinolones, and the existence of products that protect the microorganism from the lethal effects of quinolones, such as Qnr (17). The most common mutation observed in quinolone-resistant E. coli is at position 83 of gyrA (18, 21). In our studies, glutamine at position 83 and serine at position 84 are found in B. pertussis strains tested instead of serine at position 83 and alanine at position 84 in E. coli (data not shown). Since these two amino acids are consistent with the gyrA sequences in both the NAL-susceptible and NAL-resistant strains, their amino acids may be conservative and do not affect quinolone resistance in B. pertussis. Substitution of aspartic acid at position 87 is the second most commonly observed mutation in clinical isolates of quinolone-resistant gram-negative and gram-positive microorganisms (18). In this study, aspartic acid to glycine at position 87 of gyrA is the only substitution observed in all of the six NAL-resistant strains compared to NAL-sensitive strains.A few studies have examined the two remaining mechanisms of quinolone resistance in Bordetella and related genera. Kadlec et al. (12) demonstrated that efflux-mediated resistance to NAL in Bordetella bronchiseptica strains was due to a FloR or CmlB1 exporter that could also export chloramphenicol and be inhibited by the efflux pump inhibitor Phe-Arg-β-naphthylamide (PAβN). This study found no differences in the MICs of chloramphenicol against both NAL-resistant and NAL-sensitive B. pertussis strains (Table (Table2).2). PAβN at 80 μg/ml (one-fourth of the MIC) could change the MIC of NAL against NAL-resistant strains but similarly decreased the MIC of NAL against NAL-sensitive strains. The presence of Qnr increases the MICs of fluoroquinolones by 16- to 125-fold but affects the MIC of NAL only 2- to 8-fold (17). These quinolone resistance patterns are quite different from those of our NAL-resistant strains. Therefore, the single QRDR mutation in gyrA may constitute the main mechanism of quinolone resistance among the strains tested here.Once a single mutation of gyrA occurs in gram-negative bacteria, additional mutations in gyrA or parC occur more frequently than in wild-type bacteria (6), and such a stepwise accumulation of multiple mutations increases resistance to fluoroquinolones. Continued surveillance of antimicrobial resistance among B. pertussis strains is clearly needed to control and eradicate B. pertussis transmission.  相似文献   

17.
Using time-kill methodology, we investigated the interactions of an imipenem-colistin combination against 42 genetically distinct Klebsiella pneumoniae clinical isolates carrying a blaVIM-1-type gene. Irrespective of the imipenem MIC, the combination was synergistic (50%) or indifferent (50%) against colistin-susceptible strains, while it was antagonistic (55.6%) and rarely synergistic (11%) against non-colistin-susceptible strains (with synergy being observed only against strains with colistin MICs of 3 to 4 μg/ml). The combination showed improved bactericidal activity against isolates susceptible either to both agents or to colistin.During the past decade, VIM metallo-β-lactamases (MBLs) have spread rapidly among Enterobacteriaceae (4). MBL producers commonly exhibit a multiple-drug resistance phenotype as a result of combined chromosomally encoded or plasmid-mediated resistance mechanisms. Frequently, colistin and tigecycline remain the only therapeutic choices. Tigecycline has demonstrated in vitro activity against MBL producers (18), but evidence of in vivo efficacy against a variety of clinical infections (i.e., bacteremia or pneumonia) is still limited. On the other hand, randomized controlled trials supporting the use of colistin as a single-drug regimen, as well as studies on its pharmacokinetic/pharmacodynamic properties, are lacking (11). Recently, the emergence of colistin resistance among Klebsiella pneumoniae isolates further jeopardized the already limited treatment options in the intensive care unit setting (2). For all these reasons, combination therapies are frequently used in clinical practice, especially in hospitals with high rates of infections by MBL producers.(Some of these data were presented at the 45th Infectious Disease Society Annual Meeting, 2007 [15a].)We investigated the in vitro activities of imipenem and colistin alone and in combination against 42 unique clinical isolates of MBL-producing K. pneumoniae isolated in Greek hospitals from February 2004 to September 2006. MICs were determined by Etest (AB Biodisc, Solna, Sweden) and interpreted according to CLSI breakpoints for imipenem (3) and EUCAST breakpoints for colistin (7). The presence of a blaVIM gene was confirmed by PCR (15). On the basis of PCR-restriction fragment length polymorphism analysis (9), all isolates carried a blaVIM-1-type gene. Extended-spectrum β-lactamase production was detected with a modified CLSI confirmatory test (8). Genetic relatedness among studied isolates was evaluated with repetitive extragenic palindromic PCR methods (10). Patterns that differed by more than one amplification band were characterized as different. In vitro interactions between imipenem and colistin were tested using time-kill methodology. Antibiotic concentrations used were 10 μg/ml for imipenem (Merck, Rahway, NJ) and 5 μg/ml for colistin sulfate (Sigma, St. Louis, MO) because these concentrations represent the steady state achievable in human serum during treatment (12, 16) and thus are clinically relevant. For susceptible strains, if 4× MIC was not higher than 10 or 5 μg/ml for imipenem or colistin, respectively, this concentration was also tested.Synergy was defined as a ≥2-log10 decrease in CFU/ml between the combination and the most active single agent at the different time points, with the number of surviving organisms in the presence of the combination being ≥2 log10 CFU/ml below the number of organisms in the starting inoculum. Antagonism was defined as a ≥2-log10 increase in CFU/ml between the combination and the most active single agent. All other interactions were characterized as indifferent. Bactericidal activity of single antibiotics or combinations was defined as a ≥3-log10 reduction in the CFU/ml of the initial inoculum after 24 h of incubation (1, 6). The lower limit of detection was 1.6 log10 CFU/ml. For analysis of the results, isolates were classified into four groups according to susceptibility to imipenem and colistin. The chi-square test was used to compare proportions of killing activity or synergy between groups by using Yates continuity correction in two-by-two tables. P values of <0.05 were considered statistically significant.The results are shown in Table Table1.1. The imipenem-colistin combination exhibited synergy against 12 of 24 (50%) colistin-susceptible MBL-producing K. pneumoniae isolates tested, but it was antagonistic against 10 of 18 (55.6%) non-colistin-susceptible isolates. Of note, isolates showing colistin MICs of 3 to 4 μg/ml behaved more like colistin-susceptible isolates, since in two of them (50%) a synergistic interaction was noted after exposure to the combination.

TABLE 1.

MICs (μg/ml) of imipenem and colistin against blaVIM-1-type MBL-producing K. pneumoniae isolates and in vitro interaction of the combination
StrainMIC (μg/ml)
Presence of ESBLInteractiona (time of growth [h])No. of isolates showing synergy (or antagonism, if indicated)/total no. of isolates (%)
ImipenemColistin
Imipenem- and colistin-susceptible isolates3/8 (37.5)
    71630.25YesIndifference
    631 CΙ0.750.38YesSynergy (24)b
    2596 II20.5YesSynergy (24)b
    1057 Β ΙΙ1.50.5YesIndifference
    75710.25YesSynergy (3, 5, 24)b
    270 E ΙΙ1.50.25YesIndifference
    235420.25YesIndifference
    1037 E ΙΙ20.38YesIndifference
Non-imipenem-susceptible and colistin-susceptible isolates9/16 (56.3)
    350 I60.5NoSynergy (1, 24)c
    498 II80.3YesSynergy (24)c
    1587 I>320.5YesSynergy (24)c
    266 E>320.4YesIndifference
    1526240.4NoSynergy (24)c
    760 C>320.4NoSynergy (24)c
    329 Β Ι>320.2YesSynergy (24)c
    175 ΙΙΙ>320.3YesIndifference
    4412 Β ΙΙ>320.5YesIndifference
    377 II>320.5YesSynergy (5c or 5 and 24d)
    513 E I>320.38YesIndifference
    682 E I>320.19YesSynergy (24)c,d
    735 E II>320.38YesIndifference
    1437 B II>320.5YesSynergy (24)c
    761 E I>320.2YesIndifference
    332 E>320.25YesIndifference
Non-imipenem-susceptible and non-colistin-susceptible isolates7/15 (46.7) (antagonism), 2/15 (13.3) (synergy)
    748 A IV>3248YesIndifference
    231 D>3248YesAntagonism (3, 5)d
    1171 C II12256YesIndifference
    1014 A I>3296YesAntagonism (5)d
    1459>32256YesAntagonism (3, 5)d
    1057 A>3248YesAntagonism (5)d
    1326 A>3296YesAntagonism (3, 5)d
    712 B I2416NoIndifference
    1478 C I8128YesAntagonism (3, 5)d
    4090 B1232YesIndifference
    1110 B II664YesAntagonism (24)d
    674 C II>323NoSynergy (3, 5, 24)d
    963 II>324NoIndifference
    1919>324YesSynergy (24)d
    680 A>324NoIndifference
Imipenem-susceptible and non-colistin-susceptible isolates3/3 (100) (antagonism)
    1119248YesAntagonism (3,e 5d)
    318 G Ι324YesAntagonism (24)d
    240 Β Ι464YesAntagonism (24)d
Open in a separate windowaAll combinations tested at all time points exhibited indifference unless otherwise specified.bConcentrations tested were as follows: imipenem and colistin, 4× MIC.cConcentrations tested were as follows: imipenem, 10 μg/ml; and colistin, 4× MIC.dConcentrations tested were as follows: imipenem, 10 μg/ml; and colistin, 5 μg/ml.eConcentrations tested were as follows: imipenem, 4× MIC; and colistin, 5μg/ml.The combination was rapidly bactericidal against all isolates susceptible to both agents (n = 8) compared to imipenem and colistin alone (4× MIC), which were bactericidal against two and three isolates, respectively (P < 0.05). In the subgroup of 16 isolates that were non-imipenem-susceptible and colistin susceptible, the combination of imipenem (10 μg/ml) and colistin (4× MIC) was bactericidal against 10 (62.5%) isolates, while another combination of imipenem (10 μg/ml) and colistin (5 μg/ml) was bactericidal against 12 (75%) isolates compared to imipenem (10 μg/ml) and colistin (4× MIC) alone, which were bactericidal against zero and two isolates, respectively (P < 0.05). The antibiotic combination was bactericidal against only 2 of 15 (13.3%) isolates that were nonsusceptible to both imipenem and colistin. In the subgroup of isolates that were susceptible to imipenem but nonsusceptible to colistin (n = 3), the combination exhibited an antagonistic effect, and regrowth was noted for all isolates after 24 h of incubation. Overall, in the group of imipenem-susceptible isolates, imipenem alone at a concentration of 10 μg/ml or 4× MIC demonstrated killing activity against 7/11 (63.6%) or 2/8 (25%) isolates, respectively, at 24 h.In order to evaluate the development of resistance as a reason for bacterial regrowth after 24 h of incubation with the studied combination, viable colonies were subjected to susceptibility testing in comparison with the respective wild-type strain, using agar dilution as described by CLSI (3). This evaluation was performed only for isolates that were initially susceptible to at least one of the tested antimicrobials.For 7 of 12 isolates (58.3%) that were initially susceptible to colistin, a colistin-resistant clone (MIC range, 64 to >256 μg/ml) was selected after incubation with the tested combination. Conversely, among four isolates initially susceptible to imipenem that showed regrowth after 24 h of incubation with the combination, none developed resistance (MIC range, 1 to 4 μg/ml).To our knowledge, the present study is the first to assess the in vitro interaction of imipenem and colistin against a large number of VIM-1-type MBL-producing K. pneumoniae isolates exhibiting a wide range of susceptibilities to these agents. Carbapenem resistance levels of MBL-positive Enterobacteriaceae are variable and often below the proposed resistance breakpoint (20) as a result of differences in outer membrane permeability or in the levels of VIM-1 production (13), but most experts recommend against the use of carbapenem monotherapy for treatment (4, 17), based on evidence of a strong inoculum effect in vitro (14). Other experimental data suggested that an increased imipenem dosage could be efficacious against susceptible isolates (5). In the era of multidrug resistance, our findings concerning the killing activity of imipenem as a single agent against selected susceptible MBL-producing strains merit further investigation. Importantly, the imipenem-colistin combination demonstrated improved bactericidal activity compared to either agent alone and yielded synergy against 14 of 42 (33.3%) K. pneumoniae isolates tested. Synergy was observed only against isolates exhibiting susceptibility or low-level resistance to colistin. In contrast, antagonism was observed against 10 of 42 (23.8%) strains tested, all of which exhibited high-level resistance to colistin. These differences underscore the importance of accurate susceptibility testing of colistin, with MIC determination. In concordance with these findings, the previous experiences of our group suggest that colistin-containing regimens are successful for the treatment of infections by VIM-1-type MBL-producing Enterobacteriaceae (19). The results of the present study merit further investigation in animal models and clinical trials. While waiting for these data, the coadministration of imipenem and colistin should probably be avoided for colistin-resistant VIM-producing K. pneumoniae because it could result in antagonism.  相似文献   

18.
A total of 489 clinical isolates of Pseudomonas aeruginosa was investigated for metallo-β-lactamase (MBL) production. Molecular analysis detected a blaVIM-1 gene in the chromosome of one isolate and a blaVIM-2 gene carried on the plasmid in seven isolates. Moreover, we showed that an initial screening by combined susceptibility testing of imipenem and ceftazidime followed by a confirmatory EDTA combination disk test represents a valid alternative to the molecular investigation of MBL genes, making MBL detection possible in routine diagnostic laboratories.Metallo-β-lactamase (MBL)-producing Gram-negative bacteria are an increasing public health problem worldwide because of their resistance to all β-lactams except aztreonam (3). MBL genes are typically part of an integron and are either carried on transferable plasmids or are part of the bacterial chromosome (28). The most common transferable MBL families include the VIM-, IMP-, GIM-, SPM-, and SIM-type enzymes which have been detected primarily in Pseudomonas aeruginosa but were also found in other Gram-negative bacteria, including nonfermenters and members of the family Enterobacteriaceae (22). Recently, two new subgroups of MBLs, designated NDM-1 and DIM-1, were identified in a clinical isolate of Klebsiella pneumoniae in India and in a clinical isolate of Pseudomonas stutzeri in the Netherlands, respectively (21, 31).In most studies, reduced susceptibility to imipenem has been adopted as the sole criterion for further phenotypic or molecular investigations in order to detect MBLs (11, 19, 20, 23). However, this criterion seems to be suboptimal, as it does not allow exclusion of isolates characterized by the loss of the OprD porin, the most common mechanism of resistance to imipenem in P. aeruginosa (14, 22). Moreover, several phenotypic tests for detecting MBLs, such as the MBL Etest (20, 25), EDTA combination disk test (20, 25, 30), EDTA disk synergy test (10), and imipenem lysate MBL assay (27), have been developed and evaluated. However, the performance of each of these tests seems to be strongly affected by the local rate of MBL-producing isolates.In Germany, VIM-, and GIM-type enzymes in P. aeruginosa isolates have already been detected (1, 8, 26). Recently, Elias et al. described a nosocomial outbreak caused by a blaVIM-2-positive P. aeruginosa in patients of the Department of Urology of the university hospital of the University of Würzburg, Würzburg, Germany, between November and December 2007 (4). However, to the best of our knowledge, no systematic surveys of the occurrence of MBLs in clinical isolates of P. aeruginosa have been conducted in Germany so far.Accordingly, this study was designed with the following aims: (i) to develop improved screening criteria for the detection of MBL-producing P. aeruginosa; (ii) to determine the proportion of MBL-producing isolates in clinical isolates of P. aeruginosa in the university hospital of the University of Würzburg in Germany; (iii) to assess the relatedness of MBL-producing isolates; (iv) to determine the locations of the MBL genes detected; and (v) to evaluate two phenotypic tests as confirmatory tests for the detection of MBL production.Since no standard imipenem MIC breakpoints for MBL producers are available, we first analyzed the MICs of imipenem in 10 well-characterized P. aeruginosa control strains shown previously to produce IMP-, VIM-, GIM-, SIM-, and SPM-type enzymes (Table (Table1).1). Identification to the species level was confirmed using Vitek 2 GN cards (bioMérieux, Nürtingen, Germany). Antimicrobial susceptibility testing was carried out with Vitek 2 AST-N021 and AST-N110 cards (bioMérieux). MICs were interpreted as recommended by the CLSI (2). All MBL-producing positive-control strains were intermediate sensitive or resistant to imipenem (MIC ≥ 8 μg/ml). Subsequently, we investigated 26 consecutive nonreplicate clinical isolates of P. aeruginosa with an imipenem MIC of ≥8 μg/ml for the presence of MBL genes by multiplex PCR as described by Ellington et al. (5). All clinical isolates were collected in our laboratory throughout 2007 before the outbreak at the university hospital of the University of Würzburg, Würzburg, Germany (4), and all isolates were shown by multiplex PCR to be negative for MBL genes. Since the loss of the OprD porin, the most common mechanism of resistance to imipenem in P. aeruginosa, does not confer ceftazidime resistance (15), we analyzed the ceftazidime MICs in the MBL-producing positive-control strains and in the 26 MBL-negative isolates. While all 10 MBL-producing positive-control strains were resistant to ceftazidime (MIC ≥ 32 μg/ml), only 6 of the 26 MBL-negative isolates were resistant to ceftazidime (P < 0.01) (Table (Table1).1). Consequently, we adopted MIC breakpoints for the initial screening of MBL producers of ≥8 μg/ml for imipenem and ≥32 μg/ml for ceftazidime.

TABLE 1.

Comparison of imipenem and ceftazidime MICs in well-characterized MBL-producing positive-control strains and in MBL-negative clinical isolatesa
Strain or isolatebMBL enzymecGeographic originReferenceMIC (μg/ml)
ImipenemCeftazidimed
Control strains
    PA431IMP-1Turkey178≥64
    PA552IMP-1UK5≥16≥64
    PA386IMP-13Italy18≥16≥64
    PA550VIM-1UK5≥16≥64
    PA373VIM-2Italy9≥16≥64
    PA399VIM-2Germany4≥16≥64
    PA430VIM-4Hungary13≥16≥64
    PA554GIM-1Germany1≥16≥64
    AB551SIM-1Korea12≥16≥64
    PA553SPM-1Brazil6≥16≥64
Clinical isolates
    PA339GermanyThis study≥168
    PA340GermanyThis study84
    PA341GermanyThis study≥16≥64
    PA342GermanyThis study≥164
    PA346GermanyThis study≥164
    PA349GermanyThis study≥168
    PA350GermanyThis study≥164
    PA352GermanyThis study≥164
    PA355GermanyThis study≥164
    PA356GermanyThis study≥164
    PA360GermanyThis study816
    PA361GermanyThis study≥16≥64
    PA362GermanyThis study832
    PA364GermanyThis study84
    PA368GermanyThis study≥164
    PA370GermanyThis study≥168
    PA376GermanyThis study≥1616
    PA377GermanyThis study≥168
    PA380GermanyThis study84
    PA381GermanyThis study≥164
    PA382GermanyThis study≥164
    PA395GermanyThis study≥16≥64
    PA734/CFGermanyThis study≥164
    PA736/CFGermanyThis study8≥64
    PA758/CFGermanyThis study≥164
    PA774/CFGermanyThis study≥16≥64
Open in a separate windowaTen well-characterized MBL-positive strains (9 P. aeruginosa strains and one Acinetobacter baumanii strain) (control strains) and 26 MBL-negative P. aeruginosa isolates with reduced susceptibility to imipenem (clinical isolates) are compared.bP. aeruginosa strains and isolates are indicated by PA at the beginning of the strain or isolate designation, and P. aeruginosa isolates from patients with cystic fibrosis are indicated by CF after a slash at the end of the isolate designation. One Acinetobacter baumanii (AB) control strain is shown.cThe type of MBL enzyme is given for the control strains that produce MBL. The clinical isolates were MBL negative (−).dAll 10 MBL-producing positive-control strains were resistant to ceftazidime (MIC ≥ 32 mg/ml), but only 6 of the 26 MBL-negative clinical isolates were resistant to ceftazidime (P < 0.01 by Fisher''s exact test).From June 2008 until May 2009, a total of 489 consecutive nonreplicate isolates of P. aeruginosa from diverse clinical specimens were screened for MBL production. Sixty-eight of these isolates (13.9%) showed reduced susceptibility to imipenem (MIC ≥ 8 μg/ml). Adding resistance to ceftazidime (MIC ≥ 32 μg/ml) as an additional screening criterion for MBL production reduced the number of isolates to be consecutively tested by multiplex PCR (5) to 15. Following PCR, sequencing of the purified amplicons (QIAquick PCR purification kit; Qiagen, Hilden, Germany) was performed with an ABI 3130 genetic analyzer (Applied Biosystems, Foster City, CA). Molecular analysis revealed a blaVIM-1 gene in one isolate and a blaVIM-2 gene in seven additional isolates. These results were confirmed by amplification and sequencing of the entire VIM gene of isolates PA500 (blaVIM-1) and PA399 (blaVIM-2) by using the primer pairs VIM-F (5′-GTTATGCCGCACTCACCCCCA-3′)/VIM-R (5′-TGCAACTTCATGTTATGCCG-3′) (29) and VIM2004A/VIM2004B (20). Furthermore, we could demonstrate the association of the MBL genes with a class 1 integron by PCR using the integron-specific primers 5-CS and 3-CS in combination with the MBL-specific primers VIM2004A and VIM2004B (20). The clinical origins and antimicrobial susceptibilities of the MBL-positive isolates are shown in Table Table22.

TABLE 2.

Epidemiological data and resistance phenotypes of all P. aeruginosa isolates with reduced susceptibility to imipenem and resistance to ceftazidime
IsolateSpecimenDate of recoveryWardaMBL enzymebMIC (μg/ml)c
PIPTZPCAZFEPATMIMPMEMCOLAMKGENTOBCIP
MBL-producing isolates
    PA399UrineNovember 2007UrologyVIM-2≥128≥128≥64≥6416≥16≥1618≥16≥16≥4
    PA462Ascitic fluidJune 2008General surgeryVIM-2≥128≥128≥64≥6416≥16≥1614≥16≥16≥4
    PA465DrainageJuly 2008General surgeryVIM-2≥12864≥64≥6416≥16≥1628≥16≤1≥4
    PA469WoundAugust 2008General surgeryVIM-2≥128≥128≥64≥6416≥16≥1618≥16≥16≥4
    PA475WoundSeptember 2008General surgeryVIM-2≥128≥128≥64≥6416≥16≥1614≥16≥16≥4
    PA477DrainageSeptember 2008General surgeryVIM-2≥128≥128≥64≥6416≥16≥1618≥16≥16≥4
    PA481UrineOctober 2008UrologyVIM-2≥128≥128≥64≥6416≥16≥1618≥16≥16≥4
    PA500Tracheal secretionFebruary 2009Surgical ICUVIM-1≥128≥128≥64≥644≥16≥161≤2≥168≥4
    PA510Tracheal secretionMay 2009Medical ICUVIM-2≥128≥128≥64≥64≥64≥16≥1614≥16≥16≥4
MBL-negative isolates
    PA459Tracheal secretionJune 2008Surgical ICU≥128≥12832816≥16812110.25
    PA460WoundJune 2008General surgery≥128≥128321616≥16814≥16≥16≥4
    PA479UrineOctober 2008General surgery≥128≥128≥64≥64328≥1618≥16≥16≥4
    PA483WoundOctober 2008Surgical ICU≥128≥128≥643264≥16≥1612110.25
    PA494UrineDecember 2008Neurosurgery≥128≥12832816≥1681881≥4
    PA507DrainageApril 2009Surgical ICU≥128≥128321616≥164116≥1610.25
    PA509WoundMay 2009Internal medicine≥128≥12832168≥16≥16264≥16≥16≥4
Open in a separate windowaICU, intensive care unit.bThe type of MBL enzyme is given for the isolates that produce MBL. Some isolates did not produce MBL (−).cAbbreviations for antimicrobial agents: PIP, piperacillin; TZP, piperacillin-tazobactam; CAZ, ceftazidime; FEP, cefepime; ATM, aztreonam; IMP, imipenem; MEM, meropenem; COL, colistin; AMK, amikacin; GEN, gentamicin; TOB, tobramycin; CIP, ciprofloxacin.On the basis of these results, the proportion of isolates producing MBL was 1.6% with regard to all P. aeruginosa isolates investigated and 11.7% with regard to the isolates with reduced susceptibility to imipenem. These data are in accordance with the MBL-producing isolate proportions recently reported for, e.g., Italy and Korea. However, we observed only the presence of the VIM-type MBLs, not the IMP-type MBLs, which are also commonly detected worldwide (11, 23).All P. aeruginosa isolates with reduced susceptibility to imipenem and resistance to ceftazidime were analyzed for genomic relatedness by a randomly amplified polymorphic DNA (RAPD) typing technique using primers 208 and 272 (16). From the banding pattern, a dendrogram using the Ward clustering algorithm was generated based on the Dice coefficients with 0.5% optimization and 0.5% position tolerance for band matching and comparison using the GelComparII software program (Applied Maths, Sint-Martens-Latem, Belgium). RAPD typing with primer 208 revealed that all blaVIM-2-positive isolates clustered together (Fig. (Fig.1).1). The cluster also included the first isolate of the outbreak at the university hospital of the University of Würzburg (PA399) (4). These results were confirmed by RAPD typing with primer 272 (data not shown) and suggest a common clonal origin of all blaVIM-2-positive isolates.Open in a separate windowFIG. 1.Genotypic comparison of imipenem-nonsusceptible and ceftazidime-resistant isolates based on the RAPD profile using primer 208. The isolate designation, presence and type of MBL, and presence (+) or absence (−) of a plasmid in the P. aeruginosa isolate is shown to the right of the RAPD profile. The clusters are shown to the left of the RAPD profile.Plasmid extraction was attempted from all imipenem-nonsusceptible and ceftazidime-resistant isolates by the alkaline lysis method (24). Extracted plasmid DNA was subsequently subjected to 0.7% agarose gel electrophoresis, followed by ethidium bromide staining, which revealed the presence of a plasmid in all blaVIM-2-positive isolates but not in isolate PA500, which therefore carried the blaVIM-1 gene on its chromosome. To further characterize the genetic location of the blaVIM-2 genes by the Southern blot technique, total DNA from all imipenem-nonsusceptible and ceftazidime-resistant isolates was subjected to pulsed-field gel electrophoresis (PFGE) using a modified version of the protocol obtained from the home page of the Health Protection Agency of the United Kingdom (http://www.hpa.org.uk/) and hybridized using digoxigenin (DIG)-labeled probes for blaVIM-2 and 16S rRNA gene. The probes were generated using the DIG-DNA labeling kit (Roche Diagnostics, Mannheim, Germany) with the purified PCR product amplified with the primer pair VIM2004A and VIM2004B (20) for blaVIM-2 and primer pair pc3 and bak for the 16S rRNA gene (7). Hybridization with the blaVIM-2-specific gene probe revealed a band about 160 kb in size in all blaVIM-2-positive isolates, corresponding to a plasmid carrying blaVIM-2. The presence of the plasmid was further confirmed by large-scale preparation of plasmid DNA from isolates PA399 and PA462 using CsCl density gradient ultracentrifugation (24), followed by PFGE of the purified plasmids for size determination.Furthermore, we successfully performed the conjugative transfer of the blaVIM-2 gene from isolate PA462 (MBL positive), which showed meropenem and amikacin MICs of ≥16 and 4 μg/ml, respectively, to isolate PA507 (MBL negative), which showed meropenem and amikacin MICs of 4 and 16 μg/ml, respectively. In detail, bacterial suspensions of both isolates in brain heart infusion (BHI) were adjusted to a McFarland standard of 0.5, and 3 ml of each suspension was mixed together and incubated at 37°C for 2 h without shaking. Transconjugant selection was performed on Mueller-Hinton (MH) agar plates containing 6 μg/ml each of meropenem and amikacin. The transconjugant was positive for blaVIM-2 as revealed by PCR with primers VIM2004A and VIM2004B. Moreover, RAPD typing using primers 208 and 272 and colony morphology on Mueller-Hinton agar revealed that the transconjugant and the MBL-negative isolate PA507 were geno- and phenotypically closer to each other than to isolate PA462, which clearly demonstrated that the plasmid carrying blaVIM-2 was transferred to the MBL-negative isolate PA507. The molecular analyses thus demonstrate the spread of a clone of P. aeruginosa harboring blaVIM-2 as part of a class 1 integron on a large conjugative plasmid in the geographically and temporarily restricted setting of a German university hospital.In addition, all isolates with reduced susceptibility to imipenem and resistance to ceftazidime were also subjected to a phenotypic analysis by MBL Etest (AB Biodisk, Solna, Sweden) and EDTA combination disk test. The EDTA combination disk test was performed as previously described (20) using antibiotic disks (Oxoid, Wesel, Germany) containing 10 μg imipenem alone and in combination with 930 μg EDTA. The MBL Etest correctly identified all MBL-positive isolates but falsely identified six of the seven MBL-negative P. aeruginosa isolates as MBL positive. In contrast, the EDTA disk test was able to discriminate between all MBL-positive and MBL-negative isolates by using a breakpoint of ≥14 mm. Of note, the same test interpreted with a breakpoint of ≥7 mm as suggested by Pitout et al. (20) falsely identified all MBL-negative isolates as MBL positive. These data therefore suggest that the optimal breakpoint may depend on the strain collection studied. An initial screening by combined susceptibility testing of imipenem and ceftazidime, followed by a confirmatory EDTA combination disk test, thus represents a valid and less expensive alternative to the molecular investigation of MBL genes. This aspect is particularly important, as it makes MBL detection possible not only in reference laboratories but also in routine diagnostic microbiology laboratories.  相似文献   

19.
During an island-wide PCR-based surveillance study of beta-lactam resistance in multidrug-resistant (MDR) Escherichia coli, Klebsiella pneumoniae, Pseudomonas aeruginosa, and Acinetobacter calcoaceticus-baumannii complex isolates obtained from 17 different hospitals, 10 KPC-positive Acinetobacter isolates were identified. DNA sequencing of the blaKPC gene identified KPC-2, -3, and -4 and a novel variant, KPC-10. This is the first report of a KPC-type beta-lactamase identified in Acinetobacter species.The Acinetobacter calcoaceticus-baumannii complex has been recognized, during the last few decades, as an important opportunistic pathogen associated with life-threatening nosocomial infections and hospital outbreaks, as well as an agent of serious infections in injured U.S. military personnel returning from the Middle East war zones (9, 25, 33). In addition to their ability to survive in adverse environmental conditions (14), the organisms have intrinsic and acquired mechanisms of antimicrobial resistance, such as porin downregulation, overexpression of efflux pumps, chromosomal and plasmid-acquired beta-lactamases, aminoglycoside-modifying enzymes, and fluoroquinolones resistance, among others (11). Class B metallo-beta-lactamases and class D oxacillinase with carbapenemase activity have been identified in Acinetobacter species as a mechanism of broad-spectrum beta-lactam resistance (11, 20).The molecular class A beta-lactamases of the KPC family are a group of potent carbapenemases identified initially in a Klebsiella pneumoniae isolate from the United States and later in other members of the Enterobacteriaceae family and in other geographical regions worldwide (17, 20). Pseudomonas aeruginosa positive for the blaKPC gene has been recently identified in Colombia, Puerto Rico, and Trinidad and Tobago (3, 28, 29).Up to date, eight different KPC variants (KPC-2 to -9) have being identified differing by 1 or 2 two amino acid substitutions. KPC-2 and -3 are the most common variants identified in Enterobacteriaceae and P. aeruginosa. KPC-6, -7, and -8 have been identified only in K. pneumoniae, while KPC-9 was detected in Escherichia coli and KPC-5 in P. aeruginosa. All the KPC variants except for KPC-7 and -9 have been detected in Puerto Rico (21-24, 29). In this report, we describe for the first time the presence of the KPC gene in clinical isolates of Acinetobacter species in Puerto Rico and the identification of a novel KPC variant, KPC-10, in one of the isolates.A PCR-based surveillance study of beta-lactam resistance was started in January 2009 in 17 hospitals across the island. After Institutional Review Board approval, the participating hospitals sent isolates of all unique, consecutive, multidrug-resistant Acinetobacter species with their corresponding susceptibility report and basic epidemiologic information. PCR screening with family-specific beta-lactamase primers for KPC and IMP and multiplex PCR for OXA carbapenemases and for CTX-Ms were performed as previously described (29, 31, 32). The Vitek DCS-R5 system confirmed the identification of the isolates as those in the A. calcoaceticus-baumannii complex. Bidirectional sequencing of the full-length blaKPC gene PCR product was independently generated at least twice to identify the type of the KPC variant. DNA sequencing was commercially performed by Davis Sequencing. Sequence alignment and analysis were done online by utilizing the BLAST program (www.ncbi.nlm.nih.gov). Trek GNXF Gram-negative MIC plate microdilution panels (Westlake, OH) were utilized to perform antibiotic susceptibility tests by following the manufacturer''s instruction, and the results were interpreted as recommended by the Clinical and Laboratory Standards Institute (8). No attempts were made to evaluate patients'' therapies or clinical outcomes.From a total of 274 multidrug-resistant Acinetobacter isolates collected from January to May of 2009, 10 (3.4%) were identified as KPC positive. Seven were detected in the metropolitan San Juan area, two in the north, and one in the central region of Puerto Rico.Table Table11 shows the patients'' pertinent clinical information and type of KPC variants identified. There were six male and four female patients, whose average age was 69 years, ranging from 40 to 95 years. Six were isolated from patients in the intensive care units either from sputum (four isolates) or blood (two isolates) samples. The average hospital length of stay was 41.6 days (range, 11 to 96 days). Six patients had ventilator-associated pneumonia with or without sepsis. Six patients had skin and soft-tissue infections. All patients had significant comorbid conditions, such as cardiovascular, renal, neurologic, or traumatic injuries. The crude mortality rate for the 10 patients was 40%. Nine patients were treated with multiple antibiotic courses during their hospitalization (data not shown). Our patients shared similar characteristics to those described in the literature, such as significant comorbidities, prolonged hospitalizations, invasive procedures, exposure to multiple antibiotics regimes, and high crude mortality rates (1, 2, 10, 27). In our island-wide surveillance study, all hospitals with KPC-positive Acinetobacter isolates also had KPC-positive Klebsiella pneumoniae, Pseudomonas aeruginosa, and/or Escherichia coli isolates (21).

TABLE 1.

Pertinent clinical information and type of KPC variants
Acinetobacter isolateHospaHosp unitbGeographic areaSample siteAge (yr)GenderdDate of (mo/day/yr):
Length of stay (days)Diagnosise
OutcomeKPC variantPCR OXAf
AdmissionSpecimen collectionRelatedOther
MC1AC9-26MC1Gen wardMetrocBlood73M3/11/093/22/0938VAP, sepsis, UTIAMI, ARF, RF, ICHDeathKPC-351
MC4AC9-30MC4ICUMetroSputum40M3/14/094/7/0954VAP, UTI, skin abscessGun-shut woundDischargeKPC-251
M3AC9-2M3Gen wardMetroUrine68F1/19/091/19/099Skin ulcerCVA, dementiaDeathKPC-351
M3AC9-4M3ICUMetroSputum73F1/6/091/22/0996VAP, UTI, skin UlcerAM1, CVA, RFDischargeKPC-351
M3AC9-6M3ICUMetroBlood52M12/27/081/28/0932VAP, sepsisCHF, CRFDeathKPC-351
M3AC9-7M3ICUMetroBlood58F1/18/092/7/0974Catheter sepsisDVT, osteomyelitisDischargeKPC-351
M2AC9-31M2Gen wardMetroSkin wound95M3/23/093/24/0911Skin ulcerDementia, seizuresDischargeKPC-451
N2AC9-18N2ICUNorthSputum58M3/5/094/2/0949VAP, sepsis, skin ulcerRFDeathKPC-351
N2AC9-19N2Gen wardNorthSkin wound88M3/30/094/1/0914Skin ulcerGastric CA, anemiaDischargeKPC-351, 58
C1AC9-24C1ICUCentralSputum88F3/20/093/31/0939VAPRF, intestinal obstructionDischargeKPC-1051
Open in a separate windowaHosp, hospital.bGen Ward, general ward; ICU, intensive care unit.cMetro, metropolitan area.dM, male; F, female.eVAP, ventilator-associated pneumonia; UTI, urinary tract infection; AMI, acute myocardial infarct; ARF, acute renal failure; RF, respiratory failure; ICH, intracerebral hemorrhage; CVA, cerebrovascular accident; CHF, congestive heart failure; CRF, chronic renal failure; and DVT, deep vein thrombophlebitis.fOXA type found by multiplex PCR.The susceptibility tests shown in Table Table22 showed that colistin, polymyxin B, tigecycline, and minocycline had the lowest MICs of the tested antibiotics. Resistance to quinolones, aminoglycosides, and to the beta-lactam antibiotics, including carbapenems, was observed. The therapeutic options available for the treatment of multidrug-resistant Acinetobacter are limited (5, 15). Our results, as well as others (4, 7, 13, 26, 34), suggested that tigecycline, polymyxins, colistin, and possibly minocycline have consistent in vitro activity against this organism.

TABLE 2.

MICs of 10 KPC-positive Acinetobacter calcoaceticus-baumannii complex isolates to selected antibiotics
Test agentaRange (μg/ml)MIC (μg/ml)
C1AC9-24M2AC9-31M3AC9-2M3AC9-4M3AC9-6M3AC9-7MC1AC9-26MC4AC9-30N2AC9-18N2AC9-19W2AC8-S10dE. coli ATCC 25922P. aeruginosa ATCC 27853
CTX1-32>32>32>32>32>32>32>32>32>32>324116
CAZ1-16>16>16>16>16>16>16>16>16>16>16212
FEP2-16>16>16>16>16>16>16>16>16>16>1622<2
TZP8/4-64/4>64/4>64/4>64/4>64/4>64/4>64/4>64/4>64/4>64/4>64/48/48/4<8/4
DORb0.12-2>2>2>2>2>2>2>2>2>2>20.120.120.5
IPM1-848/I>8>8>8>888>8>8112
MEM1-84>8>8>8>8>8>8>8>8>811<1
CIP0.25-2>2>2>2>2>2>2>2>2>2>20.250.250.25
LVX1-8>8>8>8>8>8>8>8>8>8>811<1
GEN1-8>8>82>82<18>8>8>811<1
AMK4-32>32>32>32>32>32>32>32>32>32>3244<4
SXT0.5/9.5-4/76>4/76>4/76>4/76>4/76>4/76>4/76>4/76>4/76>4/76>4/760.5/9/.50.5/9.5>4/76
MIN2-168<2<24<2<2<24<2<22216
TGCc0.25-8110.50.250.50.520.512<0.250.258
CST0.25-410.50.51110.50.50.50.50.250.251
PMB0.25-40.5<0.25<0.25<0.250.50.5<0.250.5<0.25<0.250.50.251
Open in a separate windowaCTX, cefotaxime; CAZ, ceftazidime; FEP, cefepime; TZP, piperacillin-tazobactam; DOR, doripenem; IPM, imipenem; MEM, meropenem; CIP, ciprofloxacin; LVX, levofloxacin; GEN, gentamicin; AMK, amikacin; SXT, trimethoprim-sulfamethoxazole; MIN, minocycline; TGC, tigecycline; CST, colistin; and PMB, polymyxin B.bDoripenem breakpoint value obtained from Doribax package insert.cNo breakpoints available for Acinetobacter species.dA. calcoaceticus-baumannii complex strain W2AC8-S10 is a KPC-negative, OXA-51-positive susceptible isolate.The PCR results obtained with family-specific beta-lactamase primers showed that all isolates were positive for blaKPC and blaOXA-51-like groups and negative for blaIMP and blaCTX-M genes. Sequencing of blaKPC detected the following variants: KPC-3 in seven isolates; KPC-4, KPC-2, and a novel KPC-10 (GenBank accession number GQ140348) in one isolate each. The identification of a class D blaOXA-51-like gene in all isolates was expected, since it had been shown to be chromosomally located in A. baumannii (11, 20). One isolate had the blaOXA-58-like group, which has been shown to be carried on a plasmid, recovered from diverse geographic regions, and associated with nosocomial outbreaks (6, 12, 18, 19).To our knowledge, this is the first report of multidrug-resistant A. calcoaceticus-baumannii complex clinical isolates harboring the KPC gene. The presence of this gene suggests the possibility of horizontal transmission, as this carbapenemase has been associated with mobile genetic elements (transposons) which can be transferred from one bacterium to another (11, 16, 17, 30).The presence of the blaKPC gene in Acinetobacter species adds another important element to an organism already harboring multiple innate and acquired mechanisms of resistance with the real possibility of horizontal transfer of a very troublesome and potent carbapenemase. This clearly emphasizes the importance of judicious use of antibiotics and strict infection and environmental control practices in acute and chronic care facilities to reduce the possibility of nosocomial transmission and potential outbreaks.  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