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 93 毫秒
1.
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   

2.
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   

3.
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   

4.
Thirteen human subjects scheduled for elective anterior segment eye surgery received hourly 2% voriconazole eye drops 4 hours presurgery. No side effects were reported. Significantly, the voriconazole concentration in the aqueous humor of the eye was similar to that reported for the 1% voriconazole solution, suggestive of concentration-independent absorption.Fungal keratitis accounts for 6 to 50% of all ocular infections and is one of the most difficult infections to treat (13). Seventy percent of fungal keratitis infections are attributed to Candida albicans, Aspergillus fumigatus, and some Fusarium species (2). At the Royal Victorian Eye and Ear Hospital (RVEEH), Melbourne, Victoria, Australia, Candida albicans is the causative fungus in 37% of the treated cases (2).Voriconazole (Vfend) is effective (Table (Table1)1) against common pathogens associated with fungal eye disease (i.e., Candida, Aspergillus, and Fusarium species [3, 6, 8]) and some less common keratitis-causative fungi such as Paecilomyces spp., Histoplasma spp., Scedosporium spp., Curvularia spp., and Acremonium (3, 8). The typical corneal fungal pathogens (8) are shown in Table Table1.1. Voriconazole is available commercially in oral and intravenous forms (Vfend package insert; Pfizer Inc., New York, NY). Case reports relating to the use of topical 1% voriconazole for ocular fungal infections are promising but limited (8). A recent study by Lau et al. reported that the concentration of voriconazole in the aqueous humor (after topical administration with 1% voriconazole solution) was sufficiently high to treat some common types of fungal infections but may be inadequate for common but less-sensitive keratitis-causative fungi (12).

TABLE 1.

In vitro MICs at which 90% of isolates are inhibited by voriconazole
OrganismMIC90 (mg/liter)Reference
Yeast and yeastlike species
    Candida albicans0.0616
    Candida parapsilosis0.12-0.2516
    Candida tropicalis0.25->16.016
    Cryptococcus neoformans0.06-0.2516
Moniliaceous molds
    Aspergillus fumigatus0.5016
    Aspergillus flavus0.5016
    Fusarium species0.25-88
    Fusarium solani210
    Paecilomyces lilacinus0.5016
    Acremonium alabamensis0.2516
Dimorphic fungi
    Blastomyces dermatitidis0.2516
    Coccidioides immitis0.2516
    Histoplasma capsulatum0.2516
    Penicillium marneffei0.0316
Dematiaceous fungi
    Cuvularia species0.06-0.2516
    Scedosporium species0.516
    Scedosporium apiospermum0.516
Open in a separate windowTo date, there has only been one case report on the topical application of 2% voriconazole solution (14). The solution was used in combination with intravenous voriconazole and was successful against Fusarium solani keratitis. Importantly, the extent of the role played by the 2% solution in the reported case and the extent of penetration of topically applied 2% voriconazole solution into the aqueous humor remain unknown.The aim of the current study was to investigate the penetration of 2% voriconazole eye drops into the aqueous humor as a potential alternative therapy for the management of ophthalmic fungal keratitis.(Preliminary data of the work described in this paper were presented at the 13th International Congress on Infectious Diseases, Kuala Lumpur, Malaysia, June 2008.)The current study was an open-label prospective study conducted between July 2007 and August 2008 at the RVEEH. The study was approved by the human ethics committee of RVEEH and Monash University.Participants ≥18 years old scheduled to undergo routine anterior eye segment surgery were recruited. Exclusion criteria were liver or kidney failure, breast feeding, pregnant, trying to conceive, allergy to voriconazole or benzalkonium chloride, use of latanoprost or medications contraindicated for voriconazole (Vfend package insert; Pfizer Inc., New York, NY), and active ocular inflammatory conditions. Written informed consent was obtained from each participant before enrolment.In accordance with previous studies (12, 16), for an expected standard deviation of ±1.0 mg/liter and a ±0.5-mg/liter margin of error of the mean associated with a 95% confidence level, the required sample size was 13 participants.The 2% solution was manufactured aseptically by the RVEEH Pharmacy Department. It was prepared by reconstituting Vfend IV (200-mg vial; Pfizer Australia Pty. Ltd., Sydney, NSW, Australia) with 9 ml of water for injection (Pfizer Pty. Ltd., Perth, WA, Australia) containing 0.01% benzalkonium chloride solution (benzalkonium chloride 50%; Professional Compounding Centers of America, TX; Australian Compounding Pharmacy, Australia).A single drop (0.05 ml) of the 2% voriconazole solution was administered by a nurse to the eye to be operated on every hour for 4 hours prior to operation. The last dose was administered approximately 1 hour before surgery. A drug administration diary was used to document the date and time of administration and any side effects experienced.At the start of surgery and before infusion of any intraocular irrigation solution, 0.1 to 0.2 ml of aqueous humor was aspirated through a paracentesis site with a 30-gauge needle attached to a syringe. Samples were immediately refrigerated at 4°C and analyzed within 7 days of collection.Voriconazole levels in the aqueous humor were quantified by a validated high-performance liquid chromatography assay with photodiode array detection. The analytical methods have been published previously (12).The Mann-Whitney test was used to investigate any significant difference in measured voriconazole concentrations between participants with and without diabetes mellitus. Student''s t test was used to compare differences between means.The 13 participants had a mean age (±standard deviation) of 70 ± 7.7 years (Table (Table2).2). All participants had phakic lens status and required unilateral cataract surgery in a noninflamed eye. No side effects or toxicities were reported.

TABLE 2.

Patient characteristics and voriconazole concentrations in aqueous humor
Age (yr)SexDiabetes mellitusSampling time after last voriconazole dose (h)Voriconazole concn in aqueous humor (mg/liter)
65MaleNo1.20.9
79FemaleNo1.31.0
69MaleYes1.11.0
61FemaleYes1.30.9
82MaleYes1.22.6
73MaleNo1.31.3
73MaleNo1.93.6
56MaleNo1.60.8
67MaleNo1.51.4
61MaleNo1.61.0
77MaleNo1.01.5
69MaleNo0.93.4
75MaleNo1.32.4
Open in a separate windowThe mean voriconazole concentration in the aqueous humor was 1.67 ± 0.97 mg/liter, while the mean sampling time after the last eye drop administration was at 1.3 ± 0.3 h (Table (Table2).2). There was no statistical difference (P = 0.67) in the mean aqueous humor voriconazole concentrations between the 10 nondiabetic participants (1.72 ± 1.03 mg/liter) and the 3 diabetic participants (1.50 ± 0.91 mg/liter).The mean aqueous humor voriconazole concentration in this study (using 2% voriconazole solution) was not significantly different (P = 0.68) from that reported by Lau et al. (12) using 1% voriconazole solution.In the Lau et al. study (12), the mean aqueous humor voriconazole concentration was 1.90 ± 1.12 mg/liter and the mean sampling time after the last dose was 1.1 ± 0.5 h. Results from the current study are comparable with those of Lau et al. (12). Both studies had identical numbers and frequencies of doses administered. In both studies, the trough voriconazole levels were measured (samples were collected approximately 1 h after the last one-hour drop). Furthermore, both studies administered the same volumes (∼0.05 ml) of eye drops at each dose (11) and used benzalkonium chloride (a transcorneal penetration enhancer [9]) as the preservative. The eye drops prepared in this study and those used by Lau et al. contained the same concentration of benzalkonium chloride (0.01%).The results from this study are consistent with previous studies (12, 16), demonstrating that topically administered voriconazole solutions achieve therapeutic concentration in the aqueous humor for treatment of fungi in Table Table1,1, including those encountered at our institution (RVEEH). Importantly, the concentration resulting from the 2% voriconazole eye drops is not significantly different from that reported for the 1% solution (12). It appears that the penetration of voriconazole through an intact infection-free cornea is not concentration dependent, at least for the concentration range studied (1% to 2%). This appears counterintuitive and challenges the hypothesis of the study but is consistent with observations in a recent animal study, where the voriconazole level in the corneas of horses with fungal keratitis did not change when the administered voriconazole eye drop concentration was changed from 1% to 3% (4).While the precise mechanism of fluid/drug transport through the cornea remains obscure (5), it is important to recognize that, in this and previous studies (12, 16), the eye drops were applied to noninfected eyes. To attempt this study in patients with active fungal keratitis would be difficult, given the very low incidence of fungal keratitis and the need for long-term treatment. Nonetheless, the voriconazole concentration in the infected eye depends neither on the size of the epithelial defect nor on epithelial removal, and thus it has been suggested that epithelial damage is not necessary for voriconazole penetration (15).We have previously demonstrated that the 2% eye drops are stable for up to 16 weeks when stored between 2 to 8°C and at 25°C (1). Furthermore, the 2% solutions have a pH range of 6.02 to 6.16 (1), which is usually well tolerated by the eye (7). It is unlikely, therefore, that any eye irritation resulting from the use of these eye drops (none reported in our study) would be a consequence of low pH. Systemic side effects resulting from the topical administration of the 2% voriconazole solution will be negligible as each administered dose (0.05 ml) contains approximately 1 mg of voriconazole, which, compared to the standard systemic daily dose of 400 mg, is unlikely to result in a systemic concentration that will cause side effects.In conclusion, the 2% voriconazole eye drops appear to be well tolerated. The concentration of voriconazole achieved in the aqueous humor was adequate (at least theoretically) to treat typical keratitis-causative fungi. Our data suggest that the penetration of voriconazole through an intact noninflamed cornea is unlikely to be concentration dependent for the concentration range 1 to 2%.  相似文献   

5.
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   

6.
7.
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   

8.
9.
Amphotericin B (AMB) concentrations were determined in pulmonary epithelial lining fluid (ELF) of 44 critically ill patients, who were receiving treatment with liposomal AMB (LAMB) (n = 11), AMB colloidal dispersion (ABCD) (n = 28), or AMB lipid complex (ABLC) (n = 5). Mean AMB levels (± standard errors of the means) in ELF amounted to 1.60 ± 0.58, 0.38 ± 0.07, and 1.29 ± 0.71 μg/ml in LAMB-, ABCD-, and ABLC-treated patients, respectively (differences are not significant).Invasive pulmonary mycoses exhibit a high mortality, particularly in critically ill patients (19). Amphotericin B (AMB) lipid formulations—liposomal AMB (AmBisome; Gilead) (LAMB), AMB colloidal dispersion (Amphotec [Three Rivers] and Amphocil [Torrex-Chiesi]) (ABCD), and AMB lipid complex (Abelcet; Zeneus) (ABLC)—display differences in plasma pharmacokinetics and tissue distribution (15, 26, 28). During treatment with AMB lipid formulations, AMB concentrations were investigated in epithelial lining fluid (ELF), which is a well-established model for pulmonary drug penetration (1-4, 6-10, 12, 17).This study was approved by the local ethics committee. Patients on lipid-formulated AMB requiring bronchoalveolar lavage (BAL) were enrolled (Table (Table1).1). AMB concentrations were assessed in 8-ml aliquots of BAL samples obtained by a standard procedure (16). BAL fluid was concentrated by evaporation, and AMB was quantified by high-performance liquid chromatography as described previously, with modifications for BAL samples (13). The concentrations were assessed by using a linear standard curve (R between 0.995 and 0.999), obtained from standards comprising 0.9% saline solution spiked with AMB. The lower detection limit of AMB in BAL fluid was 0.005 μg/ml. The assay has been found to be linear over the concentration range of 0.005 to 2.5 μg/ml for AMB in BAL fluid. The intraday and interday precisions were 3.2% and 4.7%, respectively. AMB concentrations in ELF were calculated by the urea dilution method (23), AMBELF = AMBBAL × (ureaPLA/ureaBAL), where AMBELF is the AMB concentration in ELF, AMBBAL is the AMB concentration in BAL fluid, ureaPLA is the urea concentration in plasma, and ureaBAL is the urea concentration in BAL fluid (23). Two ml of the BAL fluid was separated for urea quantification, which was performed using an enzymatic assay (urea/blood urea nitrogen; Roche) as with plasma.

TABLE 1.

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

TABLE 2.

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

10.
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   

11.
We determined the prevalence of fluoroquinolone resistance among the isolates of Mycobacterium tuberculosis from 605 pulmonary tuberculosis patients in Shanghai, China. Mutations in gyrA were found in 81.5% of phenotypically fluoroquinolone-resistant isolates and were used as a molecular marker of fluoroquinolone resistance. gyrA mutations were detected in 1.9% of strains pan-susceptible to first-line drugs and 25.1% of multidrug-resistant strains. Fluoroquinolone resistance was independently associated with resistance to at least one first-line drug and prior tuberculosis treatment.Fluoroquinolones are among the most promising antibiotic drugs for tuberculosis (TB) treatment and have the potential to become part of a new first-line treatment regimen against TB (12, 17). Fluoroquinolones were introduced into clinical practice in China nearly 20 years ago and have been widely used to treat common bacterial infections, TB patients infected with Mycobacterium tuberculosis strains resistant to first-line drugs, and TB patients with severe adverse reaction to first-line agents (2, 13). Although high levels of fluoroquinolone resistance have been detected among many common bacterial pathogens (16, 19), little is known about the fluoroquinolone resistance of M. tuberculosis. A previous study reported that the risk that a TB patient would acquire fluoroquinolone resistance was correlated with the patient''s previous exposure to fluoroquinolones (14). If TB patients are infected with M. tuberculosis strains that are resistant to fluoroquinolones, it will not be possible to use fluoroquinolones in anti-TB treatment regimens.To estimate the prevalence and to identify the risk factors associated with fluoroquinolone resistance among pulmonary TB patients in Shanghai, we performed a retrospective case control study using specimens and data collected and stored in the Tuberculosis Reference Laboratory, Shanghai Municipal Center for Disease Control and Prevention. The incidence rate of pulmonary TB in Shanghai in 2005 was 39.4 per 100,000 persons. From March 2004 through November 2007, clinical isolates from 4,663 patients with pulmonary TB were collected. Drug susceptibility testing for the major first-line drugs, specifically, isoniazid (0.2 μg/ml), rifampin (rifampicin) (40 μg/ml), ethambutol (2 μg/ml), and streptomycin (4 μg/ml), were routinely performed on each isolate by using the proportion method (4). A total of 85.1% of the TB patients were infected with M. tuberculosis strains susceptible to isoniazid, rifampin, ethambutol, and streptomycin; these patients are hereafter referred to as pan-susceptible. A total of 14.9% of the strains were resistant to at least one first-line drug, and 5.6% were multidrug resistant (MDR) (18). We selected a random sample of TB patients infected with pan-susceptible strains (n = 257), a random sample of TB patients infected with a strain monoresistant to isoniazid (n = 60), and a random sample of TB patients infected with a strain that was polyresistant but not MDR (n = 77), all TB patients infected with a strain monoresistant to rifampin (n = 36), and all 175 (70.9%) strains available from 247 reported MDR-TB patients. In total, the initial clinical isolates of M. tuberculosis and sociodemographic data from 605 pulmonary TB patients were included in the study. The study was approved by the ethics committee of Fudan University.Mutations in the fluoroquinolone resistance-determining region of gyrA gene are the most important mechanism of fluoroquinolone resistance in M. tuberculosis (1, 7, 8, 11). To confirm that mutations in the fluoroquinolone resistance-determining region of gyrA can be used as a reliable molecular marker for detection of fluoroquinolone-resistant M. tuberculosis in Shanghai, we compared ofloxacin (2 μg/ml) susceptibility testing and gyrA sequencing results for 175 MDR strains. The sensitivity of gyrA mutations among 54 isolates with the ofloxacin-resistant phenotype was 81.5% (95% confidence interval [CI], 68.6% to 90.8%). The specificity of gyrA sequencing was 100% (97.5% CI, 97.0% to 100.0%). Next, we sequenced the gyrA gene of the initial isolate from 605 pulmonary TB patients (Table (Table1).1). The prevalence of gyrA mutations was lowest among pan-susceptible strains of M. tuberculosis (1.9%) and higher among strains that were resistant to one or more first-line drugs (17.0%), particularly MDR strains (25.1%). By univariate analysis, gyrA mutations were more likely to occur among strains resistant to first-line anti-TB drugs, especially MDR strains, than among pan-susceptible strains (Table (Table2).2). By multivariate logistic regression modeling, with adjustment for age, MDR was the strongest independent predictor of a gyrA mutation (Table (Table3).3). We tested the multivariate model for goodness of fit (P = 0.607), and interaction terms did not significantly improve the model.

TABLE 1.

Estimates of prevalence levels and 95% CIs for gyrA mutations by drug susceptibility test result
CharacteristicaNo. of isolates in study population
% Prevalence estimate (95% CI)
TotalWith gyrA
Pan-susceptibility25751.9 (0.6-4.5)
Resistance to one or more first-line drugs3485917.0 (13.2-21.3)
Any resistance to INH3025317.5 (13.4-22.3)
Any resistance to RIF2205022.7 (17.4-28.8)
Any resistance to EMB691927.5 (17.5-39.6)
Any resistance to SM1964020.4 (15.0-26.7)
Monoresistance to INH6046.7 (1.8-16.2)
Monoresistance to RIF36411.1 (3.1-26.1)
Polyresistance7779.1 (3.7-17.8)
Any drug resistance except MDR173158.7 (4.9-13.9)
MDR1754425.1 (18.9-32.2)
Open in a separate windowaINH, isoniazid; RIF, rifampin; EMB, ethambutol; SM, streptomycin; polyresistance, resistance to two or more first-line drugs but not MDR; MDR, resistance to at least INH and RIF.

TABLE 2.

Results of univariate analysis of characteristics of TB patients by presence or absence of gyrA mutation, a marker of fluoroquinolone resistance
CharacteristicaNo. (%) of isolates with gyrA mutation
ORc (95% CI)P
Present (n = 64)Absent (n = 541)
Drug susceptibility
    Pan-susceptibility5 (7.8)252 (46.6)1.0
    Resistance to one or more drugsb59 (92.2)289 (53.4)10.3 (4.1-33.3)<0.00005
    Any INH resistanceb5324910.7 (4.2-34.9)<0.00005
    Any RIF resistanceb5017014.8 (5.8-48.4)<0.00005
    Any EMB resistanceb195019.2 (6.4-67.8)<0.00005
    Any SM resistanceb4015612.9 (4.9-42.6)<0.00005
    Monoresistance to INHb4563.6 (0.7-17.2)0.069
    Monoresistance to RIF4326.3 (1.2-30.6)0.016
    Polyresistanceb7705.0 (1.3-20.7)0.0018
    MDRb4413116.9 (6.5-55.7)<0.00005
    Any category of drug resistance except MDRb151584.8 (1.6-17.1)0.0002
    MDR vs any other category of drug resistance441313.5 (1.8-7.1)0.0000
Patient status
    Retreatment case32 (50.0)113 (20.9)3.8 (2.1-6.7)<0.00005
    New case32 (50.0)428 (79.1)1.0
Patient origin
    Resident of Shanghai46 (71.9)333 (61.6)1.6 (0.9-3.0)0.107
    Migrant18 (28.1)208 (38.4)1.0
Age
    ≥46 yr43 (67.2)264 (48.0)2.2 (1.2-3.9)0.005
    <46 yr21 (32.8)277 (51.2)1.0
Sex
    Male47 (73.4)409 (75.6)1.1 (0.6-2.1)0.704
    Female17 (26.6)132 (24.4)1.0
Open in a separate windowaINH, isoniazid; RIF, rifampin; EMB, ethambutol; SM, streptomycin; polyresistance, resistance to two or more first-line drugs but not MDR; MDR, resistance to at least INH and RIF.bPan-susceptible isolates were used as the reference group.cOR, odds ratio.

TABLE 3.

Results of multivariate logistic regression analysis for identification of covariates that were independently associated with gyrA mutations, a marker of fluoroquinolone resistancea
CharacteristicAdjusted OR (95% CI)P
MDR13.8 (5.2-36.5)<0.0005
Monoresistance to rifampin6.3 (1.6-25.1)0.010
Polyresistance4.5 (1.3-14.8)0.015
Age ≥46 yr2.4 (1.3-4.4)0.005
TB retreatment2.1 (1.2-3.8)0.014
Open in a separate windowaOR, odds ratio; MDR, resistance to at least isoniazid and rifampin; polyresistance, resistance to two or more first-line drugs but not MDR.A history of prior TB treatment was also an independent predictor of gyrA mutations (Table (Table3).3). Medical records were available for 15 of 25 Shanghai residents but none of the 7 migrants with retreatment TB and a gyrA mutation. For 53% (8/15) of the TB patients whose medical records were reviewed, there was documentation indicating that the patient had received at least 2 weeks of fluoroquinolone treatment in a previous TB treatment regimen. Fluoroquinolone use during anti-TB treatment likely contributed to acquired fluoroquinolone resistance in retreatment patients.The classical fluoroquinolone drug susceptibility test recommended by the World Health Organization is the proportion method (5, 15), but this method is relatively time-consuming and labor-intensive. The positive predictive value of screening for gyrA mutations to detect fluoroquinolone resistance varies, depending on the study site and sampling, the prevalence of fluoroquinolone resistance, and the study design. Eighty-six percent (23/30) of fluoroquinolone-resistant isolates in a study performed in North America had a gyrA mutation (3), but only 35.7% (5/14) fluoroquinolone-resistant isolates had gyrA mutation in a study with a small sample of fluoroquinolone-resistant patients in Taiwan (14). A study in Beijing, China, that used denaturing high-pressure liquid chromatography and DNA sequencing reported that 56% of 87 ofloxacin-resistant M. tuberculosis clinical strains had a mutation in gyrA (10). In our study, the sensitivity of the gyrA mutation for detection of the ofloxacin-resistant phenotype was 81.5%, and this molecular marker was reasonable for prediction of phenotypic fluoroquinolone resistance. However, we did not screen for all possible mutations that have been reported to confer fluoroquinolone resistance, such as mutations in the gyrB gene and the efflux pump (6, 9, 14), and the mechanisms of fluoroquinolone resistance in M. tuberculosis are still not fully understood. gyrA mutations do not perfectly predict fluoroquinolone resistance phenotypes, and we may have underestimated the true prevalence of fluoroquinolone resistance in our study population.Fluoroquinolones have the potential to become part of a new first-line treatment regimen against TB but will not be effective if the prevalence of fluoroquinolone resistance among new TB cases is high. In our retrospective study, 1.9% of the pan-susceptible strains of M. tuberculosis from pulmonary TB patients in Shanghai had a gyrA mutation. Fluoroquinolone resistance was independently associated with resistance to first-line drugs and prior TB treatment. Although our data did not permit a statistical comparison, fluoroquinolone resistance is likely to be associated with prior fluoroquinolone use during prior TB treatment. Inappropriate regimens, such as monotherapy or a regimen delivered without directly observed therapy, likely contribute to acquired fluoroquinolone resistance in M. tuberculosis. Thus, more attention should be paid to fluoroquinolone use and potential acquired fluoroquinolone resistance during anti-TB therapy, especially in populations where TB treatment guidelines are not well established and MDR TB occurs.  相似文献   

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

TABLE 1.

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

14.
This study compared the efficacies of two N-methylglucomine antimoniate (MA) dose regimens for treating macaques with Leishmania braziliensis-induced chronic skin disease. Whereas all animals treated with the full dose (20 mg MA/kg/day) were cured, 50% of the monkeys receiving a low-dose regimen (5 mg MA/kg/day) relapsed. The antimony concentrations in macaque plasma and tissue samples were greater in the full-dose group than in that receiving a subtherapeutic MA regimen. Our data also suggest the presence of drug-induced hepatic pathology.Leishmaniasis is a cause of significant morbidity and mortality throughout the world, with two million new cases of human infection worldwide each year. However, an approved vaccine to prevent leishmaniasis does not exist (7). Treatment for leishmaniasis relies largely on pentavalent antimony (Sbv) compounds (meglumine antimoniate and sodium stibogluconate). A number of other therapeutic agents may be employed, but high costs have limited the large-scale use of the most potent drugs (4). Factors limiting the usefulness of SbV therapy include their adverse toxic effects (arthralgias, myalgias, cardiac arrhythmia, pancreatitis, and hepatic or renal function impairment) and the increasing occurrence of parasite resistance (2, 4). Nevertheless, responses to SbV vary considerably, depending on both the parasite''s intrinsic drug sensitivity and the host''s immune status (1, 4, 15). Poor clinical responses can also be attributed to inadequately dosed antimony regimens (1, 2) and problems concerning drug pharmacokinetics (PK) or biodisposal (5, 13).The Leishmania-macaque model has proven to be a valuable in vivo system for anti-infectious disease drug and vaccine development studies (7, 14). The aim of the present study was to compare the pharmacological parameters (PK, toxicity, and efficacy) of a low dose of MA with those of a standard dose of MA in macaques with L. braziliensis infection. All animal studies were performed under the guidance and with the approval of the Institutional Animal Care and Use Committee. Groups of six outbred adult rhesus macaques (Macaca mulatta) were used in this study. A high dose (107 promastigotes) of virulent L. braziliensis (MHOM/BR/2000/CP13396 strain) was injected intradermally above the left upper eyelid of each monkey. The infection was allowed to proceed until the macaques reached skin disease progression. At 9 weeks postinfection, macaques received a 21-day course of a low dose (5 mg/kg/day) or a full dose (20 mg/kg/day) of MA administered through an intramuscular route. A vehicle-only-treated control group was included to assess the development of a local skin lesion caused by the infection. Animals were euthanized on days 55 (140, 142, O6, L30, M2, and O34) and 95 (S62, T32, U48, U12, U46, and X53) after the completion of treatment, and selected necropsy tissue specimens (liver, spleen, and kidneys) were removed from drug-cured and control macaques to determine residual tissue antimony concentrations and for histological examination to assess antimony-induced histopathological changes. Antimony concentrations in macaque plasma and tissues were determined by inductively coupled plasma mass spectrometry (ICP-MS) under the optimized conditions previously described (11).Although L. braziliensis-infected macaques showed a high degree of variability in lesion size (area range, 20 to 540 mm2) before antimonial therapy, the ulcerative cutaneous lesion persisted in untreated animals until the end of the observation period (Fig. (Fig.1).1). Treatment with both therapeutic schemes rapidly reduced the lesion size (in comparison with that of untreated lesions) after treatment. However, while complete healing was achieved in all animals receiving a regular MA schedule, three out of six monkeys treated with a low-dose regimen relapsed with the presentation of macroscopic wound inflammation after a clinical cure and wound reopening 3 to 4 months after the cessation of therapy. The concentration-time profiles of Sb in macaque plasma (Fig. (Fig.2)2) confirmed that drug exposure was much lower in that group (range, 11.3 to 149.3 ng Sb/g) than in macaques treated with a full-dose regimen (range, 36.4 to 150.5 ng Sb/g). Except for days 16, 23, 61, and 68, the differences in plasma Sb concentrations between the two groups were statistically significant (P < 0.001 to 0.05).Open in a separate windowFIG. 1.Response of L. braziliensis cutaneous infections in rhesus macaques to either low-dose (5 mg/kg/day; group A) or standard-dose (20 mg/kg/day; group B) MA treatment. To assess therapy, lesion size development was scored weekly following infection and treatment. All values represent the mean ± standard deviation. Before treatment, there was no statistically significant difference (P > 0.05) in mean lesion size over time between the macaque groups.Open in a separate windowFIG. 2.Time course of plasma Sb concentrations after intramuscular injection of either a low dose (group A) or a full dose (group B) of MA into L. braziliensis-infected macaques. For detection of total Sb, heparin-anticoagulated blood samples were analyzed by ICP-MS as described previously (11). *, Significant differences (P < 0.001 to 0.05) in plasma Sb concentrations between the two groups.Sustained nadir blood levels of Sb gave rise to residual drug accumulation in different soft tissues. Approximately 2 and 3 months after the last dose of MA, Sb was detected in the livers, spleens, and kidneys of treated macaques, with concentrations in renal tissue significantly lower than those found in hepatic tissue (Table (Table1).1). No overt clinical signs of toxicity were observed in any macaque, regardless of the administration scheme. Nevertheless, histological evidence of MA-induced hepatic injury (Fig. (Fig.3)3) was observed in all treated macaques. There was a clear correlation (r = 0.94; P = 0.001) between tissue Sb levels and the extent to which hepatocytes were affected. In contrast, no histopathological alterations were found in any other tissue examined.Open in a separate windowFIG. 3.Histopathology of liver samples from representative drug-cured L. braziliensis-infected macaques (A, L30; B, 140; C, U46; D, O34). Micrographs show focal hepatocellular acidophilic necrosis (arrows) with surrounding inflammatory infiltrates (circled). These obliterate the sinusoids and protrude into the parenchyma and are associated with fatty changes in stellate cells (arrowheads). Hypotrophy of the hepatic parenchyma at the center of the lobules (boxed) and numerous hemosiderin deposits within liver cells and Kupffer cells (inset in panel D) are also illustrated. Hematoxylin-and-eosin staining was used.

TABLE 1.

Residual concentrations of antimony found in selected necropsy tissue specimens from drug-cured L. braziliensis-infected macaquesa
Posttreatment time,b group, and monkey or parameterResidual Sb concn (ng/g) in:
LiverSpleenKidneys
Group A
    55 days
        1402,4901,800340
        1422,8901,050330
        063,3601,450830
        Mean ± SD2,913 ± 1,435.51,433 ± 375.3500 ± 285.8
    95 days
        S62831370169
        T32430360125
        U48747700162
        Mean ± SD669.3 ± 211.5476.7 ± 193.5152 ± 23.6
Group B
    55 days
        L304,1802,630580
        M2NDc1,1001,000
        0345,4201,050515
        Mean ± SD4,8001,593 ± 898.1698.3 ± 263.3
    95 days
        U121,8002,150320
        U462,3701,190374
        X531,2303,320245
        Mean ± SD1,800 ± 5702,220 ± 1,067313 ± 64.8
Open in a separate windowaPrimates were treated with either a low dose (5 mg/kg/day; group A) or a full dose (20 mg/kg/day; group B) of MA administered intramuscularly.bNecropsies were performed at different time points (as described in Materials and Methods) after the completion of treatment. For this analysis, specimens from infected but untreated (n = 2) monkeys were also included as negative controls. Residual levels of Sb were compared among organs (matched per monkey) for each dose group by the Friedman test, followed by Dunn''s multiple-comparison test: kidneys versus livers in groups A and B, P < 0.05.cND, not done.Our main findings in macaques receiving a low SbV dosage regimen are consistent with those obtained in our previous study (14) but differ from the observations of Oliveira-Neto et al. (12), who recommended subtherapeutic MA concentrations for treating L. braziliensis-infected patients. This is likely linked to variabilities in parasite drug susceptibility (1, 4, 15). In fact, taxonomic studies have shown a high degree of genetic variation in natural populations of L. braziliensis from different geographic areas in Brazil (6). It should be noted that sublethal doses contribute to the selection of drug-resistant parasites (9), with the parasites that are inherently most drug resistant being favored (10). Moreover, drug-resistant clones of Leishmania spp. may exhibit cross-resistance (8).The plasma PK profile of Sb in L. braziliensis-infected macaques treated with MA was similar to that reported for human cases receiving SbV drugs (3, 11, 13). Accordingly, most of the Sb absorbed from the injection site was eliminated rapidly, but long-lived nadir plasma Sb concentrations caused a gradual accumulation of the drug in tissues after repeated daily dosing. With the exception of the spleen in groups treated with the standard dose, tissue Sb concentrations in monkeys necropsied 95 days after the end of treatment were lower than the levels found in monkeys examined 40 days earlier (Table (Table1).1). It remains to be determined if the observed variability in tissue Sb levels in macaques in the same dose group results from the outbred genetics of these primates. The species of Sb in the organism and the mechanisms by which it is transported, distributed to tissues, and eliminated from the body remain unclear (4). We also provided evidence for an association between tissue Sb levels and the extent of hepatocyte damage. The drug-induced hepatic injury could be related to the conversion of SbV to SbIII, which has been demonstrated to be considerably more toxic than SbV in different test systems (3).  相似文献   

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

TABLE 1.

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

TABLE 2.

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

TABLE 3.

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

TABLE 4.

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

17.
In vitro activity of tebipenem, a new oral carbapenem antibiotic, against clinical Haemophilus influenzae isolates was compared with those of 8 reference agents. Isolates were classified into 6 resistance classes after PCR identification of β-lactamase genes and ftsI gene mutations. For all isolates, the minimal concentration at which 90% of isolates were inhibited was lower for tebipenem than for the reference oral antibiotics, except for cefditoren. Tebipenem also showed excellent bactericidal activity against β-lactamase-nonproducing, ampicillin-resistant isolates.Tebipenem-pivoxil (TBM-PI) is a new oral carbapenem agent whose active metabolite, tebipenem (TBM), shows broad-spectrum activity against Gram-positive and -negative bacteria, except for Pseudomonas aeruginosa (4, 7). Unlike other clinically available oral β-lactam antibiotics, TBM displays excellent activity against Streptococcus pneumoniae strains, including penicillin-resistant strains (5). In consideration of these characteristics, the clinical efficacy of TBM-PI for treatment of community-acquired pediatric infections such as pneumonia, acute otitis media (AOM), and sinusitis was investigated. In a phase III clinical study of AOM, the efficacy of TBM-PI (3.5 to 5 mg/kg twice a day [b.i.d.]) was equal to that of high-dose cefditoren-pivoxil (CDN-PI; 4.2 to 6 mg/kg three times a day [t.i.d.]) (11). Among the important pathogens causing pneumonia and AOM, the antimicrobial characteristics of TBM against Haemophilus influenzae are not yet verified. Furthermore, β-lactamase-nonproducing, ampicillin-resistant (BLNAR) strains are increasing in Japan and various other countries (9). In the present study, we evaluated in vitro antibacterial and bactericidal activities of TBM and reference agents against H. influenzae strains, including BLNAR strains.A total of 232 H. influenzae strains were collected between October 2005 and December 2008 from pediatric patients at random. The details of the strains were as follows: 112 were obtained from tympanic effusions in patients with AOM, 30 were obtained from patients with pneumonia, and 90 were obtained from cerebrospinal fluid samples collected from patients with meningitis. To clarify the genetic background as related to β-lactam resistance, we used PCR identification to place H. influenzae strains in 1 of 6 genetic resistance classes (classes indicated with a “g” prefix) (Table (Table1)1) (2). Susceptibility testing was performed using an agar dilution method (3). Muller-Hinton (MH) agar (Becton Dickinson, Sparks, MD) with 0.5% yeast extract, 2% defibrinated, heat-treated horse blood, and 15 μg of β-NAD+ per ml (subsequently referred to as “the 3 supplements”) was used. Plates were examined after incubation at 37°C in a 5% CO2 atmosphere for 20 h. Table Table22 shows the MIC ranges, MIC50s, and MIC90s of TBM and 8 reference antibiotics against H. influenzae strains in each of the 6 genetic resistance classes (total n = 232). Antimicrobial activities of TBM for genetically β-lactamase-nonproducing, ampicillin-susceptible (gBLNAS), genetically low β-lactamase-nonproducing, ampicillin-resistant (gLow-BLNAR), and genetically β-lactamase-nonproducing, ampicillin-resistant (gBLNAR) strains were essentially equal to those of meropenem. The increased ratios of MIC90s between gBLNAS and gBLNAR strains for TBM, cefditoren, and meropenem were lower (4 to 8 times) than those for cefotaxime, amoxicillin, and cefdinir (32 to 64 times). The susceptibility distribution of the 232 H. influenzae strains to TBM according to resistance class is shown in Fig. Fig.1.1. The MIC peaks of TBM against gBLNAS and gBLNAR strains were 0.063 and 0.5 μg/ml, respectively. Antimicrobial activities of cephalosporins, except for cefditoren, were greatly decreased by amino acid substitutions in penicillin-binding protein 3 (PBP3) of BLNAR strains (9). The reason why TBM was affected relatively little by these amino acid substitutions appears to involve the binding affinities of TBM for PBPs; TBM was found to bind not only to PBP3 but also to PBP1B, PBP2, and PBP4 (10).Open in a separate windowFIG. 1.Distribution of MICs of TBM for clinical isolates of H. influenzae (n = 232) classified into groups consisting of gBLNAS, genetically β-lactamase-producing ampicillin-resistant (gBLPAR), gLow-BLNAR, gBLNAR, genetically β-lactamase-producing, amoxicillin-clavulanic acid-resistant I (gBLPACR-I), and gBLPACR-II strains (“I” and “II” indicate different substitutions in PBP3) (Table (Table11).

TABLE 1.

Identification of 6 resistance classes based on PCR
Resistance classGenetic background results
Amino acid substitutions in PBP3a
β-lactamase gene
Asn526Lys or Arg517HisSer385Thr and Asn526Lys or Ser385Thr and Arg517His
gBLNAS
gLow-BLNAR+
gBLNAR+
gBLPAR+
gBLPACR-I++
gBLPACR-II++
Open in a separate windowaAmino acid substitutions in penicillin-binding protein 3 (PBP3) that affect decreases of susceptibilities of H. influenzae strains for β-lactam antibiotics were selected (12).

TABLE 2.

Comparison of in vitro activities of tebipenem and those of reference antibiotics against H. influenzae strains classified genotypically by PCR
Resistance class (no. of H. influenzae strains) and antibioticaMIC (μg/ml)b
Range50%90%
gBLNAS (65)
    Tebipenem0.008-0.250.0630.12
    Ampicillin0.12-10.250.5
    Amoxicillin0.12-10.50.5
    Cefdinir0.12-10.250.5
    Cefditoren0.002-0.0630.0160.031
    Cefotaxime0.004-0.0630.0160.031
    Meropenem0.016-0.120.0630.12
    Clarithromycin4-16816
    Azithromycin0.25-824
gLow-BLNAR (32)
    Tebipenem0.031-0.50.250.5
    Ampicillin0.5-211
    Amoxicillin0.5-424
    Cefdinir0.5-40.52
    Cefditoren0.016-0.0630.0310.063
    Cefotaxime0.016-0.120.0310.12
    Meropenem0.063-0.50.120.25
    Clarithromycin4-16816
    Azithromycin0.5-422
gBLNAR (119)
    Tebipenem0.031-10.251
    Ampicillin0.5-3228
    Amoxicillin0.25-64832
    Cefdinir2-32832
    Cefditoren0.031-10.250.25
    Cefotaxime0.063-40.51
    Meropenem0.031-0.50.250.5
    Clarithromycin4-32816
    Azithromycin0.5-824
gBLPAR (TEM-1 [6])
    Tebipenem0.063-0.12
    Ampicillin8-32
    Amoxicillin8-16
    Cefdinir0.25-0.5
    Cefditoren0.008-0.031
    Cefotaxime0.008-0.031
    Meropenem0.031-0.063
    Clarithromycin8-16
    Azithromycin1-2
gBLPACR-I (2)
    Tebipenem0.12-0.25
    Ampicillin8-32
    Amoxicillin8
    Cefdinir1
    Cefditoren0.016
    Cefotaxime0.031-0.063
    Meropenem0.063-0.12
    Clarithromycin8
    Azithromycin2
gBLPACR-II (8)
    Tebipenem0.25-0.5
    Ampicillin2->64
    Amoxicillin2->64
    Cefdinir8-16
    Cefditoren0.12
    Cefotaxime0.25-0.5
    Meropenem0.063-0.25
    Clarithromycin8->64
    Azithromycin2-64
Open in a separate windowaThe oral antibiotics used in this study were tebipenem, ampicillin, amoxicillin, cefditoren, cefdinir, clarithromycin, and azithromycin. The parenteral antibiotics used in this study were cefotaxime and meropenem. A total of 232 H. influenzae strains were used.bThe breakpoint MICs (in μg/ml) of each antibiotic against susceptible strains are as follows: ampicillin, ≤1; amoxicillin-clavulanic acid, ≤4/2; cefdinir, ≤1; clarithromycin, ≤8; azithromycin, ≤4; cefotaxime, ≤2; and meropenem, ≤0.5 (1).Time-kill curves for TBM and reference antibiotics in BLNAR strains JPH002 and JPH1306 were determined at concentrations corresponding to the MIC and double the MIC. Colonies precultured on chocolate II agar plates (Nippon Becton Dickinson, Tokyo, Japan) were suspended in tubes of MH broth, with the turbidity adjusted to a 0.5 McFarland standard. This bacterial suspension, diluted 10-fold using MH broth with the 3 supplements, was then grown at 37°C for 120 min. The culture (500 μl) then was inoculated into 9.5 ml of fresh MH broth containing each of the antibiotics with the 3 supplements and 9.5 ml of broth additionally supplemented with 10% fresh human serum to approximate conditions in vivo. Tubes then were incubated without shaking, and cultures were sampled at predetermined intervals.Figure Figure22 shows time-kill curves for TBM and β-lactam reference antibiotics at concentrations corresponding to the MIC and double the MIC for gBLNAR strain JPH002. A 3-log10 reduction of bacterial cells by TBM in the supplemented MH broth was achieved at 4 h of exposure at the MIC. The same reduction by cefditoren, ampicillin, or amoxicillin required 6 h or more of exposure at the MIC. Using supplemented MH broth containing 10% human serum, a 3-log10 reduction of bacterial cells by TBM was achieved with 2 h of exposure at the MIC. In contrast, similar reductions by cefditoren, ampicillin, and amoxicillin required 6, 4, and 6 h of exposure at the MIC, respectively. Using this approximation of conditions in vivo, the bactericidal activity of TBM against gBLNAR strains was observed within a short time after initiating exposure to TBM at equal to or greater than the MIC and was higher than those of cefditoren, ampicillin, and amoxicillin. The activity of TBM in the supplemented MH broth including serum was even better than that in supplemented broth without serum. The bactericidal activity of TBM against gBLNAR strain JPH1306 was similar to that against gBLNAR strain JPH002 (data not shown).Open in a separate windowFIG. 2.Time-kill curves for TBM and 3 other β-lactam antibiotics at the MICs (•) and double the MICs (▴) for gBLNAR strain JPH002. ○, control (no drug). HTM, MH broth supplemented with 0.5% yeast extract, 2% defibrinated and heat-treated horse blood, and 15 μg of β-NAD+ per ml.The functions of PBPs in H. influenzae can be deduced from those in Escherichia coli, considering that PBPs in H. influenzae and E. coli show high homology (6). Inhibition of PBP1A and -B in E. coli causes a rapid lysis reaction (8). In the present study, the bactericidal activity of TBM against gBLNAR strains was superior to those of cefditoren, ampicillin, and amoxicillin. This excellent bactericidal activity of TBM against H. influenzae in vitro also may relate to the lysis reaction caused by inhibition of PBP1A and -B.In summary, the antibacterial and bactericidal activities of TBM against BLNAR H. influenzae were different from those of penicillins and cephalosporins. These phenomena might be related to the characteristic binding affinity of TBM for PBPs.  相似文献   

18.
The in vitro activity of iclaprim, a novel diaminopyrimidine derivative, was evaluated against 5,937 recent gram-positive clinical isolates collected in the United States and Europe. Iclaprim demonstrated potent activity against Staphylococcus aureus (including methicillin-resistant S. aureus [MRSA]), beta-hemolytic Streptococcus spp., and Enterococcus faecalis strains tested. In addition, iclaprim exhibited bactericidal activity against all S. aureus strains tested, including MRSA.Staphylococcus aureus strains, including methicillin-resistant S. aureus (MRSA) strains, beta-hemolytic streptococci (most commonly Streptococcus pyogenes and Streptococcus agalactiae), and Enterococcus spp. are the principal gram-positive pathogens responsible for complicated skin and skin-structure infections (cSSSI). The increasing prevalence of MRSA in hospital and community settings (5, 20), as well as the potential for resistance or emergence of resistance during therapy to drugs such as vancomycin (1, 19), linezolid (9), and daptomycin (12), underscores the urgent need for additional well-differentiated therapeutic agents (17). Iclaprim is a new-generation diaminopyrimidine that potently and selectively inhibits bacterial dihydrofolate reductase (DHFR) (11, 18, 21). It was designed by a rational drug design approach based on structural information available on trimethoprim (TMP), the best known compound of the diaminopyrimidine class which, either alone or in its synergistic 1:19 combination with sulfamethoxazole, has been widely used in medical practice for over four decades. Iclaprim has been shown to have very potent activity against gram-positive bacteria that are susceptible to TMP. Iclaprim inhibits bacterial DHFR in a similar manner to TMP but possesses higher affinity due to increased hydrophilic interactions between iclaprim and DHFR (15). Thus, iclaprim may retain activity against some TMP-resistant isolates, but it would be lower than that against TMP-susceptible strains.Unlike TMP, the spectrum of activity of iclaprim is more focused against gram-positive pathogens, including MRSA strains. Iclaprim has been shown to be rapidly bactericidal against these strains and to possess a low potential for resistance development when used on its own, without the synergistic combination of a sulfonamide agent (8). For these reasons, iclaprim is under development as a monotherapy, and an intravenous formulation of iclaprim has completed two phase 3 trials for the treatment of cSSSI caused by gram-positive pathogens (11, 16). In addition, an oral formulation of iclaprim as a step-down therapy for patients with cSSSI is ongoing (16).We evaluated the potency and bactericidal activity of iclaprim against a large collection of contemporary gram-positive isolates from hospitalized patients in the United States (26 centers), the European Union (22 centers), Israel (1 center), and Turkey (1 center) from 2004 to 2006. In total, 5,937 clinical isolates representative of predominant gram-positive pathogens in cSSSI were tested. All organisms were collected from skin and soft tissue, bloodstream, and respiratory clinical specimens. The numbers of individual strains for each species tested are shown in Tables Tables11 and and2.2. Susceptibility testing by broth microdilution, including the appropriate quality controls, was performed according to the documents CLSI M7-A7 (2) and CLSI M100-S18 (3). Antimicrobial agents tested included iclaprim (Arpida Ltd., Reinach, Switzerland) and comparators. Minimum bactericidal concentration (MBC) values were determined for 101 randomly selected strains. MBC experiments were performed by plating the broth from wells from at least five log2 dilutions from the MIC (13, 14) onto growth medium. The lowest concentration of the antimicrobial that killed ≥99.9% of the starting test inoculum was defined as the MBC endpoint. Cidality was defined as an MBC/MIC ratio of ≤4. In all cases, the thymidine content of the test medium was assessed to ensure that no artifactual inhibition of iclaprim activity occurred (2, 6).

TABLE 1.

Activity of iclaprim and comparator agents tested against Staphylococcus aureus isolates from the United States and the European Union
Organism and antimicrobial agent (no. of isolates tested)MIC (μg/ml)
% Susceptible/resistant isolatesa
50%90%Range
S. aureus (4,516)
    Methicillin-sensitive (1,513)
        Iclaprim0.060.120.008-8−/−
        Trimethoprim110.06->6498.7/1.3
        Trimethoprim-sulfamethoxazole0.060.060.015->899.3/0.7
        Erythromycin0.25>4≤0.12->477.9/21.1
        Clindamycin≤0.12≤0.12≤0.12->496.4/3.4
        Tetracycline≤0.5≤0.5≤0.5->1695.8/4.0
        Ciprofloxacin0.51≤0.12->492.0/5.9
        Levofloxacin0.250.5≤0.12->494.2/5.6
        Linezolid220.5-2100.0/−
        Vancomycin11≤0.5-2100.0/0.0
    Methicillin-resistant (3,003)
        Iclaprim0.060.12≤0.004-8−/−
        Trimethoprim120.06->6492.9/7.1
        Trimethoprim-sulfamethoxazole0.060.250.015->896.1/3.9
        Erythromycin>4>4≤0.12->414.6/84.3
        Clindamycin≤0.12>4≤0.12->452.7/47.0
        Tetracycline≤0.5>16≤0.5->1686.5/12.8
        Ciprofloxacin>4>4≤0.12->416.3/83.1
        Levofloxacin>4>4≤0.12->416.8/82.4
        Linezolid22≤0.25-2100.0/−
        Vancomycin11≤0.5-2100.0/0.0
Beta-hemolytic streptococci (808)
    Group A (604)
        Iclaprim0.0150.03≤0.004-0.12−/−
        Trimethoprim0.250.5≤0.03-2−/−
        Trimethoprim-sulfamethoxazole0.060.120.015-0.25−/−
        Erythromycin≤0.12>4≤0.12->483.3/16.1
        Clindamycin≤0.12≤0.12≤0.12->496.4/3.3
        Tetracycline≤0.5≤0.5≤0.5->1690.2/9.6
        Levofloxacin0.510.25-2100.0/0.0
        Vancomycin≤0.5≤0.5≤0.5-1100.0/−
        Linezolid110.5-2100.0/−
        Penicillin≤0.06≤0.06≤0.06100.0/−
    Group B (204)
        Iclaprim0.120.250.015-0.5−/−
        Trimethoprim140.25-8−/−
        Trimethoprim-sulfamethoxazole0.060.120.03-0.25−/−
        Erythromycin≤0.12>4≤0.12->473.0/26.0
        Clindamycin≤0.12>4≤0.12->488.7/10.3
        Tetracycline>16>16≤0.5->1621.6/77.9
        Levofloxacin0.510.25->499.5/0.5
        Vancomycin≤0.5≤0.5≤0.5100.0/−
        Linezolid110.5-2100.0/−
        Penicillin≤0.06≤0.06≤0.06100.0/−
E. faecalis (310)
    Iclaprim0.0154≤0.004->8−/−
    Trimethoprim0.25>64≤0.03->64−/−
    Trimethoprim-sulfamethoxazole0.03>80.008->8−/−
    Erythromycin>4>4≤0.12->413.5/52.9
    Clindamycin>4>4≤0.12->4−/−
    Tetracycline>16>16≤0.5->1628.4/71.3
    Ciprofloxacin1>40.25->464.8/34.2
    Levofloxacin1>40.25->465.2/34.2
    Teicoplanin≤0.5≤0.5≤0.5->1698.4/1.0
    Vancomycin12≤0.5->1697.7/1.9
    Linezolid120.5-2100.0/0.0
    Ampicillin≤4≤4≤4100.0/0.0
E. faecium (303)
    Iclaprim2>8≤0.004->8−/−
    Trimethoprim32>64≤0.03->64−/−
    Trimethoprim-sulfamethoxazole>8>8≤0.004->8−/−
    Erythromycin>4>4≤0.12->43.0/84.5
    Clindamycin>4>4≤0.12->4−/−
    Tetracycline≤0.5>16≤0.5->1667.3/32.0
    Ciprofloxacin>4>40.25->46.9/86.5
    Levofloxacin>4>40.5->414.2/82.2
    Teicoplanin≤0.5>16≤0.5->1663.4/34.0
    Vancomycin1>16≤0.5->1655.8/43.2
    Linezolid120.5->899.3/0.7
    Ampicillin>16>16≤4->1611.9/88.1
Open in a separate windowaCriteria as published by the CLSI (1). β-Lactam susceptibility should be directed by the oxacillin test results. −, no criteria have been established by the CLSI (1).

TABLE 2.

Iclaprim, trimethoprim, and trimethoprim-sulfamethoxazole MIC distributions for MSSA, MRSA, beta-hemolytic streptococcal strains, and enterococci evaluated by reference methods
Organism and antimicrobial agent (no. of isolates tested)aNo. of isolates (cumulative %) inhibited at indicated MIC (μg/ml)b
≤0.0080.0150.030.060.120.250.51248163264>c
MSSA (1,513)
    Iclaprim1 (0.1)3 (0.3)88 (6.1)797 (58.8)557 (95.6)45 (98.5)1 (98.6)1 (98.7)1 (98.7)17 (99.9)2 (100.0)
    Trimethoprim0 (0.0)2 (0.1)2 (0.3)32 (2.4)483 (34.3)845 (90.2)116 (97.8)11 (98.5)2 (98.7)0 (98.7)1 (98.7)0 (98.7)19 (100.0)
    TMP-SMX0 (0.0)3 (0.2)174 (11.7)1,238 (93.5)69 (98.1)12 (98.9)2 (99.0)3 (99.2)2 (99.3)2 (99.5)1 (99.5)7 (100.0)
MRSA (3,003)
    Iclaprim2 (0.1)11 (0.4)370 (12.8)1,707 (69.6)646 (91.1)35 (92.3)20 (92.9)26 (93.8)23 (94.6)99 (97.9)64 (100.0)
    Trimethoprim0 (0.0)1 (<0.1)5 (0.2)166 (5.7)1,260 (47.7)1,226 (88.5)109 (92.1)11 (92.5)12 (92.9)22 (93.6)15 (94.1)13 (94.6)163 (100.0)
    TMP-SMX0 (0.0)2 (0.1)580 (19.4)1,824 (80.1)246 (88.3)164 (93.8)22 (94.5)22 (95.2)25 (96.1)17 (96.6)33 (97.7)68 (100.0)
GAS (604)
    Iclaprim184 (30.5)285 (77.6)104 (94.9)27 (99.3)4 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)
    Trimethoprim5 (0.8)47 (8.6)225 (45.9)234 (84.6)84 (98.5)6 (99.5)3 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)
    TMP-SMX0 (0.0)3 (0.5)58 (10.1)369 (71.2)153 (96.5)21 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)
GBS (204)
    Iclaprim0 (0.0)1 (0.5)0 (0.5)10 (5.4)102 (55.4)73 (91.2)18 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)
    Trimethoprim0 (0.0)0 (0.0)0 (0.0)1 (0.5)11 (5.9)92 (51.0)78 (89.2)21 (99.5)1 (100.0)0 (100.0)0 (100.0)0 (100.0)
    TMP-SMX0 (0.0)0 (0.0)7 (3.4)101 (52.9)92 (98.0)4 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)0 (100.0)
E. faecalis (310)
    Iclaprim79 (29.0)100 (57.7)19 (63.9)10 (67.1)6 (69.0)5 (70.6)0 (70.6)1 (71.0)3 (71.9)77 (96.8)6 (1.9)4 (100.0)
    Trimethoprim2 (0.6)11 (4.2)79 (29.7)83 (56.5)20 (62.9)5 (64.5)10 (67.7)7 (70.0)2 (70.6)0 (70.6)1 (71.0)6 (72.9)84 (100.0)
    TMP-SMX1 (0.3)43 (14.2)130 (56.1)32 (66.5)17 (71.9)11 (75.5)4 (76.8)4 (78.1)9 (81.0)4 (82.3)4 (83.5)
E. faecium (303)
    Iclaprim102 (33.7)8 (36.3)4 (37.6)0 (37.6)1 (38.0)1 (38.3)3 (39.3)9 (42.2)48 (58.1)51 (74.9)2 (75.6)
    Trimethoprim42 (13.9)50 (30.4)17 (36.0)4 (37.3)1 (37.6)1 (38.0)1 (38.3)2 (38.9)8 (41.6)17 (47.2)24 (55.1)17 (60.7)119 (100.0)
    TMP-SMX4 (1.3)10 (4.6)40 (17.8)47 (33.3)10 (36.6)5 (38.3)0 (38.3)0 (38.3)1 (38.6)1 (38.9)14 (43.6)171 (100.0)
Open in a separate windowaTMP-SMX, trimethoprim-sulfamethoxazole.b−, concentration not tested.cMICs greater than the highest concentration tested, which was 8 μg/ml for iclaprim and trimethoprim-sulfamethoxazole and 64 μg/ml for trimethoprim.Numbers of strains and activity summaries of all drugs tested are shown in Table Table1.1. Whereas the vast majority of methicillin-sensitive S. aureus (MSSA; 1,513 strains) strains were susceptible to most of the compounds tested in this study, a large proportion of MRSA strains were resistant to erythromycin (84.3%), clindamycin (47.0%), and both quinolones tested (83.1% and 82.4% for ciprofloxacin and levofloxacin, respectively) (Table (Table1).1). Iclaprim was highly active against both MSSA and MRSA (MIC50/MIC90, 0.06/0.12 μg/ml for both) (Tables (Tables11 and and2),2), with 98.7% of MSSA strains and 94.6% of MRSA strains being inhibited at a MIC of ≤2 μg/ml (Table (Table2).2). For group A streptococci (GAS), resistance rates to erythromycin, clindamycin, and tetracycline were higher in the European Union isolates (26.1%, 5.9%, and 14.2%, respectively) than those isolated in the United States (6.0%, 0.7%, and 5.0%, respectively; data not shown). Iclaprim inhibited 100% of GAS at an MIC of ≤0.12 μg/ml and an MIC50/MIC90 of 0.015/0.03 μg/ml (Tables (Tables11 and and2).2). After penicillin (MIC50/MIC90, ≤0.06/≤0.06 μg/ml) (Table (Table1),1), iclaprim was the most active among all the antibacterial agents tested. Unlike GAS, group B streptococci (GBS) from the United States showed higher rates of resistance to erythromycin (39.2% versus 12.7%), clindamycin (14.7% versus 5.9%), and tetracycline (82.4% versus 73.5%) than those isolated in the European Union (data not shown). Iclaprim was also highly potent against GBS (MIC50/MIC90, 0.12/0.25 μg/ml; MIC100, 0.5 μg/ml) (Tables (Tables11 and and2),2), and its activity was not affected by the organisms'' resistance to erythromycin, clindamycin, or tetracycline. As expected, E. faecalis isolates were generally more susceptible than Enterococcus faecium isolates to the drugs tested (Tables (Tables11 and and2),2), and resistance rates for E. faecalis did not vary much between the United States and European Union isolates (data not shown). Resistance rates were generally high among E. faecium isolates and differed significantly between those from the United States and European Union. Most notably, resistance to vancomycin was 70.8% among isolates in the United States compared to 14.8% among non-United States isolates. Iclaprim demonstrated the typical bimodal activity of its class against enterococci (Table (Table2).2). However, it showed high potency (MIC50/MIC90, 0.015/4 μg/ml) (Table (Table1)1) against E. faecalis, with approximately 72% and 97% of the isolates being inhibited at MICs of ≤2 and ≤4 μg/ml, respectively (Table (Table2).2). Iclaprim was also active (MIC50/MIC90, 2/>8 μg/ml) against E. faecium (Table (Table1),1), and its activity was not affected by resistance to vancomycin; approximately 58% and 75% of the isolates were inhibited at iclaprim MICs of ≤2 and ≤4 μg/ml, respectively (Table (Table22).Iclaprim was bactericidal against the S. aureus strains tested (Table (Table3).3). Against MRSA, iclaprim demonstrated MBC/MIC ratios of ≤4 for 100% of strains compared to 60% of strains for vancomycin. Against MSSA, iclaprim and vancomycin demonstrated MBC/MIC ratios of ≤4 for 86% of strains (Table (Table3).3). Iclaprim exhibited bactericidal activity against 45.0% of GAS and 65.0% of GBS, compared to ratios of ≥32 for vancomycin for most GAS and 100.0% of GBS tested (Table (Table3).3). Against enterococci, MBC/MIC ratios of ≤4 were seen in two out of five E. faecium strains and 3 out of 15 E. faecalis strains for iclaprim, whereas all the enterococcal strains tested exhibited MBC/MIC ratios of ≥32 for vancomycin (Table (Table3).3). The bactericidal activity of iclaprim when tested against staphylococci corroborates the results of previous studies (8, 10). In contrast, the fact that iclaprim demonstrated higher bactericidal activity against staphylococci than against streptococci and enterococci has not been previously reported and warrants further evaluation.

TABLE 3.

MBC/MIC ratios for iclaprim and vancomycin against 101 organisms
Organism (no. of isolates) and MBC/MIC ratioNo. of isolates at MBC/MIC ratio for indicated antimicrobial agent
IclaprimVancomycin
Staphylococcus aureus
    Methicillin-susceptible (21)
        1214
        272
        492
        811
        1600
        ≥3222
    Methicillin-resistant (20)
        148
        2131
        433
        804
        1601
        ≥3203
Beta-hemolytic streptococci
    Group A (20)
        110
        260
        420
        800
        1600
        ≥321118a
    Group B (20)
        110
        2100
        420
        810
        1600
        ≥32620
Enterococcus faecalis (15)
    100
    210
    420
    800
    1610
    ≥321115
Enterococcus faecium (5)
    100
    210
    410
    810
    1600
    ≥3225
Open in a separate windowaMBC could not be evaluated for two strains because both MIC and MBC values were beyond the dilution range tested (≤0.5 μg/ml).Results from this study support previously reported nonclinical susceptibility data (4, 7, 21) and demonstrate the potent bactericidal activity of iclaprim in vitro against both MSSA and MRSA. This finding, together with the high activity of iclaprim against beta-hemolytic streptococci and enterococcal species, confirms iclaprim as an important addition to the existing panel of therapies for the treatment of cSSSI caused by gram-positive organisms, including MRSA.  相似文献   

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

TABLE 1.

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

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