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 109 毫秒
1.
2.
An AmpC-type β-lactamase conferring high-level resistance to expanded-spectrum cephalosporins and monobactams was characterized from an Acinetobacter baumannii clinical isolate. This class C β-lactamase (named ADC-33) possessed a Pro210Arg substitution together with a duplication of an Ala residue at position 215 (inside the Ω-loop) compared to a reference AmpC cephalosporinase from A. baumannii. ADC-33 hydrolyzed ceftazidime, cefepime, and aztreonam at high levels, which allows the classification of this enzyme as an extended-spectrum AmpC (ESAC). Site-directed mutagenesis confirmed the role of both substitutions in its ESAC property.Acinetobacter baumannii is commonly associated with serious nosocomial infections (4, 11, 23). The emergence of multidrug-resistant A. baumannii strains, specifically those resistant to carbapenems, has created severe therapeutic challenges. A growing number of β-lactamases conferring resistance to expanded-spectrum cephalosporins have been identified in Acinetobacter spp. Resistance to oxyiminocephalosporins (ceftazidime and cefotaxime) is usually related to the overproduction of the resident AmpC-type β-lactamase (1, 2, 10) encoded by the gene blaAmpC. The overexpression of that gene has been associated with the insertion sequence ISAba1 providing a strong promoter (2, 9). Recently, the AmpC variants have been named according to a nomenclature specific to A. baumannii (ADC, for Acinetobacter-derived cephalosporinases [10]).Most AmpC-type β-lactamases naturally produced by Gram-negative bacteria hydrolyze amino- and ureidopenicillins, cephamycins (cefoxitin and cefotetan), and, at a lower level, oxyiminocephalosporins, such as ceftazidime, cefotaxime, and ceftriaxone, and monobactams, such as aztreonam (3). Zwitterionic cephalosporins (cefepime and cefpirome) together with carbapenems are usually excluded from the spectrum of activity of AmpC β-lactamases (8). However, natural cephalosporinases possessing broadened substrate activity in the taxa Enterobacteriaceae and Pseudomonas aeruginosa have been reported (12, 14, 22). These extended-spectrum AmpCs (ESACs) confer reduced susceptibility to all cephalosporins (12, 14, 22). They differ from “regular” cephalosporinases by amino acid substitutions or insertions/deletions in four specific regions that are all located in the vicinity of the active site: the Ω-loop, the H-10 helix, the H-2 helix, and the C-terminal extremity of the protein (12, 14, 22).Here, we describe the first ESAC-type β-lactamase from A. baumannii with a broadened substrate activity toward expanded-spectrum cephalosporins and monobactams.The A. baumannii KI clinical isolate was obtained from a rectal screening of a patient who had been transferred from the Guadeloupe Islands (French Caribbean islands) and admitted at the Raymond Poincaré Hospital (Garches, Paris, France) in January 2009. This patient did not receive any antibiotic treatment. A. baumannii KI was selected for further study on the basis of its uncommon pattern of resistance to β-lactam antibiotics, including high-level resistance to expanded-spectrum cephalosporins and reduced susceptibility to carbapenems. In addition, it was resistant to cotrimoxazole and fluoroquinolones and susceptible to aminoglycosides.Susceptibility testing was performed by use of the Etest (17) and interpreted according to the CLSI guidelines (5). It showed that A. baumannii KI was resistant to amino-, carboxy-, and ureidopenicillins, to expanded-spectrum cephalosporins, including ceftazidime, cefotaxime, ceftriaxone, cefepime, and cefpirome, and to aztreonam and had reduced susceptibilities to imipenem and meropenem (see Table Table2).2). By use of cloxacillin-containing plates as described previously (22), the susceptibilities to ceftazidime and cefepime were restored, suggesting (i) overproduction of the AmpC β-lactamase and (ii) the likelihood for this AmpC to possess ESAC properties. Mating-out assays and electrotransformation were performed as described previously (20) but failed to transfer any cephalosporin resistance marker from A. baumannii KI to A. baumannii BM4547 or to Escherichia coli TOP10 recipient strains, suggesting that this resistance to expanded-spectrum cephalosporins was chromosomally conferred. Whole-cell DNA of A. baumannii KI was extracted as described previously (22). Primers PreAmpC-PISAba1 and PreAmpC-Ab1 were used in combination with primer PreAmpC-Ab2 to amplify 1,521-bp and 1,254-bp fragments, respectively, encompassing the entire blaAmpC gene with and without the PISAba1 promoter, respectively (Table (Table11 ). All the inserts containing the PISAba1 promoter are named with “P+” accordingly. The amplification products were cloned into E. coli TOP10 by using the ZeroBluntTOPOPCR cloning kit (Invitrogen, Cergy-Pontoise, France) followed by selection on plates containing 50 μg/ml of amoxicillin and 30 μg/ml of kanamycin. A. baumannii strain AYE and the CIP7010 reference strain were used to clone two regular blaAmpC β-lactamase genes (7, 18). Strain AYE corresponds to a multidrug-resistant isolate from France (19) whose complete genome sequence was determined (6, 24). DNA sequence analysis showed that the β-lactamase from strain KI, named ADC-33, had 8 and 7 amino acid changes compared to regular β-lactamases ADC-50 and ADC-11, corresponding to the ADC β-lactamases of strains CIP7010 and AYE, respectively (Fig. (Fig.1).1). No amino acid change was identified among the conserved SVSK, YSN, and KTG motifs, neither in helix H-2 nor in helix H-10, previously associated with the extended-spectrum activity of ESACs (12, 14, 22). However, we identified a single amino acid substitution, Pro210Arg, associated with a duplication of the Ala residue at position 215, both of them located inside the Ω-loop. Thus, the molecular basis of the extended-spectrum hydrolysis profile of ADC-33 might be related to those specific substitutions. The presence of the insertion sequence ISAba1 providing strong promoter sequences (−35 [TTAGAA] and −10 [TTATTT]) immediately upstream of the blaAmpC gene was noticed, as previously observed when the blaAmpC gene was overexpressed (9).Open in a separate windowFIG. 1.Amino acid sequence alignment including ADC-33 (-33) as an ESAC, ADC-50 (-50) and ADC-11 (-11) as regular AmpCs, and ADC-7 (-7) taken as the reference sequence as published (10). Conserved amino acid identities are indicated by dashes. The typical AmpC β-lactamase domains (SVSK, YSN, and KTG) are underlined. Helices H-2 and H-10 are boxed in gray. The Ω-loop is boxed in gray and double underlined. Differences observed inside the Ω-loop are in boldface. The vertical arrow indicates the position of the +1 amino acid (cleavage site for signal peptide). Numbering is according to the sequence of the mature protein.

TABLE 1.

Primers used in this work
PrimerSequence (5′ to 3′)aPCR product size (bp)Purpose
PreAmpC-Ab1GAGCTAATCATGCGATTTAAA1,254Cloning AmpC coding sequence region
PreAmpC-Ab2GCTTAGGATATGTTTGGTTCTT
PreAmpC-PISAba1GACCTGCAAAGAAGCGCTGC1,521 (with PreAmC-Ab2)Cloning AmpC plus PISAba1
AmpC-Ab1GGAAAGGTTGTGGCTTTGTCTAmpC sequencing
Site-directed mutagenesis primers
    KI-A215-DelFGGC CCA CTC GAT GCC CCA GCA TAT GGCDeletion of Ala215
    KI-A215-DelRGCC ATA TGC TGG GGC ATC GAG TGG GCC
    Aye-A215-InsFGGC CCA CTC GAT GCC GCC CCA GCA TAT GGCInsertion of Ala215
    Aye-A215-InsRGCC ATA TGC TGG GGC GGC ATC GAG TGG GCC
    Aye-P210R-FATT CGA GTT AAC CGC GGC CCA CTC GAT GCCSubstitution Pro210Arg
    Aye-P210R-RGGC ATC GAG TGG GCC GCG GTT AAC TCG AAT
    KI-R210P-FATT CGA GTT AAC CCC GGC CCA CTC GAT GCCSubstitution Arg210Pro
    KI-R210P-RGGC ATC GAG TGG GCC GGG GTT AAC TCG AAT
Open in a separate windowaModified base pairs are underlined.

TABLE 2.

MICs of β-lactams for bacteriaa
β-Lactam(s)bMIC (μg/ml)
A. baumannii KIA. baumannii CIP70.10A. baumannii AYEE. coli TOP10 (pADC-33)E. coli TOP10 (pADC-50)E. coli TOP10 (pADC-11)E. coli TOP10 (pADC-33-P+)cE. coli TOP10 (pADC-11-P+)E. coli TOP10 (pADC-33-P+ (A215Del))dE. coli TOP10 (pADC-11-P+ (A215Ins))eE. coli TOP10 (pADC-33-P+(A215Del+ R210P))E. coli TOP10 (pADC-11-P+ (A215Ins+ P210R))E. coli TOP10 (pTOPO)
Amoxicillin>25632>256>256256>256>256>256>256>256>256>2562
Amoxicillin + CLA6416128>256256256>256>256>256>2562562562
Ticarcillin>2568>2566488>256256128>256128>2562
Ticarcillin + CLA6481286488>256256128>256128>2562
Piperacillin>25616>25632168>256>256>256>256>256>2561
Piperacillin + TZB128412832168>2561282562561282561
Cefuroxime>25632>256>25664128>256>256>256>256>256>2562
Ceftazidime>2562>256>2560.50.5>256816>2568>2560.125
Cefotaxime>2568>2561280.50.5>2568162568>2560.047
Cefepime2561>2560.1250.0320.03210.1250.250.50.12520.023
Cefpirome>2562>2560.50.0470.03240.1250.520.12580.023
Aztreonam2568>25680.1250.094256261622560.047
Imipenem20.2510.250.250.190.50.380.50.380.250.250.125
Meropenem20.2510.0230.0230.0230.0320.0320.0320.0320.0320.0320.023
Open in a separate windowaβ-Lactam MICs for A. baumannii KI, A. baumannii AYE expressing the ESBL VEB-1, reference strain A. baumannii CIP7010, E. coli TOP10 strains harboring recombinant plasmid pADC-33, pADC-11, pADC-50, pADC-33-P+, pADC-11-P+, pADC-33-P+(A215Del), pADC-11-P+(A215Ins), pADC-33-P+(A215Del+R210P), pADC-11-P+(A215Ins+P210R), and E. coli TOP10 reference strain.bCLA, clavulanic acid at a fixed concentration of 4 μg/ml; TZB, tazobactam at a fixed concentration of 4 μg/ml.cP+ corresponds to entire blaAmpC gene, including the PISAba1 promoter.dA215Del corresponds to the deletion of this alanine residue at position 215.eA215Ins corresponds to the insertion of this alanine residue at position 215.MICs of all β-lactams for the different clones were higher in the presence of the PISAba1 promoter (Table (Table2).2). E. coli TOP10(pADC-33-P+) was resistant to most β-lactams except cefepime, cefpirome, and carbapenems. E. coli TOP10(pADC-11-P+), producing a regular cephalosporinase, was resistant to amoxicillin and to narrow-spectrum cephalosporins and remained susceptible to the expanded-spectrum cephalosporins, including ceftazidime, cefotaxime, cefepime, and cefpirome, and aztreonam (Table (Table2).2). The recombinant E. coli strain harboring the blaADC-33 gene showed a broadened hydrolytic activity against ceftazidime, cefotaxime, cefepime, cefpirome, and aztreonam compared to that of the regular variant, increasing the resistance rates significantly (from 8- to 128-fold), suggesting that ADC-33 might be considered an ESAC enzyme. In order to better define the clinical impact of this ADC-33 variant, recombinant plasmids pADC-33-P+ and pADC-11-P+ were additionally electrotransformed into E. coli HB4 (lacking OmpC and OmpF porins) (13). By using E. coli HB4, our aim was to mirror the intrinsic weak permeability of A. baumannii and therefore better appreciate the impact of the ESAC enzymes. MICs of cefepime, cefpirome, and aztreonam for the ESAC and regular variants were 32, 128, and >256 μg/ml and 2, 4, and 12 μg/ml, respectively (data not shown). Those MIC values were higher than those observed for the wild-type E. coli strain, as expected. The significantly different impacts on the resistance level shown by the ESAC and non-ESAC AmpC variants in a porin-deficient E. coli strain would also likely apply to A. baumannii.Isoelectric focusing analysis performed as described previously (21) using crude culture extracts of A. baumannii KI and of the different clones obtained by sonication gave a single β-lactamase signal for all of them, corresponding to a pI value of 9.3, consistent with the production of an AmpC β-lactamase.β-Lactamases ADC-33 and ADC-11 were purified to near homogeneity (>99%) as previously described (20). The molecular mass of those proteins determined by SDS-PAGE analysis was 43 kDa. The specific activities, determined with 100 μM benzylpenicillin as the substrate, were 45 and 55 μmol/min/mg of protein for ADC-33 and ADC-11, respectively.Purified β-lactamases ADC-33 and ADC-11 were used for kinetic measurements (Km and kcat) calculated as described previously (20). The kinetic parameters for penicillins and carbapenems for the two AmpC enzymes were similar, whereas the catalytic efficiency of the purified β-lactamase ADC-33 against cephalosporins was higher than that of ADC-11 (Table (Table3).3). This higher catalytic efficiency observed for ADC-33 was noticeable in particular for ceftazidime, cefotaxime, cefepime, cefpirome, and aztreonam (Table (Table3).3). Those data confirmed the ESAC property of ADC-33, which associates two specific features, namely, (i) a greater ability to hydrolyze ceftazidime, cefotaxime, and aztreonam than a regular AmpC and (ii) an extended-spectrum activity toward cefepime and cefpirome, which are substrates usually weakly hydrolyzed by regular AmpCs.

TABLE 3.

Kinetic parameters of β-lactamases ADC-33 (ESAC) and ADC-11 of A. baumanniia
β-LactamADC-33
ADC-11
ADC-33/ADC-11 kcat/Km ratio
Km (μM)kcat (s−1)kcat/Km (μM−1 s−1)Km (μM)kcat (s−1)kcat/Km (μM−1 s−1)
Benzylpenicillin3155124541.3
Ampicillin940.57510.5
Piperacillin1640.252550.21.2
Cephaloridine10022022506002.50.8
Cephalothin12014012105502.50.5
Cefoxitinb0.10.330.51.531
Cefotaximeb0.5122.50.20.120
Ceftazidime3040.15100.010.001150
Cefepime1,300100.011,80010.000520
Cefpirome700350.051,70080.00412.5
Aztreonamb0.0070.01523NDcND
Imipenem50.010.00260.0150.00250.8
Open in a separate windowaData are the means of three independent experiments. Standard deviations were within 15% of the means.bFor β-lactams with a Km value less than 5 μM, Ki values were determined instead of Km values, with cephalothin used as the substrate.cND, no detectable hydrolysis (<0.01 s−1), for a maximum amount of 5 μg of purified enzyme and up to 200 nmol of substrate.No significant difference in 50% inhibitory concentrations (IC50s) for clavulanic acid was observed between the two enzymes (1.6 and 2.1 mM, for ADC-33 and ADC-11, respectively).Compared to the regular ADC-11 β-lactamase, ADC-33 contained several amino acid substitutions, including the substitution Pro210Arg together with a duplication of the Ala residue at position 215 located inside the Ω-loop. A site-directed mutagenesis strategy was used with a protocol described by the manufacturer (QuikChange site-directed mutagenesis kit; Stratagene) and as reported previously (21) for evaluation of the amino acid changes that could be involved in the extended-spectrum resistance spectrum of ADC-33. By use of recombinant plasmids pADC-33-P+ and pADC-11-P+ as templates and primers KI-A215-DelF/KI-A215-DelR and Aye-A215-InsF/Aye-A215-InsR, deletion and insertion of the Ala215 residue, respectively, were performed (Table (Table1).1). This allowed us to obtain recombinant plasmids pADC-33-P+(A215Del) (with a single A215 residue) and pADC-11-P+(A215Ins) (with two A215 residues). Then, the Arg210Pro and Pro210Arg mutants were generated by using primer pairs KI-R210P-F/KI-R210P-R and Aye-P210R-F/Aye-P210R-R, respectively (Table (Table1),1), with other recombinant strains obtained in a second step as indicated in Table Table2.2. Sequence analysis of the different inserts confirmed the presence of the expected amino acid changes.Duplication of the Ala215 residue in β-lactamase ADC-11 resulted in a pattern of increased resistance toward expanded-spectrum cephalosporins (Table (Table2)2) which was, however, not sufficient to explain MIC differences observed between E. coli(pADC-11-P+) and E. coli(pADC-33-P+). Thus, as a second step, the Pro210Arg substitution was performed, allowing us to achieve an ESAC phenotype identical to that of ADC-33. Conversely, the Arg210Pro substitution performed with ADC-33 allowed the recovery of the regular AmpC phenotype, supporting the hypothesis that both amino acid modifications (Ala215 duplication and Pro210Arg substitution) were responsible for the extended-spectrum profile of ADC-33 (Table (Table2).2). Whether the other identified amino acid differences between ADC-33 and ADC-11 (Fig. (Fig.1)1) could play a specific role in the functionality or stability of β-lactamase ADC-33 remains unclear.Extension of the Ω-loop for AmpC β-lactamase from Enterobacter cloacae has been shown to broaden its hydrolysis spectrum (15, 16). In that case, the mutant enzyme exhibited an increased opening of the entrance of the substrate-binding pocket (6). The Ala215 duplication observed here in AmpC ADC-33 might have similar consequences for its hydrolysis spectrum.We describe here the first ESAC conferring resistance to expanded-spectrum cephalosporins in A. baumannii after the reports of those in Enterobacteriaceae and, recently, in P. aeruginosa. Noticeably, ADC-33 expression did not have any impact on carbapenem resistance. Clinical implications and spread of this resistance trait shall now be evaluated with A. baumannii isolates of worldwide origins.  相似文献   

3.
Van-M-02, a novel glycopeptide, was revealed to exert potent activities against Gram-positive bacteria, including vancomycin-resistant enterococci (VRE) and vancomycin-resistant Staphylococcus aureus (VRSA). A crude assay system was then used to study the mode of action of Van-M-02 as a peptidoglycan synthesis model of both vancomycin-susceptible and -resistant strains. The results suggested that Van-M-02 inhibits the synthesis of lipid intermediates irrespective of their termini. This inhibitory activity may contribute to the anti-VRE and anti-VRSA activities observed.The increasing incidence of vancomycin resistance in clinical settings has prompted research into new antibiotics against vancomycin-resistant strains (7, 12). We previously reported the synthesis of a novel glycopeptide, Van-M-02 (Fig. (Fig.1),1), during the course of our study of a vancomycin dimer (8). In this report, we describe the potent activities of Van-M-02 against the Gram-positive bacteria, including vancomycin-resistant enterococci (VRE) and vancomycin-resistant Staphylococcus aureus (VRSA), and the investigation of its mode of action using a crude assay system.Open in a separate windowFIG. 1.Structures of Van-M-02, ΔN-Van-M-02, vancomycin, and ΔN-vancomycin.MICs were determined by the broth dilution method in accordance with CLSI document M7-A7 (2). Van-M-02 showed potent activities against VRSA, VanA-type VRE, and constitutive VanB-type VRE, with MICs of 4 μg/ml or less despite its structural similarity to vancomycin (Table (Table1).1). In order to assess the contribution of the d-Ala-d-Ala binding pocket to the activity of Van-M-02, N-terminally degraded Van-M-02 (ΔN-Van-M-02, Fig. Fig.1)1) was prepared (see Materials and Methods in the supplemental material). In a previous report, the corresponding compound ΔN-vancomycin (6) lost its antibacterial activities even against vancomycin-sensitive strains due to the lack of tight binding to d-Ala-d-Ala. In the present study, however, ΔN-Van-M-02 was effective against both vancomycin-sensitive (MIC = 2 μg/ml) and -resistant strains (MICs of 4 to 16 μg/ml).

TABLE 1.

Antibacterial activities of Van-M-02 and ΔN-Van-M-02
StrainRelevant phenotypeaMIC (μg/ml)
Sourceb or reference
Van-M-02ΔN-Van-M-02VancomycinΔN-vancomycin
Staphylococcus aureus
    RN4220Mcs0.12521>643
    SmithMcs0.12521>645
    SR3637Mcr0.2522>649
    HIP11714Vmr (VanA)416>64>641
    ATCC 700787VISA248>6410
Enterococcus faecium
    SR16972Vms0.2520.5>64Clinical isolate
    SR7940Vmr (VanA)18>64>64Clinical isolate
    SR23598Vmr (VanB)0.254>64>64Clinical isolate
    SRM1101Vmr (VanB const)c416>64>64This study
Enterococcus faecalis
    SR1004Vms0.2522>64Clinical isolate
    SR7914Vmr (VanA)28>64>64Clinical isolate
    SR23630Vmr (VanB)0.5432>64Clinical isolate
Open in a separate windowaAbbreviations: Mc, methicillin; Vm, vancomycin; s, susceptible; r, resistant; VISA, vancomycin-intermediate Staphylococcus aureus.bClinical isolates were from Japan.cConstitutively vancomycin-resistant strain.To further investigate the mode of action of Van-M-02, a macromolecule precursor incorporation assay was performed. Van-M-02 specifically inhibited the incorporation of N-acetyl-d-[1-3H]glucosamine into cells of S. aureus and VRE (see Materials and Methods and Fig. S1A and D in the supplemental material). ΔN-Van-M-02 retained the specific inhibition of incorporation (see Fig. S1B and E in the supplemental material). These results suggested that cell wall synthesis is the primary target by which Van-M-02 exerts its antibacterial activity.We next evaluated the inhibition of in vitro peptidoglycan synthesis by Van-M-02 with a crude assay system using wall membrane particulate-containing membrane enzymes of S. aureus and UDP-MurNAc-pentapeptide (for the susceptibility model) or UDP-MurNAc-depsipeptide (for the resistance model). UDP-MurNAc-pentapeptide (d-Ala-d-Ala terminus) was prepared from S. aureus, and UDP-MurNAc-depsipeptide (d-Ala-d-lactate terminus) was prepared from VRE (see Materials and Methods in the supplemental material). The formation of lipid intermediates (d-Ala-d-Ala terminus or d-Ala-d-lactate terminus) and nascent peptidoglycan (d-Ala-d-Ala terminus or d-Ala-d-lactate terminus) via successive reactions catalyzed by several enzymes, i.e., MraY, MurG, FemX, FemA, FemB, and transglycosylase, etc., were detected by thin-layer chromatography (TLC) utilizing the incorporation of [14C]glycine into them (see Fig. S2 in the supplemental material). In this report, “lipid intermediates” refers to lipid I and lipid II. Among the lipid intermediates, only lipid II can be glycylated, as reported previously (11). Lipid II was the lipid intermediate detected by TLC in the present study. The amounts of radioactivity incorporated into lipid intermediates and peptidoglycan in the resistance model were slightly reduced, by 14% and 22%, respectively, compared to those in the susceptibility model, which could result in a reduction of the growth rate of VRSA (4). To the best of our knowledge, this is the first report that the UDP-MurNAc-depsipeptide served as a substrate in an enzymatic assay for the peptidoglycan synthetic pathway in S. aureus.In the susceptibility model of the crude assay system, the inhibition of peptidoglycan formation by vancomycin was far more potent (50% inhibitory concentration [IC50] = 4.5 ± 0.4 μg/ml) than that of the lipid intermediates (IC50 = 180 ± 46 μg/ml) (9). This result is reasonable, because the primary target of vancomycin is transglycosylase, an enzyme catalyzing the formation of peptidoglycan from lipid intermediates. The inhibition of peptidoglycan formation by Van-M-02 was similarly potent (IC50 = 4.2 ± 0.1 μg/ml). The inhibition of lipid intermediate formation was moderate in this case (IC50 = 32 ± 7.0 μg/ml) (Table (Table2).2). Thus, peptidoglycan polymerization by transglycosylase may also be a primary target of Van-M-02 in the susceptibility model. In the resistance model, on the other hand, the levels of the inhibitory activities of Van-M-02 were not very different between lipid intermediate formation (IC50 = 76 ± 0.6 μg/ml) and peptidoglycan formation (IC50 = 36 ± 4.2 μg/ml). Therefore, the processes of lipid intermediate formation would also be important targets of Van-M-02, in addition to transglycosylase, in the resistance model. It is also noteworthy in lipid intermediate formation that the IC50 of Van-M-02 (IC50 = 76 ± 0.6 μg/ml) was significantly lower than that of vancomycin (IC50 = 2,300 ± 1,400 μg/ml). ΔN-Van-M-02, which lacks a d-Ala-d-Ala binding motif, also showed inhibition of peptidoglycan and lipid intermediate formation in both the susceptibility and resistance models (Table (Table22).

TABLE 2.

Antibacterial activities and inhibition of lipid intermediates and peptidoglycan formation by antibiotics
AntibioticMIC (μg/ml)
IC50 (μg/ml)
Lipid intermediate formation
Peptidoglycan formation
RN4220HIP11714Susceptible modelResistant modelSusceptible modelResistant model
Van-M-020.125432 ± 7.076 ± 0.64.2 ± 0.136 ± 4.2
ΔN-Van-M-02216140 ± 9593 ± 2811 ± 0.753 ± 18
Vancomycin1>64180 ± 462,300 ± 1,4004.5 ± 0.485 ± 28
ΔN-vancomycin>64>643,900 ± 183,100 ± 1,100130 ± 20920 ± 700
Open in a separate windowWe next examined whether the antibacterial activity of Van-M-02 was antagonized by external Nα,Nɛ-diacetyl-l-Lys-d-Ala-d-Ala (dKAA) or Nα,Nɛ-diacetyl-l-Lys-d-Ala-d-lactate (dKAL) in order to investigate the contribution of substrate-binding properties to the mechanism of Van-M-02. In the presence of 64 μg/ml dKAA, the antibacterial activity of vancomycin was decreased considerably (MIC = 16 μg/ml) but the activity of Van-M-02 was not substantially affected (MIC = 0.5 μg/ml) (Table (Table3).3). Thus, Van-M-02 should exert its antibacterial activities via a less substrate-dependent mechanism than vancomycin. Since dKAA had a smaller effect on the antibacterial activity of Van-M-02 than on that of vancomycin (Table (Table3),3), the potent activity of Van-M-02 would not be attributable to its enhanced binding affinity for the d-Ala-d-Ala terminus of the lipid intermediates of vancomycin-susceptible strains. In other words, another target of Van-M-02 may exist.

TABLE 3.

Antagonism of the antibacterial activity of antibiotics by externally added dKAA or dKAL
AntibioticMIC (μg/ml) with:
NAadKAA at:
dKAL at:
8 μg/ml64 μg/ml8 μg/ml64 μg/ml
Van-M-020.1250.1250.50.1250.0625
ΔN-Van-M-0222221
Vancomycin141610.5
Open in a separate windowaNA, no addition.Since dKAL did not decrease the antibacterial activities of Van-M-02 (Table (Table3),3), the activities of Van-M-02 would not be attributable to high binding affinity for the d-Ala-d-lactate terminus of the lipid intermediates of VRE (depsilipids I and II).These in vitro crude assays and antagonism studies suggested that Van-M-02 inhibits the lipid intermediate synthesis irrespective of the termini of lipid intermediates in bacteria. The inhibitory activity may contribute to anti-VRE and -VRSA activities.Although the modified semiquantitative crude assay system enabled us to obtain insights into the actions of Van-M-02, their molecular mechanisms have not been fully characterized. Elucidation of the molecular target(s) of Van-M-02 would be a worthy topic of future research. Possible targets would be (i) enzymes catalyzing the synthesis of lipid intermediates (MurG, etc.) and transglycosylase, (ii) lipid intermediates, or (iii) the bacterial membrane.In conclusion, inhibition of lipid intermediate synthesis should be considered a possible antibacterial target of Van-M-02 for vancomycin-susceptible and -resistant strains. We are currently working to establish quantitative assays with purified enzymes involved in cell wall synthesis (MurG, transglycosylase, etc.) and their substrates (depsilipid I, depsilipid II, etc.) that should shed light on the mechanism of action of modified vancomycins.   相似文献   

4.
In this study, application of a dual absorbance/fluorescence assay to a chemical library screen identified several previously unknown inhibitors of mycobacteria. In addition, growth conditions had a significant effect on the activity profile of the library. Some inhibitors such as Se-methylselenocysteine were detected only when screening was performed under nutrient-limited culture conditions as opposed to nutrient-rich culture conditions. We propose that multiple culture condition library screening is required for complete inhibitory profiling and for maximal antimycobacterial compound detection.Recent data from the World Health Organization show that there are more than 9 million new cases of tuberculosis (TB) each year, half a million of which are caused by drug-resistant Mycobacterium tuberculosis (26, 27, 28). The spread of antibiotic resistance has necessitated the identification of new anti-infective molecules for the treatment of TB (25). TB drug discovery research is often dependent on the robustness of the upstream biological assay used. In terms of whole-cell antimycobacterial assays, screens are commonly performed under optimal growth conditions such as in the presence of excess nutrients. Evidence suggests, however, that M. tuberculosis persists in a nutrient-deprived state in the host lung (4, 12-14, 21). Screening under nutrient-rich conditions may fail to detect compounds that are preferentially active against nutrient-limited M. tuberculosis during infection.Traditionally, antimycobacterial assays use optical density (OD) as an indicator of growth, which can be distorted by the intrinsic absorbance of some compounds and the propensity of mycobacterial cells to form aggregates. Alternatives to OD measurement include the use of reporter molecules, such as the green fluorescent protein (GFP). Collins et al. (10) demonstrated that when expressed in M. tuberculosis, the levels of GFP paralleled the numbers of CFU during growth. The MICs of a range of antitubercular drugs determined using GFP were consistent with those obtained using Alamar Blue (7) and the BACTEC 460 system (9). In this work, we used both OD and GFP fluorescence to screen the library of pharmacologically active compounds (LOPAC) (LO1280; Sigma-Aldrich, St. Louis, MO) for antimycobacterial compounds.For expression of GFP, a vector was constructed by PCR amplification of the gfpmut2 gene (11) from pOT11 (19) using primers GFP_RBS_F1 (5′-GGGGGTACCTTTAAGAAGATATACATATGAGTAAAGGAGAA-3′) and GFP_R1 (5′-GGGGGCATGCTTATTATTTGTATAGTTCATCCATGCC-3′). The product was cloned into the KpnI and SphI restriction sites of pTKmx (16), generating plasmid pTKmxGFP. The pAL5000 origin of replication of pTKmxGFP was excised by NheI restriction digestion and replaced with the replicon of the high-copy-number plasmid pHIGH100 (5), amplified using PCR primers OriM_F (5′-GGGGGCTAGCAACGAGGACAGTCGCACGAC-3′) and OriM_R (5′-GGGGGCTAGCATCGAGCCGAGAACGTTATC-3′), generating plasmid pSHIGH. The hsp60 gene promoter from Mycobacterium bovis BCG, amplified using PCR primers Hsp60_F (5′-GGGGGGTACCGGTACCGGTGACCACAACGACGCGCCCGCT-3′) and Hsp60_R (5′-GGGGGGTACCCGCAATTGTCTTGGCCATTGCGAA-3′), was cloned into the KpnI site of pSHIGH, generating plasmid pSHIGH+hsp60. (Underlining indicates position of restriction site for each primer.)Mycobacterium smegmatis mc2155 harboring plasmid pSHIGH+hsp60 was inoculated into Luria-Bertani broth (LB) containing 50 μg/ml kanamycin and supplemented with 0.1% (vol/vol) Tween 80 and 100 μg/ml d-arabinose to reduce cell aggregation as previously described (2, 20). A selection of first- and second-line antitubercular drugs and tetracycline were used to test the validity of the antimycobacterial assay. The cultures were grown to the mid-logarithmic phase and diluted to an OD at 600 nm of 0.2 (10-mm path length). Two hundred microliters of sterile deionized water was added to each well on the perimeter of the plate to minimize evaporation of the growth medium during the assay. Fifty microliters of LB containing 50 μg/ml kanamycin, 0.1% Tween 80, and 100 μg/ml d-arabinose were added to the remaining wells. Starting at 50 μM, twofold serial dilutions were performed for the experimental compounds and control antibiotics. Fifty microliters of the cell culture, corresponding to approximately 5 × 106 CFU per well, was added to each one of the inner wells, except the medium control wells. The plates were sealed, wrapped in parafilm, and incubated at 37°C for 96 h with 200 rpm shaking. OD and GFP fluorescence measurements were performed at 0- and 96-h incubation using a Wallac Envision multilabel plate reader (Perkin-Elmer). Data were analyzed with SigmaPlot 11 (SYSTAT) using four-parameter logistic standard curve analysis. The MIC and 50% inhibitory concentration (IC50) values were determined with respect to the controls at 96 h. For each of the drugs tested, use of OD and GFP measurements produced identical MICs and good correlation for the IC50s (Table (Table11 and Fig. Fig.11).Open in a separate windowFIG. 1.(A) Correlation of the data from the OD and GFP fluorescence-based library screens. The LOPAC inhibitory data from the OD and GFP fluorescence assays were compared to determine the level of correlation in the presence of different growth media. Pearson coefficients of r = 0.64 (nutrient-rich culture conditions), r = 0.5 (carbon-limited culture conditions), and r = 0.33 (nitrogen-limited culture conditions) were obtained indicating medium to large positive correlation between the two assays. (B) Profile of LOPAC inhibitory activity under various growth conditions. Examination of the LOPAC library index determined that a number of compounds were not detected in all of the culture conditions. This resulted in modifications in the inhibitory profile of the LOPAC library as a function of various growth conditions. Mycobacterium smegmatis was grown under nutrient-rich (LB), carbon-limited (C−), and nitrogen-limited (N−) culture conditions with each LOPAC compound present at a single concentration of 20 μM. Inhibitory activity is expressed as percentage inhibition of M. smegmatis growth. Symbols: ○, LB medium; ▵, carbon-limited culture conditions; □, nitrogen-limited culture conditions. The opacity of symbols increases with increasing inhibition. Significant results in both axes are displayed with solid symbols, i.e., LB medium (•), carbon-limited culture conditions (▴), nitrogen-limited culture conditions (▪). Scatter plots were drawn using Scatterplot3d (17).

TABLE 1.

Comparison of the optical density- and fluorescence-based inhibitor assays using known antibiotics in Mycobacterium smegmatisa
AntibioticMIC (μM) by OD and fluorescence assaysIC50 (μM) (mean ± SEM) by the following assay:
ODFluorescence
Capreomycin12.52.12 ± 0.691.78 ± 0.71
Ciprofloxacin12.51.98 ± 0.251.07 ± 0.24
Ethambutol12.53.66 ± 0.103.20 ± 0.13
Ethionamide1007.66 ± 0.156.75 ± 0.75
Rifampin6.254.36 ± 0.493.11 ± 0.51
Streptomycin1.560.50 ± 0.090.48 ± 0.05
Tetracycline6.250.41 ± 0.200.33 ± 0.30
Open in a separate windowaMIC and IC50s for M. smegmatis were compared using both OD and fluorescence measurements for a suite of known antibiotics including first- and second-line antitubercular drugs. Use of OD and fluorescence measurements produced the same MIC and good correlation in terms of the IC50s for all of the antibiotics tested.The LOPAC library was screened for inhibitory activity toward M. smegmatis grown under nutrient-rich conditions. Two microliters of a 1 mM concentration of each compound from the 16 LOPAC stock plates was transferred to the wells of columns 2 to 11 using a Cybi-Well robotic liquid handling station (Cybio) to obtain a final chemical concentration of 20 μM. Two hundred microliters of sterile-distilled water were added to each well in column 1 to minimize evaporation and medium, solvent, and antibiotic controls were established in column 12 of each plate. Solvent controls, consisting of 2% dimethyl sulfoxide, did not produce any significant inhibition of M. smegmatis growth. Antibiotic controls, consisting of rifampin (rifampicin) and ciprofloxacin at 20 μM, yielded complete growth inhibition.The library screens were performed three times, and the resulting data were normalized to control for plate-to-plate variation by linear scaling to match the most extreme plates in each data set and taking their natural log transforms. Control values that were more than 3 standard deviations from the mean were considered outliers and discarded. Z-factors were calculated for each assay control to give an indication of assay reliability, taking into account both dynamic range and assay variability (29). Under standard growth conditions, the Z-factors for capreomycin were Z = 0.87 (OD) and Z = 0.88 (GFP) and the Z-factors for rifampin were Z = 0.68 (OD) and Z = 0.87 (GFP), interpreted as an excellent assay. To enable comparison with previous work (7), the signal-to-noise ratios for capreomycin were 23.21 (OD) and 24.20 (GFP) and the ratios for rifampin were 9.44 (OD) and 22.78 (GFP). Following validation of potential hits using MIC and IC50 determination, 14 compounds that inhibited M. smegmatis growth under nutrient-rich conditions at concentrations of 12.5 μM or below were identified (Table (Table2).2). A number of these compounds, e.g., calcimycin, demeclocycline, doxycycline, lomefloxacin, minocycline, ofloxacin, and vancomycin, are known antibacterials, and therefore, activity against mycobacteria was not unexpected. In addition, antimycobacterial activity has previously been recorded for clotrimazole and niclosamide by the Southern Research Institute in Alabama (http://pubchem.ncbi.nlm.nih.gov) and other researchers (6, 22). From our review of the literature, antimycobacterial activity has not been previously reported for the remaining compounds, calmidazolium, diphenyleneiodonium, idarubicin, and methoctramine.

TABLE 2.

Validated inhibitors from the LOPAC chemical library for Mycobacterium smegmatis grown under nutrient-rich and nutrient-limited culture conditionsa
LOPAC compoundMIC (μM)b,c
IC50 (μM) (mean ± SEM)b
LBC−N−LBC−N−
BAY 11-7085506.2512.530.60 ± 0.592.58 ± 0.486.22 ± 0.11
Calcimycin3.12512.512.51.04 ± 0.245.21 ± 0.163.43 ± 0.19
Calmidazolium chloride12.55012.513.20 ± 1.1713.85 ± 0.227.02 ± 1.51
Carboplatin50255014.07 ± 4.2712.14 ± 0.677.41 ± 0.25
4-Chloromercuribenzoic acid256.2512.57.07 ± 0.757.65 ± 0.176.33 ± 0.11
Cisplatin>5012.512.5>508.02 ± 3.4410.31 ± 5.45
Clotrimazole12.550506.54 ± 0.7718.21 ± 3.2412.19 ± 0.07
Demeclocycline3.1256.256.251.88 ± 0.245.92 ± 0.284.40 ± 0.25
Dequalinium analog C-146.256.253.1253.95 ± 0.552.73 ± 0.401.74 ± 0.30
Diphenyleneiodonium6.253.1253.1256.77 ± 0.438.53 ± 1.811.63 ± 0.30
Doxycycline1.563.1253.1250.76 ± 0.052.98 ± 0.161.95 ± 0.57
Idarubicin12.52512.53.61 ± 2.016.68 ± 6.535.13 ± 0.23
Lomefloxacin12.56.256.254.33 ± 0.104.88 ± 0.251.44 ± 0.15
LY-36726550255025.11 ± 6.9625.26 ± 1.4422.31 ± 0.34
Methoctramine12.512.512.515.90 ± 1.986.88 ± 1.227.45 ± 0.40
Minocycline12.512.512.56.77 ± 1.125.02 ± 1.157.23 ± 0.13
Mitoxantrone50505021.84 ± 5.279.70 ± 0.845.02 ± 0.25
Niclosamide12.56.256.256.85 ± 0.723.31 ± 1.214.73 ± 0.26
Ofloxacin3.1256.256.252.03 ± 0.653.13 ± 1.551.73 ± 0.37
Pentamidine50255023.53 ± 0.1511.05 ± 0.179.10 ± 0.26
1,10-Phenanthroline5012.512.541.64 ± 0.314.99 ± 0.285.76 ± 0.32
Se-methylselenocysteine>5012.512.5>502.19 ± 0.180.60 ± 0.09
Ruthenium red>505050>5017.71 ± 0.199.47 ± 0.11
Trifluoperazine50505012.26 ± 0.0612.35 ± 0.183.15 ± 1.04
U-83836505050>50>5034.25 ± 0.38
Vancomycin12.525253.18 ± 0.4316.69 ± 0.1911.12 ± 0.17
WB645050258.15 ± 1.3711.70 ± 5.736.00 ± 0.12
Open in a separate windowaMIC and IC50s for M. smegmatis were determined using OD and/or fluorescence measurements for compounds that were detected as inhibitory in the LOPAC library screens. Validation of the hits was carried out using a starting concentration of 50 μM, and compounds were tested against M. smegmatis grown under nutrient-rich, carbon-limited, and nitrogen-limited culture conditions.bThe mycobacteria were grown under nutrient-rich (LB) and carbon-limited (C−) and nitrogen-limited (N−) culture conditions.cUnderlined MIC values indicate the growth condition under which an inhibitor was detected in screening.In the next stage of this work, we tested whether the inhibitory profile of the LOPAC library toward mycobacteria varied significantly under different culture conditions. M. smegmatis was grown in carbon- and nitrogen-limited Hartman-de Bonts (HdeB) medium (18, 21) in place of LB. Many hit compounds detected in rich media were also active under conditions of carbon and nitrogen limitation (Table (Table2).2). Clotrimazole was more active under nutrient-rich conditions. However, the screens identified additional inhibitors that were not detected using nutrient-rich culture conditions. These include Bay 11-7085, 4-chloromercuribenzoic, cisplatin, and Se-methylselenocysteine, which exhibit significantly lower MIC values under nutrient limitation (Table (Table2).2). These findings indicate that antimycobacterial activity contained within chemical libraries is significantly affected by the growth conditions used for screening. Screening protocols need to be able to detect drugs that are active under nutrient limitation and other conditions that are considered relevant to the host environment.Of the compounds detected in the library screens against mycobacteria, a number of compounds may have activities relevant to TB drug development. Se-methylselenocysteine (MeSeCys) is used in the dietary chemoprevention of tumors (1). It is metabolized by selenocysteine lyase (β-lyase) producing methylselenol, which causes apoptosis in cancer cells by redox cycling and protein thiol modification (1, 15, 23). Selenocysteine lyases are widely distributed among bacteria (8). Enzymes annotated at the National Center for Biotechnology Information (http://www.ncbi.nlm.nih.gov) as selenocysteine lyases are encoded in the genomes of a number of mycobacterial species, including M. smegmatis (e.g., genes MSMEG_1242 and MSMEG_4538) and M. tuberculosis (e.g., genes Rv1464 and Rv3025c). Hence, it is plausible that mycobacteria could convert MeSeCys into toxic selenium products, such as methylselenol. Mycobacteria may potentially use MeSeCys as a source of selenium or other nutrients, which may account for the more pronounced inhibitory activity seen for MeSeCys in nutrient-deprived conditions (Fig. (Fig.22).Open in a separate windowFIG. 2.Activity of the compound Se-methylselenocysteine against Mycobacterium smegmatis. M. smegmatis was grown under nutrient-rich (LB), carbon-limited (C−), and nitrogen-limited (N−) culture conditions. (A) Dose-response curve for Se-methylselenocysteine obtained using the optical density-based assay. (B) Dose-response curve for Se-methylselenocysteine obtained using the GFP fluorescence-based assay. (C) Illustration of a possible metabolic reaction through which methyl selenol, a toxic form of selenium, could be generated from Se-methylselenocysteine by selenocysteine lyase in mycobacteria.We have been unable to find reports of inhibitory activity of MeSeCys or any organic forms of selenium toward mycobacteria or their use in the treatment of TB. Micronutrient supplementation with inorganic selenium can improve health outcomes for patients with TB by improving the nutritional status of the host (3, 24); however, an antibiotic role for selenium-based compounds has not been explored previously. Therefore, further investigations on MeSeCys as a potential antitubercular compound are being conducted.In conclusion, the use of different growth media exerts a significant effect on the identification of active compounds in a chemical library screen. Incorporation of host-related physicochemical conditions into whole-cell screens could potentially augment the detection of alternative compounds that are active against M. tuberculosis infection.  相似文献   

5.
Ceftobiprole is a new cephalosporin that exhibits a high level of affinity for methicillin-resistant Staphylococcus aureus PBP 2a. It was reported that ceftobiprole did not interact with a mutated form of the low-affinity protein Enterococcus faecium PBP 5 (PBP 5fm) that, when overexpressed, confers a β-lactam resistance phenotype to the bacterium. Our results show that ceftobiprole binds to unmutated PBP 5fm to form a stable acyl-enzyme and that ceftobiprole is able to efficiently kill a penicillin-resistant Enterococcus faecium strain that produces this protein.β-Lactam antibiotics (penicillins, cephalosporins, carbapenems, and monobactams) are the most frequently prescribed antibacterial agents used to fight serious bacterial infections. They inactivate the membrane-bound d,d-transpeptidases essential for peptidoglycan synthesis by forming with them stable acyl-enzymes (9). This explains why these enzymes are generally designated penicillin-binding proteins (PBPs) (17, 19). In Gram-positive cocci, the resistance to β-lactam antibiotics is primarily conferred by the presence or the overproduction of a low-affinity PBP. In enterococci, among which is the opportunistic human pathogen Enterococcus faecium (3), the resistance to penicillin may be associated with overproduction of the intrinsic low-affinity protein PBP 5 (PBP 5fm) or to other alterations affecting PBP 5fm (15).Ceftobiprole (BPR) (BAL9141) is a novel broad-spectrum cephalosporin that is active against Gram-positive and Gram-negative bacterial groups, including methicillin-resistant Staphylococcus aureus (MRSA) (2, 6, 10, 12) and Enterococcus faecalis (1). It has been demonstrated to be a good inhibitor of the S. aureus low-affinity protein PBP 2a (6, 8). A previous study reported that ceftobiprole had poor inhibitory activity against β-lactam-resistant E. faecium and that ceftobiprole did not bind to the mutated, low-affinity PBP 5fm protein isolated from such a strain (10). To better understand the difference between these two low-affinity PBPs with respect to ceftobiprole, we have characterized the interaction between an unmutated form of PBP 5fm and ceftobiprole.To perform this study, we used the penicillin-sensitive Enterococcus faecium strain D63 (benzylpenicillin MIC = 5 μg/ml and ampicillin MIC = 15 μg/ml) and its laboratory-derived penicillin-resistant strain D63r (benzylpenicillin MIC = 70 μg/ml and ampicillin MIC = 125 μg/ml) (21). The PBP profile of the latter strain exhibits a 6-fold increase in quantity of PBP 5fm (21). The ceftobiprole MICs determined by the microdilution method (4) for E. faecium D63r and D63 were 8 and 2 μg/ml, respectively. These values contrasted with those reported by other workers, who have found that most ampicillin-resistant E. faecium clinical isolates were resistant to ceftobiprole (13). They concluded that ceftobiprole was ineffective against ampicillin-resistant enterococcal strains.To further study the killing effects of benzylpenicillin and ceftobiprole on E. faecium D63r, we exposed exponentially growing cultures of both the sensitive and the resistant strains to increasing concentrations of antibiotic corresponding to 1 and 4 times the respective MICs (Fig. (Fig.1)1) (18). For concentrations higher than their MICs, benzylpenicillin and ceftobiprole show killing effects (Fig. (Fig.11).Open in a separate windowFIG. 1.Time-kill curves for resistant E. faecium D63r and susceptible E. faecium D63 in the presence of benzylpenicillin or ceftobiprole. The MICs for D63r were 70 μg/ml for benzylpenicillin and 8 μg/ml for ceftobiprole. The MICs for D63 were 5 μg/ml for benzylpenicillin and 2 μg/ml for ceftobiprole. The surviving bacteria were counted after 0, 4, and 24 h of incubation at 37°C by subculturing serial dilutions (at least 10-fold, to minimize drug carryover).Our study was completed by the determination of the 50% inhibitory concentration (IC50) values for ceftobiprole, benzylpenicillin, cefepime, and ceftazidime for the different PBPs of E. faecium D63r by using purified membrane preparations and fluorescent ampicillin (20, 21). Membrane preparations (300 μg of proteins) were first incubated (20 min at 37°C) with increasing concentrations of ceftobiprole and next incubated (with a saturating concentration of 25 μM fluorescent ampicillin) for an hour at 37°C. The titration of the PBPs by ceftobiprole (Fig. (Fig.22 and Table Table1)1) revealed that at a 2-fold MIC, all high-molecular-mass PBPs are inhibited by the antibiotic (Fig. (Fig.2)2) and that the low-molecular-mass protein PBP 6, which acts as a d,d-carboxypeptidase (7), is not affected. The high-molecular-mass protein PBP 2 is the most sensitive to ceftobiprole (IC50 = 0.2 μg/ml). PBPs 1 and 3 show similar IC50s (<1 μg/ml), whereas the low-affinity proteins PBP 5fm, PBP 4, and PBP 5* possess slightly higher values (≥1 μg/ml). This pattern of inhibition is completely different from that obtained with benzylpenicillin, for which the resistant protein PBP 5fm is the most insensitive PBP. The IC50 for ceftobiprole on purified soluble PBP 5fm (sPBP 5fm) gives results similar to those observed for the membrane preparations (0.7 μg/ml). Note that PBP 4 and PBP 5* exhibit very similar IC50 profiles (Table (Table1).1). A similar observation was made for PBP 4 and PBP 4* in Enterococcus hirae. PBP 4* was shown to be produced by a proteolytic cleavage of the 60 N-terminal amino acid residues of PBP 4 (11). It is very likely that, in the E. faecium membranes used here, PBP 5* was the result of an N-terminal truncation of PBP 4.Open in a separate windowFIG. 2.Relative affinities of ceftobiprole (A) and benzylpenicillin (B) for E. faecium D63r PBPs. The affinities of ceftobiprole and other β-lactams for PBPs were analyzed by a competition assay using fluorescent ampicillin. Membrane proteins were prepared from D63r and D63. Nonlabeled antibiotics were incubated with membrane proteins at 37°C for 20 min, followed by the addition of fluorescent ampicillin during 1-h incubation. The membrane fractions were subjected to SDS-PAGE and fluorography.

TABLE 1.

Inhibition of PBPs from E. faecium D63r and D63
StrainAntibioticIC50a (μg/ml) for PBP
MIC (μg/ml)
123455*6
D63rBenzylpenicillin1.6 ± 0.40.06 ± 0.010.6 ± 0.38 ± 555 ± 158 ± 32 ± 170
Cefepime1.7 ± 0.60.6 ± 0.30.5 ± 0.1>200>200>200>50>100
Ceftazidime0.9 ± 0.60.5 ± 0.30.04 ± 0.01>200>200>200>50>100
Ceftobiprole0.7 ± 0.10.2 ± 0.10.5 ± 0.11.8 ± 0.21.0 ± 0.21.4 ± 0.5>168
D63Benzylpenicillin0.9 ± 0.70.1 ± 0.071.3 ± 0.610 ± 875 ± 256 ± 41.5 ± 15
Ceftobiprole0.6 ± 0.20.2 ± 0.10.6 ± 0.21.5 ± 0.30.7 ± 0.11 ± 0.2>162
Open in a separate windowaConcentration of the β-lactam antibiotic that inhibits 50% of the fluorescent ampicillin in comparison to the level for a control containing no drug.The kinetic parameters governing the acylation of PBP 5fm by ceftobiprole were determined by using sPBP 5fm (soluble PBP 5fm from which the N-terminal membrane anchoring peptide was removed). It was overproduced and purified as previously described, except that the molecular sieve was eliminated (16). The pseudo-first-order equation was applied (5) (Fig. (Fig.3).3). The opening of the ceftobiprole β-lactam ring was measured at 319 nm with a Specord 200 spectrophotometer (Analytik Jena, Germany) at 30°C in 10 mM phosphate buffer (pH 7.0) with 10 μM sPBP 5fm. ka (the observed rate constant) was estimated by fitting with (where A0 is the initial absorbance, At is the absorbance at time t, and Af is the final absorbance). The slope allowed the determination of the value of the second-order rate constant k+2/K. A higher k+2/K value indicates faster acyl-enzyme formation, which means a faster inactivation of the PBP. The 110 ± 11 M−1 s−1 value reported for the second-order rate constant k+2/K obtained for ceftobiprole was 5 to 10 times higher than the value reported for benzylpenicillin (15 to 24 M−1 s−1), indicating that ceftobiprole inactivated sPBP 5fm faster than benzylpenicillin (21). However, this value is 70 times lower than the value obtained for S. aureus PBP 2a acylation (8,900 M−1 s−1), produced and purified in our laboratory (14). The sPBP 5fm-ceftobiprole adduct was very stable. Indeed, no free enzyme could be detected after 4 h of incubation at 37°C.Open in a separate windowFIG. 3.Variation of the pseudo-first-order rate constants (ka) of reaction of sPBP 5fm with ceftobiprole concentration. Upon reaction with β-lactam compounds, the active-site serine PBPs are immobilized in the form of a very stable acyl-enzyme. The kinetic model describing their interaction is , where E is the PBP, C the β-lactam, E·C the noncovalent complex, and EC* the acylated PBP. K is the dissociation constant of E·C, and k2 is the first-order rate constant characterizing the formation of the acyl-enzyme. The reaction obeys the equation E = E0·exp (−ka·t), in which ka is (k2/KC.In conclusion, ceftobiprole efficiently inhibited the low-affinity protein E. faecium PBP 5 in our penicillin-resistant strain. It demonstrated bactericidal activity against this laboratory-derived ampicillin-resistant E. faecium mutant that overproduced an unmutated PBP 5 protein. This profile is different from that observed for most E. faecium clinical isolates bearing a mutant PBP 5 protein that had reduced affinity, where resistance is reported for all β-lactams, including ceftobiprole, suggesting that simple overexpression of PBP 5 is sufficient to elevate the MIC for ceftobiprole but that amino acid substitutions in the protein are necessary for high-level resistance. Finally, ceftobiprole is, up to now, the best tool for easily determining kinetic parameters of unlabeled β-lactams or for finding new inhibitors by high-throughput screening using the purified sPBP 5fm protein, because of its rapid acylation of the protein and the ability to directly follow the cleavage of its β-lactam ring at 319 nm.  相似文献   

6.
7.
8.
9.
Analysis of 15 European clinical Enterobacteriaceae isolates showed that differences in the genetic context of blaCMY-2-like genes reflected the replicon type, usually IncA/C or IncI1. These blaCMY-2 loci may originate from the same ISEcp1-mediated mobilization from the Citrobacter freundii chromosome as structures described in earlier studies.Among plasmid-mediated AmpC enzymes, CMY-2 is the most prevalent both in France and worldwide, especially in Salmonella species (1, 17). Previous studies in the United States revealed that large plasmids carrying blaCMY-2 have been transmitted between different genera of bacteria, and notably from enteric Salmonella strains to pathogenic Escherichia coli strains (22). American studies of plasmids encoding CMY-2 using mixed-plasmid microarrays showed that the plasmids fell into two clusters (3); likewise, replicon-typing experiments revealed two replicon types, I1 and A/C (11). Similar studies in Taiwan confirmed that plasmids from diverse isolates carrying blaCMY-2 genes have common restriction patterns, evidence of interspecies spread (23). Further investigations of Taiwanese isolates allowed the identification of a specific transposon-like element responsible for the spread of blaCMY-2 among members of the family Enterobacteriaceae (6, 19). In the American and Taiwanese studies, ISEcp1 has been presumed to be involved both in the mobilization of blaCMY-2 from the Citrobacter freundii chromosome and in the expression of ampC lacking chromosomal ampR (6, 10, 12, 14, 15, 19). Poirel et al. have shown the involvement of ISEcp1B, an ISEcp1-like element, in the expression and mobilization of the β-lactamase gene blaCTX-M-19 (18).We characterized 15 CMY-type-producing isolates, including eight E. coli isolates, four Klebsiella pneumoniae isolates, two Proteus mirabilis isolates, and one Salmonella enterica isolate. We describe the genetic organization of blaCMY-2-like genes in these European isolates and compare our findings with those from the United States and Taiwan.Table Table11 lists the 15 clinical isolates and their sources. Six isolates were previously reported (4, 7, 9, 13, 21). Total DNA was extracted by using a QIAamp DNA Mini kit (Qiagen, Courtaboeuf, France). Using repetitive-element PCR, we showed that three of the eight E. coli clinical isolates were clonally related (profile A: TN2106, TN38148, TN386) according to the interpretation criteria of van Belkum et al. (Table (Table1)1) (16, 20). Enterobacterial repetitive intergenic consensus PCR of the four K. pneumoniae isolates gave different amplification patterns (Table (Table1)1) (16).

TABLE 1.

Clinical isolates, origins, epidemiological features, and exploration by PCR mapping
StrainReferenceCollection yrCountry of origin (city, hospital)aPCR profileCMY plasmid transferbReplicon typingCMY typePCRe
ABCDEFGHIJKL
S. enterica serovar Senftenberg SENF131994AlgeriaNDTc E. coli C1a NalrA/C2+++++++
K. pneumoniae 16971999France (Paris, H2)WcTc E. coli J53A/C2+++++++
K. pneumoniae BM297441998SwedenXcTc E. coli J53A/C4+del+del+del++++
K. pneumoniae LMCMYThis studyFrance (Colombes, H3)YcTc E. coli J53A/C2+++++++
K. pneumoniae 970171997France (Paris, H1)ZcTc E. coli J53A/C4+ins+ins+ins++++
P. mirabilis H223b211996TunisiaNDTc E. coli J53A/C4+++++++
P. mirabilis RPCMYThis study2002France (Garches, H6)NDTc E. coli J53Negative2+++++++
E. coli TN2106This study2003France (Paris, H4)AdTc E. coli J53A/C2++++++
E. coli TN38148This study2004France (Paris, H4)AdTc E. coli J53I12+++++
E. coli TN10This study2002France (Paris, H4)BdTc E. coli J53I12+++++
E. coli TN1392002France (Paris, H4)CdEp E. coli DH10BI12+++++
E. coli TN386This study2002France (Paris, H4)AdTc E. coli J53I12+++++
E. coli IGR4801This study2005France (Villejuif, H5)DdTc E. coli C600Negative2+++++
E. coli IGR4872This study2005France (Villejuif, H5)EdTc E. coli C600Negative2++++ins+ins
E. coli TN44889This study2004France (Paris, H4)FdNTND2++
Open in a separate windowaH1, Hôpital Cochin; H2, Hôpital Saint-Antoine; H3, Hôpital Louis Mourier; H4, Hôpital Tenon; H5, Institut Gustave Roussy; H6, Hôpital Raymond Poincaré.bAbbreviations: Tc, transconjugant; Ep, transformant; NT, no transfer; ND, not determined.cEnterobacterial repetitive intergenic consensus PCR profile.dRepetitive-element PCR profile.e−, negative; +, positive; +ins, positive with an insertion; +del, positive with a deletion.Except for one strain, SENF, which was previously transferred in E. coli C1a Nalr, E. coli K-12 strain J53-2 (Rifr) or C600 (Nalr) transconjugants were obtained by mating and selection on cefoxitin (10 μg/ml) and either rifampin (rifampicin) (250 μg/ml) or nalidixic acid (50 μg/ml). For isolates with which mating was unsuccessful, E. coli strain DH10B (Invitrogen SARL, Cergy-Pontoise, France) was transformed with plasmid DNA by electroporation (Bio-Rad, Marnes la Coquette, France). Transformants were selected with cefoxitin (10 μg/ml). All of the transformants or transconjugants were successfully typed by PCR-based replicon typing, with the exception of three isolates (5). Most of the isolates (11/15) carried an A/C or an I1 replicon plasmid (Table (Table11).Except for the six isolates previously published, the blaCMY gene was amplified by PCR experiments with primers ampC1 and ampC2, as previously described (7), and sequenced. Three isolates (BM2974, 9701, and H223b) carried blaCMY-4, a CMY-2 variant; all of the others carried blaCMY-2 (Table (Table11).The genetic context of blaCMY was explored both by PCR mapping and cloning experiments. A 13-kb sequence surrounding blaCMY-2 from S. enterica serovar Newport was available in GenBank (accession number DQ164214); we used several PCRs (B, C, F, G, H, I, J, and K) to characterize the genetic context of blaCMY in our collection of 15 isolates (see the supplemental material). Table Table11 shows several PCR profiles. PCR products B and F did not have the expected length for three isolates. Four isolates (169, TN44889, TN38148, and RPCMY) were chosen as representative of most of the PCR profiles, and cloning experiments were performed to study the genetic organization of blaCMY more extensively.Plasmid DNA was partially digested with Sau3AI, and the fragments were ligated into the BamHI site of pACYC184. E. coli DH10B was transformed with the resulting plasmids, and transformants were selected with cefoxitin (10 μg/ml) and chloramphenicol (50 μg/ml). The largest plasmids with a blaCMY insert were selected, and both strands were sequenced. PCR primers were designed (PCRs A, D, E, and L), and PCR experiments were performed with all of the isolates (Table (Table1;1; see the supplemental material). The sequenced regions and DNA sequence analyses are shown in Fig. Fig.11.Open in a separate windowFIG. 1.Comparison of regions surrounding blaCMY-2. IRR, imperfect right inverted repeat.Previous American and Taiwanese studies suggested that ISEcp1 is involved in mobilization of the blaCMY-2-like gene from the C. freundii chromosome, with ISEcp1-IRR consistently 117 bp upstream from blaCMY-2 (6, 10, 12). By comparing the sequence of S. enterica serovar Newport DQ164214 with that of C. freundii chromosome U21727, a 2,823-bp region including a blaCMY-2-like gene, blc, sugE, and ΔecnR could be delimited. Interestingly, the last 14 bp of this region, CCACACAATTCAGG, could have been recognized as an imperfect right inverted repeat-like sequence by the transposase of ISEcp1 during the mobilization process. In the 15 isolates in this study, at least a part of ISEcp1 was present upstream from the blaCMY-2-like gene. Nevertheless, the ISEcp1-blaCMY-2-like gene region was in the same configuration in six isolates only: 169, LMCMY, SENF, H223b, TN2106, and TN44889. For the other nine isolates, ISEcp1 was present with various deletions/insertions which do not separate the blaCMY-2-like gene from its putative promoter in ISEcp1 (Fig. (Fig.1).1). These modifications are likely to have occurred after ISEcp1 had mobilized the blaCMY-2-like gene from the C. freundii chromosome to a plasmid.The plasmids from most of the isolates shared various structural similarities with the 13-kb type I structure described in S. enterica serovar Newport (accession number DQ164214), where the blaCMY-2 gene is duplicated (12); the second copy follows a partial copy of ISEcp1 in the opposite orientation (Fig. (Fig.11).Downstream from the blaCMY-2-like gene in this 13-kb type I structure, several open reading frames (ORFs) were described between the two copies of blaCMY-2, from blc to traC (Fig. (Fig.1).1). In 8 of our 15 isolates, RPCMY, 169, BM2974, LMCMY, SENF, H223b, 9701, and TN2106, the genetic organization downstream from the blaCMY-2-like gene was very similar to this 13-kb type I structure. Only RPCMY had two copies of the blaCMY-2-like gene. For the other isolates, the sequence remains unknown either downstream from traC or downstream from orf3 (Fig. (Fig.11).Upstream from the blaCMY-2-like gene, ISEcp1 was present with deletions/insertions in its sequence for three isolates. (i) In RPCMY, there is a deletion of the first 1,140 bp due to the insertion of a separate copy of ISEcp1 in the reverse orientation; this second copy has a 1,296-bp deletion due to the insertion of IS10. Interestingly, IS10 has previously been identified as being responsible for the disruption of ISEcp1 upstream from blaCTX-M-14 in E. coli TN13, but in a different position (8). (ii) In BM2974, there is a deletion of a 104 bp-segment at the 3′ end of ISEcp1. (iii) In Kp9701, there is an insertion of IS5 at position 1509 in ISEcp1 (Fig. (Fig.11).In 169, BM274, LMCMY, SENF, H223b, TN2106, and 9701, an additional 1,560-bp segment was characterized upstream from ISEcp1. It is 100% identical to the region comprising the traB, traV, and traA genes in S. enterica serovar Newport SL254 (accession no. CP000604).Of the 15 isolates, 8 had a genetic organization around the blaCMY-2-like gene highly similar to that in the 13-kb type I structure and, except for RPCMY, carried an A/C replicon plasmid.A second group included six isolates less similar to the 13-kb type I structure (TN38148, TN10, TN13, TN386, IGR4801, and IGR4872): the blaCMY-2-like gene was followed by blc, sugE, ecnR, and orf1 only. Indeed, orf1 had 1,078 bp deleted at the 3′ end and was followed by yafA, which had a 3′ deletion and was in the reverse orientation. The truncated yafA gene was followed by yafB, and both genes belonged to a 1,983-bp segment which is 98% identical to a region of S. enterica plasmid pNF1358 (accession number no. DQ017661) (Fig. (Fig.1).1). This segment in pNF1358 is located directly upstream from ISEcp1-blaCMY-2.In all six isolates of this second group, a deletion of the first 1,281 bp in the 5′ part of ISEcp1 was due to the insertion of IS1294 in the reverse orientation (Fig. (Fig.1).1). IS1294 belongs to the IS91 family and is a putative transposable element capable of mediating one-ended transposition (www-is.biotoul.fr). Downstream from IS1294, a 2,265-bp segment comprising the repY and repZ genes involved in replication initiation is 95% identical to a region of S. enterica pNF1358 (Fig. (Fig.1).1). In IGR4872, an additional IS1294 was inserted between sugE and ecnR, leading to the duplication of the GTTC target site.Four isolates of this second group carried an IncI1 plasmid; two isolates from the same hospital could not be replicon typed.In the 15th isolate, TN44889, the genetic organization of the ISEcp1-blaCMY-2-like gene region was unique with the presence of IS4 upstream from a complete copy of ISEcp1 (Fig. (Fig.1).1). Downstream from the blaCMY-2-like gene, 467 bp of blc was deleted by the insertion of a truncated ORF, yggR. That is followed by a 4,591-bp region including five ORFs, which is 95% identical to the Shigella flexneri 2a SRL pathogenicity island. This system is linked to a cluster of multiple antibiotic resistance determinants (accession number no. AF326777).Our findings are consistent with the study of Hopkins et al. (11); there is a predominance of A/C and I1 replicons among plasmids carrying blaCMY-2-like genes in 15 nonrepetitive Enterobacteriaceae isolates. Each replicon type reflected the genetic organization surrounding the blaCMY-2-like gene, with several minor rearrangements, including small insertions or deletions. Using phylogenetic methods, Barlow and Hall demonstrated that all CMY-2-type β-lactamase variants had a common ancestor which was plasmid borne (2). This suggests that a region from the C. freundii chromosome (ampC, blc, sugE, ecnR) has been mobilized only once. This study is consistent with results of Barlow and Hall, as ISEcp1 seems to have been involved in the mobilization of the blaCMY-2-like gene, irrespective of the replicon type and the plasmid backbone.  相似文献   

10.
We have evaluated the antifungal activity of micafungin in serum by using the disk diffusion method with serum-free and serum-added micafungin standard curves. Serum samples from micafungin-treated patients have been shown to exhibit adequate antifungal activity, which was in proportion to both the applied dose and the actual concentration of micafungin measured by high-performance liquid chromatography. The antifungal activity of micafungin in serum was also confirmed with the broth microdilution method.Micafungin has been shown to bind to serum proteins at a level of 99.8% (13). If the unbound drug contributes to its pharmacological activity (the free-drug hypothesis), only 0.2% of total micafungin would be available to exert antifungal activity in the presence of serum, and the MIC for micafungin in vitro would increase 500-fold. However, several studies have shown that this ratio varies from 4- to 267-fold (6, 7, 11), indicating that the antifungal activities of micafungin in serum may not follow the free-drug hypothesis; instead, observed activities are mostly superior to those predicted. Furthermore, it remains unclear whether these results can be applied to micafungin in a patient''s serum. To address this issue, we collected serum samples from micafungin-treated patients and examined the relationship between micafungin concentration and its in vitro antifungal activity in serum.This study was approved by the institutional review board, and informed consent was obtained from each patient. Patients with hematologic malignancies, admitted into Osaka University Medical Hospital, were administered micafungin at a dose of 50 to 300 mg/body once daily. The efficacy of prophylaxis was defined as the absence of proven, probable (EORTC-IFICG/NIAID-MSG) (1), or suspected (unexplained persistent fever and clinical findings) (10) fungal infection, through the end of therapy. The efficacy of the drug for suspected fungal infections was indicated by improvement of persistent fever and clinical findings.Blood samples were collected from patients just before (trough) and after (peak) micafungin infusion, at least 4 days after initiating treatment (steady state) (2). Micafungin concentration in serum was measured by high-performance liquid chromatography (HPLC) (9, 12). The disk diffusion method was performed according to National Committee for Clinical Laboratory Standards (NCCLS) M44-A guidelines (5). To obtain standard curves, we prepared two types of serial dilution disks impregnated with micafungin standard solution, one in RPMI 1640 (serum-free standard) and the other in heat-inactivated serum from volunteers (serum-added standard). Disks were applied to Sabouraud dextrose agar plates inoculated with Candida albicans FP633, a clinical isolate kindly provided by Astellas Pharma Inc., Tokyo, Japan. The diameter of the area of complete growth inhibition (inhibitory zone) was measured. Similarly, disks were impregnated with serum samples collected from patients, and the inhibitory zones were measured. The determination of antifungal activity of micafungin in a patient''s serum was based on two standard curves, as described above. To determine the inhibitory titer in a patient''s serum, we utilized the broth microdilution method based on the guidelines in NCCLS M27-A2 (4). Serum from a patient was serially diluted twofold with serum from a volunteer, supplemented with 20 mM HEPES, and inoculated with C. albicans FP633. MIC was defined as the lowest concentration where no visible growth was observed. Serum inhibitory titers were defined as the highest dilution of serum that completely inhibited fungal growth.In all seven patients, micafungin was effective for prophylaxis or treatment against fungal infections (Table (Table1).1). Serum peak concentrations (Cmax) of micafungin (measured by HPLC) ranged from 5.59 to 37.1 μg/ml at a dose of 50 to 300 mg/body and closely correlated with both daily dose and dosage in terms of body weight (Table (Table2).2). Standard curves were prepared from both serum-free and serum-added micafungin standard disks (Fig. (Fig.1).1). The antifungal activity of micafungin remained intact in serum: 20 to 50% (by measured value) or 25 to 30% (by standard curve).Open in a separate windowFIG. 1.Estimation of micafungin concentration in serum samples from patient no. 5, using the disk diffusion method. (A) Concentration measured using HPLC, 16.4 μg/ml. (B) Concentration estimated from the serum-free micafungin standard curve, 6.0 μg/ml. (C) Concentration estimated from the serum-added micafungin standard curve, 22.1 μg/ml. Ratio of concentration B to concentration A (%) = 6.0/16.4 × 100 = 37. Ratio of concentration C to concentration A (%) = 22.1/16.4 = 134.8.

TABLE 1.

Patient background
Patient no.Age (yr)GenderaBWb (kg)DiagnosiscHSCTdAntifungal treatmentDose of micafungin (mg/body)Duration of therapy (days)Clinical efficacy
143M76MLAuto-PBSCTPreemptive therapy30011Effective
259F52MLAuto-PBSCTPreemptive therapy3008Effective
333M52MSAllo-BMTEmpirical therapy7555Effective
451F47MLAllo-BMTEmpirical therapy509
15016Effective
22520
1504
547F58MLAuto-PBSCTEmpirical therapy15021Effective
3007
622F45AMLAllo-BMTProphylaxis5022Effective
1006
746F43ALLAllo-BMTProphylaxis5022Effective
1009
Open in a separate windowaM, male; F, female.bBW, body weight.cML, malignant lymphoma; MS, myelodysplastic syndrome; AML, acute myeloid leukemia; ALL, acute lymphoblastic leukemia.dHSCT, hematopoietic stem cell transplantation; PBSCT, peripheral blood stem cell transplantation; BMT, bone marrow transplantation.

TABLE 2.

Antifungal activities and inhibitory titers of serum samples from patients administered micafungin
Patient no.Dose of micafungin
Collection point
Antifungal activity of serum samples (μg/ml) measured using:
Ratio (%) of antifungal activities measured by:
Serum inhibitory titer
mgmg/kgDayTimeHPLCDisk diffusion methodc
Serum-free standard curve/HPLCaSerum-added standard curve/HPLCb
Serum-free standard curveSerum-added standard curve
13003.910Peak34.2NDND32
23005.88Peak33.6NDND32
3751.445Peak6.72.54.238634
42254.843Peak37.114.134.8389432
51502.612Peak16.46.022.13713516
6501.18Trough2.71.12.142782
8Peak8.43.88.142968
15Trough3.01.01.732584
15Peak5.62.55.245928
7501.215Trough2.30.91.640712
15Peak6.43.37.0521094
17Trough2.0NDND2
17Peak6.52.95.944918
Open in a separate windowaMean ± standard deviation is 41% ± 6%.bMean ± standard deviation is 89% ± 23%.cThese serum concentrations were estimated using the two standard curves. ND, not determined.Results for all seven successfully treated patients are summarized in Table Table2,2, as are the micafungin concentrations in serum samples measured by HPLC. The antifungal activity of micafungin in serum samples from these patients was 41% ± 6% (mean value ± standard deviation, ranging from 37% to 52%) of the actual micafungin serum concentration (the ratio of antifungal activity estimated by the disk diffusion method based on the serum-free standard curve to that measured by HPLC). Representative results for patient no. 5 are shown in Fig. Fig.1.1. Meanwhile, the antifungal activity of micafungin calculated with the serum-added standard curve was almost equal to the actual micafungin serum concentration (the ratio of antifungal activity estimated by the disk diffusion method based on the serum-free standard curve to that measured by HPLC was 89% ± 23% [mean ± standard deviation, ranging from 58% to 135%]) (Table (Table22).MIC for micafungin against C. albicans FP633 in heat-inactivated serum from a volunteer was 1 μg/ml, which was consistent with previously reported data using the same strain (3). At this concentration, micafungin induced swelling and subsequent burst of mycelia. Inhibitory titers for serum samples from all patients are summarized in Table Table2.2. Representative results from patient no. 5 are shown in Fig. Fig.2.2. These titers were in excellent agreement with both micafungin concentrations in serum samples by HPLC and those estimated from the serum-added standard curve (Table (Table22).Open in a separate windowFIG. 2.Determining the inhibitory titer values for serum from patient no. 5 using the broth microdilution method. MIC was defined as the lowest concentration at which no visible growth was observed (magnification of ×40). Serum inhibitory titers were defined as the highest dilution of serum that completely inhibited fungal growth. Insets show C. albicans morphologies (magnification of ×400).These results indicate that serum proteins certainly bind to micafungin and reduce its antifungal activity, but this binding may be reversible and weak. These data are inconsistent with the free-drug hypothesis. One or more of the following reasons could explain this discrepancy. First, micafungin binds to serum proteins at 99.8% in situations without any other competitors, such as in ultrafiltration, the method measuring the equilibrium binding (13). If fungi susceptible to micafungin are present, however, micafungin may be easily released from the protein-bound form in a rapid equilibrium, bind to target pathogens, and exert its antifungal activity. In this case, increased MIC of micafungin in serum may depend on the fungal strains being tested (6, 7). Furthermore, although albumin is supposed to bind mainly to micafungin, several other proteins in serum, such as alpha and gamma globulins, might influence the interactions among micafungin, serum proteins, and target pathogens (8).In conclusion, it seems to be unsuitable to apply the free-drug hypothesis to the pharmacodynamics of micafungin, because this may underestimate its antifungal activity. We have shown, using the disk diffusion and broth dilution methods, that serum samples from micafungin-treated patients exhibited adequate antifungal activity. Our data will be useful for understanding the pharmacodynamics of micafungin and for improving the clinical outcome of micafungin treatment.  相似文献   

11.
The incidence of naturally occurring AmpC β-lactamases with extended activities toward several cephalosporins was evaluated among 17 ceftazidime (CAZ)-resistant Acinetobacter baumannii isolates. Five AmpC β-lactamases (named ADC β-lactamases) were identified, among which those possessing the Val208Ala (inside the omega-loop) or Asn283Ser (helix H-10) substitution conferred higher levels of resistance (4- to 64-fold higher) to CAZ and to cefotaxime in Escherichia coli. This study demonstrates that peculiar AmpCs playing a role in resistance to broad-spectrum cephalosporins in A. baumannii may be identified.Acinetobacter baumannii is commonly associated with serious nosocomial infections (4, 10, 14). A growing number of β-lactamases conferring resistance to broad-spectrum cephalosporins in Acinetobacter spp. have been identified. Even though resistance to oxyiminocephalosporins (ceftazidime and cefotaxime) may be related to production of extended-spectrum β-lactamases (ESBLs), it is usually associated with overproduction of an AmpC-type β-lactamase (1, 2, 9). Overexpression of the blaAmpC gene of A. baumannii may occur as a result of an insertion sequence, ISAba1, providing strong promoter sequences (5, 8, 13), being located upstream.Most AmpC-type β-lactamases naturally produced by Gram-negative bacteria hydrolyze amino- and ureidopenicillins, cephamycins (cefoxitin and cefotetan), and, at a low level, oxyiminocephalosporins, such as ceftazidime, cefotaxime, and ceftriaxone, and monobactams, such as aztreonam (3). AmpCs possessing a broad substrate activity have been reported in Enterobacteriaceae and Pseudomonas aeruginosa (11, 12, 16, 19, 20). These AmpCs (or extended-spectrum AmpCs [ESACs]) with activities against several cephalosporins may confer reduced susceptibility to those molecules (6, 11, 12, 19). They differ from “regular” cephalosporinases by amino acid substitutions or insertions/deletions in four specific regions that are all located in the vicinity of the active site: the Ω loop, the H-10 helix, the H-2 helix, and the C-terminal extremity of the protein (15).Recently, we have identified the first ESAC (named ADC-33, according to the numbering of AmpC β-lactamases from A. baumannii [9]) from a single A. baumannii isolate (21). ADC-33, possessing a Pro210Arg substitution together with a duplication of an Ala residue at position 215 (inside the Ω loop), hydrolyzed ceftazidime, cefepime, and aztreonam at high levels (21).The present study aimed to evaluate the extent of the variability of ADC β-lactamases in A. baumannii. Seventeen nonrepetitive A. baumannii clinical isolates that were recovered in 2007-2008 from patients hospitalized at the Bicêtre Hospital and were resistant to ceftazidime were included in this study. Genotyping was performed by using pulsed-field gel electrophoresis (PFGE) according to the instructions of the manufacturer (Bio-Rad, Marnes-la Coquette, France) as previously described (17, 22), and it identified eight distinct clones among the 17 isolates (Table (Table11).

TABLE 1.

Features of the A. baumannii isolates studiedd
Clone (PFGE)AmpC overproducerbISAba1-blaADC colinearityADC variantOXA-51-like variantMIC (μg/ml)c
CAZCTXFEPCPOATMIPM
A (strain CIP)aADC-50OXA-64281280.25
B (strain AYE)++ADC-11OXA-69>256>256>256>256>2561
C++ADC-51OXA-69>256>25664>256>256>32
D++ADC-30OXA-66128>25632256324
EADC-52OXA-90326432641280.25
F++ADC-30OXA-6664>256321281282
G++ADC-51OXA-69>256>25632>256>256>32
H++ADC-30OXA-6664>25632641282
I++ADC-26OXA-6932>2561664322
J++ADC-53OXA-5164>25683232>32
Open in a separate windowaClones A and B correspond to A. baumannii CIP7010 and A. baumannii AYE, coding for regular AmpC β-lactamases.bAmpC overproduction evaluated by cloxacillin test (250 μg/ml).cCAZ, ceftazidime; CTX, cefotaxime; FEP, cefepime; CPO, cefpirome; ATM, aztreonam; IPM, imipenem.dThere is no ISAba1-blaOXA-51-like colinearity among any of the clones in the table.In addition to being resistant to ceftazidime, most of the clones were resistant to cefotaxime, ceftriaxone, aztreonam, cefepime, and cefpirome (Table (Table1).1). By use of cloxacillin-containing plates as described previously (17), the susceptibility to ceftazidime was recovered for most of the clones (excluding clones C, E, and G), indicating that overproduction of the ADC β-lactamase was responsible for the resistance to ceftazidime. Double-disk synergy tests using clavulanic acid-containing disks were performed as described previously (17) for those isolates belonging to clones C, E, and G, but no ESBL production was detected.Insertion sequence ISAba1, providing strong promoter sequences (−35 [TTAGAA] and −10 [TTATTT]) immediately upstream of the blaAmpC gene, was identified in almost all clones, similar to previous observations (8).Analysis of the naturally occurring carbapenem-hydrolyzing blaOXA-51-like sequences was also performed as described previously (Table (Table1)1) (7, 18). No colinearity between ISAba1 and the blaOXA-51-like genes was observed in any of these isolates.The blaAmpC genes of the A. baumannii isolates were sequenced and cloned with or without the PISAba1 promoter as described previously (21). The inserts containing the PISAba1 promoter were further named with “P+.” The A. baumannii AYE clinical isolate and A. baumannii CIP7010 reference strain were used to clone wild-type blaADC genes expressing ADC-11 and ADC-50, respectively. Five ADC variants (ADC-26, ADC-30, ADC-51, ADC-52, and ADC-53) were identified (Fig. (Fig.1),1), with ADC-51, ADC-52, and ADC-53 corresponding to newly identified enzymes. Amino acid changes were not identified either in the SVSK, YSN, and KTG conserved motifs or in helix H-2, previously associated with the extended activity of ESACs identified in other Gram-negative species (11, 15). However, substitutions Val208Ala located in the omega loop of ADC-53 and Asn283Ser located in helix H-10 of ADC-51 were identified.Open in a separate windowFIG. 1.Amino acid sequence alignment, including ADC-7 taken as a reference sequence as previously published (9) and ADC-11 and ADC-50 as wild-type ADCs, together with the newly characterized ADC-26, ADC-30, ADC-51, ADC-52, and ADC-53. Identical amino acids are indicated by dashes. The typical ADC β-lactamase domains (SVSK, YSN, and KTG) are underlined. Helices H-2 and H-10 are boxed in gray. The Ω loop is boxed in gray and double underlined. Differences observed inside the Ω loop and helix H-10 are in bold. The vertical arrow indicates the position of the +1 amino acid (cleavage site for signal peptide). Numbering is according to the sequence of the mature protein. Asterisks indicate ADC β-lactamases possessing ESAC properties.Subsequent cloning showed that MICs of broad-spectrum cephalosporins differed significantly among the recombinant Escherichia coli strains that did not include the strong PISAba1 promoter upstream of the blaADC genes (Table (Table2).2). Higher MICs of ceftazidime were observed for E. coli TOP10(pADC-30), E. coli TOP10(pADC-51), and E. coli TOP10(pADC-53), producing β-lactamases ADC-30, ADC-51, and ADC-53, respectively.

TABLE 2.

MICs of β-lactams for E. coli TOP10 strains harboring recombinant plasmids and a TOP10 reference strain
β-Lactam(s)aMIC(μg/ml) for E. coli TOP10 with:
pADC-26pADC-30pADC-51pADC-52pADC-53pADC-11No plasmid
Amoxicillin>256>25632256>256>2562
Amoxicillin + CLA>256>25632256>2562562
Ticarcillin8825681682
Ticarcillin + CLA8825681682
Piperacillin88481681
Piperacillin + TZB88481681
Cefuroxime64>2566464321282
Ceftazidime0.52320.540.50.125
Cefotaxime0.25140.520.50.06
Cefepime0.030.030.060.030.030.0320.03
Cefpirome0.030.030.060.030.030.0320.03
Aztreonam0.1250.250.250.1250.50.0940.03
Imipenem0.250.250.250.250.50.190.125
Meropenem0.0230.0230.0470.0230.0230.0230.03
Open in a separate windowaSusceptibility testing performed by using Etest strips (AB bioMérieux, Solna, Sweden). CLA, clavulanic acid (4 μg/ml); TZB, tazobactam (4 μg/ml).MICs of ceftazidime, cefotaxime, and aztreonam were higher for the clones expressing ADC-30, ADC-51, and ADC-53, suggesting that they might correspond to ESACs (Table (Table33).

TABLE 3.

MICs of β-lactams for E. coli TOP10 strains harboring recombinant plasmids and a TOP10 reference strain
β-Lactam(s)aMIC(μg/ml) for E. coli TOP10 with:
pADC-26-P+pADC-30-P+pADC-51-P+pADC-53-P+pADC-11-P+No plasmid
Amoxicillin>256>256>256>256>2562
Amoxicillin + CLA>256>256256>256>2562
Ticarcillin128128>256>2562562
Ticarcillin + CLA128128>256>2562562
Piperacillin>256256>256>256>2561
Piperacillin + TZB2561281282562561
Cefuroxime>256>256>256>256>2562
Ceftazidime864>25625680.125
Cefotaxime864>25612880.06
Cefepime0.1250.250.50.250.1250.03
Cefpirome0.1250.2520.50.1250.03
Aztreonam2881620.06
Imipenem0.250.50.250.250.250.125
Meropenem0.030.030.030.030.030.03
Open in a separate windowaSusceptibility testing performed by using Etest strips (AB bioMérieux, Solna, Sweden). CLA, clavulanic acid (4 μg/ml); TZB, tazobactam (4 μg/ml).In order to better evaluate the impact of the ADC variants with extended activities (ADC-30, ADC-51, and ADC-53), recombinant plasmids containing the ISAba1-blaADC fragments were electrotransformed into E. coli HB4 with permeability defects associated with a lack of OmpC and OmpF porins (12). Besides that of ceftazidime, the MICs of cefepime, cefpirome, and aztreonam were higher for the ADC-51 producer (16, 64, and 32 μg/ml, respectively) than for the wild-type ADC-26 producer (2, 4, and 8 μg/ml, respectively).Several amino acid substitutions were identified here among ADC β-lactamases (Fig. (Fig.1).1). These residues were located at positions 188, 242, 255, 274, 275, and 313 (Fig. (Fig.1),1), but none of those was located in domains supposed to be crucial for β-lactamase activity, such as the Ω loop, helix H-2, or helix-H10. A steric view of these substitutions in the overall fold structure of the modeled wild-type ADC-26 variant seems to indicate that these amino acid substitutions are located far from the active site of the enzyme (data not shown). In an analysis of sequences of ADC variants possessing ESAC properties, compared to those of wild-type β-lactamases ADC-11 and ADC-26, it is interesting that both ADC-51 and ADC-53 variants possess a Val208Ala substitution located in the omega loop and an Asn283Ser substitution located in helix H-10, respectively. These substitutions may explain their extended-spectrum activity. However, ADC-30 possesses two specific residues relative to the other ADC sequences (Lys122 and Ser139) and a Thr at position 313 that was identified in the ADC-53 ESAC. Whether those unusual amino acid substitutions (alone or combined) might play a significant role in the broadened activity of ADC-30 remains to be determined.We report here the identification and characterization of five blaADC genes from ceftazidime-resistant A. baumannii isolates. Three of them encode novel ADC variants and three of them possess significant extended activity toward ceftazidime and cefotaxime. After the identification of ESACs in Enterobacteriaceae, the demonstration of their wide diffusion among P. aeruginosa, and the recent and initial finding of such an enzyme in A. baumannii, further investigations are required to evaluate whether ADC β-lactamases with extended activities might be widely distributed in A. baumannii.  相似文献   

12.
We evaluated the in vitro activities of anidulafungin, micafungin, and caspofungin against Candida krusei by determining MIC and minimum fungicidal concentration (MFC) measurements and by the time-kill method. The geometric mean (GM)-MIC/GM-MFC values were 0.1/0.34, 0.25/0.44, and 1/2.29, respectively. The mean times to reach 99.9% growth reduction were 19.1 ± 18.2 h (mean ± standard deviation) for 2 mg/liter anidulafungin, 37.4 ± 8.8 h for 2 mg/liter caspofungin, and 30.7 ± 12.2 h for 1 mg/liter micafungin. Anidulafungin exhibited the highest time-kill rate, followed by micafungin. The three echinocandins showed fungicidal activity at concentrations reached in serum.Candida krusei is isolated mainly from immunocompromised patients and those who have received fluconazole treatment. The isolation (2 to 60%) and mortality (30 to 60%) rates depend on the institution and on the unit to which the patient is admitted (10, 13, 14, 17, 20). Treatment of invasive C. krusei infections can be difficult due to its intrinsic resistance to fluconazole and its reduced susceptibility to amphotericin B and flucytosine (2, 15). In immunocompromised patients, fungicidal agents may be required to eradicate infection since the major host defense (phagocytic killing of neutrophils and monocytes/macrophages) is reduced (9).Anidulafungin, caspofungin, and micafungin are novel antifungals that share the same spectrum of activity but differ in their MICs. Although the fungicidal activities and killing rates of echinocandins have been previously evaluated (3, 7, 8, 16), none of these studies performed a head-to-head comparison. However, these agents also may differ in their fungicidal activities. This study aimed to determine their relative antifungal activities against C. krusei by three measures: (i) MICs, (ii) minimal fungicidal concentrations (MFCs), and (iii) time-kill curves.Anidulafungin (Pfizer España, Madrid, Spain), caspofungin (Merck Sharpe & Dome, Madrid, Spain), and micafungin (Fugisawa Pharmaceutical Company, Japan) were provided by their manufacturers. Caspofungin and micafungin were dissolved in water and anidulafungin in dimethyl sulfoxide; further dilutions were prepared in standard RPMI 1640 medium (Sigma-Aldrich, Madrid, Spain). Final anidulafungin and micafungin drug concentrations ranged from 0.03 to 16 mg/liter, and that of caspofungin from 0.12 to 64 mg/liter.MIC and MFC values were obtained for 20 C. krusei bloodstream isolates, the caspofungin-resistant strain CY-118 (MIC > 8 mg/liter) (11), and C. krusei ATCC 6258. MICs were determined by following the CLSI M27-A3 microdilution method (4). MIC2 and MIC0 (minimum concentrations that produce ≥50 and 100% growth reduction, respectively) were determined at 24 and 48 h. As previously described but using a larger inoculum size (1), MFCs were obtained by plating 0.1 ml from clear MIC wells following 48 h of incubation onto Sabouraud dextrose agar plates (SDA) (1). The MFC was the lowest drug concentration that resulted in ≤1 colony (≈99% killing). C. krusei ATCC 6258 was included in each batch of experiments (5).Time-kill studies were performed for six blood isolates (randomly selected from the 20 blood isolates), CY-118, and ATCC 6258. The antifungal carryover effect and the time-kill curve for each agent were determined as previously described (RPMI 1640 medium, 1 × 106 to 5 × 106-CFU/ml inoculum, and 5-ml volume) (2). The anidulafungin and caspofungin drug concentrations evaluated were 0.03, 0.12, 0.5, 2, and 8 mg/liter and, in addition, 32 mg/liter for caspofungin. The micafungin concentrations assessed were 0.06, 0.25, 1, 8, and 16 mg/liter. These concentrations are within the range achieved clinically (6, 18, 19). At 0, 2, 4, 6, 12, 24, and 48 h, aliquots of 0.1 ml were removed to determine the number of CFU/ml. The lowest limit of accurate and reproducible detectable colony counts was 100. All experiments were performed twice with three replicates for every dilution of each time point.To our knowledge, this is the first study comparing the killing activities of anidulafungin, caspofungin, and micafungin head to head with the same strains.The geometric mean (GM)-MIC2, GM-MIC0, and GM-MFC and the concentrations that inhibited and killed 90% of isolates (MIC90 and MFC90, respectively) were calculated. Time-kill data were fitted to the exponential equation Nt = N0 × ekt, where Nt is the number of viable cells at time t, N0 is the starting inoculum, k is the kill rate, and t is the incubation time. The goodness of fit for each isolate/drug was assessed by the R2 value. The times (hours) to achieve 50, 90, 99, and 99.9% reductions in growth compared to the starting inoculum size were calculated from the kill rate value as described elsewhere (2).The MICs for C. krusei ATCC 6258 were within the established ranges (5); the anidulafungin, micafungin, and caspofungin MFCs for this strain were 0.25, 0.5, and 2 mg/liter, respectively. Table Table11 summarizes the MIC and MFC determinations for the 20 blood isolates. Anidulafungin was the most active, followed by micafungin and caspofungin. The MFC/MIC0 ratios were ≤2 for 90, 90, and 100% of isolates, respectively. The anidulafungin and micafungin MFCs were 1 to 7 dilutions lower than those of caspofungin, while the anidulafungin MFCs were only 1 to 2 dilutions lower than those of micafungin.

TABLE 1.

In vitro activitiesa of echinocandin against 20 C. krusei bloodstream isolates
DrugTime (h)MIC2
MIC0
MFC
RangeMIC50MIC90GMRangeMIC50MIC90GMRangeMFC50MFC90GM
Anidulafungin240.03-0.250.120.120.10.03-0.250.120.250.15
480.03-0.50.120.250.140.03-0.50.250.50.290.12-0.50.250.50.34
Micafungin240.12-0.50.250.50.250.25-0.50.50.50.38
480.12-0.50.50.50.280.25-0.50.50.50.410.25-0.50.50.50.44
Caspofungin240.5-21210.5-2121.03
480.5-2121.270.8-8121.681-64282.29
Open in a separate windowaValues are in mg/liter.The time-kill curve data for the six isolates were averaged, and the means and standard deviations are depicted in Fig. Fig.1A.1A. Table Table22 shows the MICs and MFCs for these isolates, and Fig. Fig.2A2A shows the relationship between killing rate and concentration. During the first 6 h, the killing activities of the three drugs were isolate or strain dependent but concentration independent (the killing rate did not increase substantially with increasing concentration). However, after 6 h, little killing or regrowth was observed depending on the concentration. In general, killing activity began at the MIC, but the ≥3 log decrease (killing/fungicidal endpoint) was achieved with ≥0.5 mg/liter at 19.1 to 20.7 h with anidulafungin; 21.4 to 37.4 h with caspofungin; and 25.9 to 31.8 h with micafungin (Table (Table33).Open in a separate windowFIG. 1.Time-kill curves of anidulafungin (AND), caspofungin (CAS), and micafungin (MCF) for the six blood isolates (A), strain CY-118 (B), and strain ATCC 6258 (C). Data plotted are the means and standard deviations from two separate experiments for each growth curve, including the control. Dotted line, 3-log decrease; CT, control without antifungal agent.Open in a separate windowFIG. 2.Means and standard deviations of the results for kill rate versus drug concentration obtained for six C. krusei blood isolates (A), strain CY-118 (B), and strain ATCC 6258 (C). K values form the slope of the regression line obtained from the kill curve. Positive and negative K values correlate with increases and decreases, respectively, in viable cell numbers. Continuous line, anidulafungin (AND); discontinuous line, caspofungin (CAS); dotted line, micafungin (MFC); L, liters.

TABLE 2.

Echinocandin MIC and MFC valuesa for the eight isolates evaluated by the time-kill curve method
Isolate or strainAnidulafungin
Caspofungin
Micafungin
MIC2MFCMIC2MFCMIC2MFC
CK-10.120.5180.250.5
CK-20.060.5110.50.5
CK-60.120.25110.50.5
CK-180.120.5120.250.5
EU-1230.120.5220.120.5
EU-2200.120.12220.250.25
CY-1180.584>64116
ATCC-62580.03-0.120.25-0.5120.12-0.250.25-0.5
Open in a separate windowaValues are in mg/liter. MIC2 values are those measured at 24 h.

TABLE 3.

Time to achieve 50, 90, 99, and 99.9% growth reduction from the starting inoculum size
AgentAmt of growth reduction (%)Mean time ± SD (h) at indicated concn (mg/liter)
0.120.250.512481632
Anidulafungin507.7 ± >486.7 ± 6.11.9 ± 1.82.1 ± 1.6
9025.5 ± >4822.4 ± 20.26.4 ± 6.16.9 ± 5.5
99>48>4812.7 ± 12.213.8 ± 10.9
99.9>48>4819.1 ± 18.220.68 ± 16.4
Caspofungin50No killing8.3 ± 5.73.7 ± 0.82.9 ± 0.62.1 ± 0.6
90No killing27.6 ± 18.912.5 ± 2.99.7 ± 2.07.1 ± 1.9
99No killing>4824.9 ± 5.819.5 ± 4.014.3 ± 3.9
99.9No killing>4837.4 ± 8.8429.2 ± 6.021.4 ± 5.9
Micafungin503.1 ± 10.13.1 ± 1.23.2 ± 1.12.6 ± 1.1
9010.3 ± 33.610.2 ± 4.110.63 ± 3.78.64 ± 3.7
9920.6 ± >4820.5 ± 8.121.25 ± 7.517.28 ± 7.5
99.931.0 ± >4830.7 ± 12.231.88 ± 11.325.91 ± 11.2
Open in a separate windowAgainst the caspofungin-resistant strain CY-118, anidulafungin and micafungin exhibited killing activities, reaching the fungicidal endpoint with 8 mg/liter of anidulafungin at 24 h and 16 mg/liter of micafungin at 48 h (Fig. (Fig.1B).1B). Against C. krusei ATCC 6258, the killing kinetics were similar to those exhibited for the blood isolates (Fig. (Fig.1C)1C) and the results reported by Ernst et al. (7).The relationship between the kill rate and concentration for each echinocandin was lineal (Fig. (Fig.2).2). Anidulafungin showed the highest kill rate against the isolates and strains tested. In contrast, the micafungin kill rate was slightly lower but parallel to that of anidulafungin against strains CY-118 and ATCC 6258. Our results are similar to those reported by other authors, although they only evaluated two C. krusei isolates (7, 8, 16). Recently, our results and the utility of killing studies have been confirmed for a murine model of C. krusei infection treated with anidulafungin (12).In summary, the three echinocandins have demonstrated fungicidal activity in both MFC and time-kill studies at total drug concentrations that are reached in serum (≥1 mg/liter) throughout the dosing interval, which is an advantage when treating infections caused by C. krusei mainly in immunocompromised patients.  相似文献   

13.
14.
15.
Acinetobacter lwoffii, a species whose natural habitat is the human skin, intrinsically possesses a chromosomal gene encoding a carbapenem-hydrolyzing class D β-lactamase, OXA-134. This species may therefore constitute a reservoir for carbapenemase genes that may spread among other Acinetobacter species.Acinetobacter baumannii, the most common Acinetobacter species isolated from humans, is an opportunistic pathogen for which resistance to carbapenems is increasing worldwide (13-15). Carbapenem resistance in A. baumannii is associated mostly with acquired carbapenem-hydrolyzing class D β-lactamases (CHDLs) (19). Four groups of acquired CHDLs in A. baumannii, OXA-23, OXA-40, OXA-58, and OXA-143, have been identified (9, 18). In addition, A. baumannii possesses a naturally occurring blaOXA-51 or blaOXA-69 CHDL-encoding gene that reduces the efficacy of carbapenems when it is overexpressed (4, 5, 7, 22). Identification of the sources of acquired and clinically relevant CHDLs is important to better understand the way and the reason why these resistance determinants are spreading. Acinetobacter radioresistens has recently been identified as the natural carrier of blaOXA-23, a gene encoding one of the most commonly acquired CHDLs in A. baumannii (16). However, the progenitors of the other acquired CHDLs identified in Acinetobacter species remain unknown. Our study aimed to evaluate whether other Acinetobacter species may represent additional reservoirs of CHDL-encoding genes.The screening panel included strains belonging to 23 Acinetobacter species, including A. junii, A. johnsonii, A. haemolyticus, A. baylyi, A. lwoffii, A. radioresistens, A. schindleri, A. ursingii, A. calcoaceticus, A. gerneri, A. tjernbergiae, A. bouvetii, A. tandoii, A. grimontii, A. towneri, A. parvus, and Acinetobacter genomospecies 3, 6, 9, 10, 13, 15, 16, and 17. Acinetobacter genomospecies 9 is now classified as A. lwoffii (15). Strains were identified at the species level by using 16S rRNA sequencing (3). Susceptibility testing was analyzed by the disk diffusion method in accordance with the guidelines of the Clinical and Laboratory Standards Institute (1), and MICs were determined by using Etest strips (AB bioMérieux, Solna, Sweden).Screening for the known CHDL-encoding blaOXA-23, blaOXA-40, blaOXA-58, and blaOXA-143 genes was performed by PCR using internal primers (8, 9). This screening was positive only for the blaOXA-23 gene and only for A. radioresistens strain 3 and A. lwoffii strain 1. After sequencing, A. radioresistens strain 3 was found to possess the blaOXA-23 gene, in accordance with previous results (16). Sequencing of the amplicon obtained from A. lwoffii strain 1 identified a gene encoding a novel OXA-type β-lactamase. Thermal asymmetric interlaced (TAIL) PCR experiments were performed in order to obtain the entire sequence of this gene (11, 12). It encoded a 273-amino-acid protein named OXA-134 that shared 63, 58, 57, and 53% amino acid identity with OXA-23, OXA-40, OXA-51, and OXA-58, respectively. OXA-134 possessed the typical features of a class D β-lactamase, including the STFK tetrad at positions 70 to 73 according to class D β-lactamase (DBL) numbering (Fig. (Fig.1)1) (2). Also, as observed for other CHDLs (except for OXA-58), an FGN motif at DBL positions 144 to 146 replaced the usual YGN motif of classical class D β-lactamases (18). Finally, a KSG element was identified at DBL positions 216 to 218, as observed in the CHDLs OXA-40 and OXA-51, whereas a KTG motif is present in most class D β-lactamases, including the CHDLs OXA-23 and OXA-58 (18). A phylogenetic analysis showed that OXA-134-like β-lactamases were constituting a separate subgroup of CHDLs but that this subgroup was more closely related to the identified class D β-lactamases from Acinetobacter spp. than to other known CHDLs (Fig. (Fig.2)2) .Open in a separate windowFIG. 1.Amino acid alignment of the seven OXA-134-like class D β-lactamases identified in this study. Dashes indicate amino acids identical to those in the OXA-187 sequence. Amino acid motifs which are well conserved (even if possibly variable) among class D β-lactamases are shaded in gray. Numbering is according to DBL numbering (2).Open in a separate windowFIG. 2.Dendrogram obtained for 32 class D β-lactamases by neighbor-joining analysis. The alignment used for tree calculation was performed with ClustalX. Branch lengths are drawn to scale and are proportional to the number of amino acid changes. The distance along the vertical axis has no significance. The different clusters identified allowed the identification of nine main groups, considering that proteins from the same group have more than 80% amino acid identity. The class D β-lactamases which are considered to be naturally occurring are indicated together with the names of the corresponding species. R. pickettii, Ralstonia pickettii; B. pseudomallei, Burkholderia pseudomallei; A. xylosoxidans, Alcaligenes xylosoxidans; A. jandaei, Aeromonas jandaei; L. gormanii, Legionella gormanii; P. aeruginosa, Pseudomonas aeruginosa; P. pnomenusa, Pandoraea pnomenusa; C. jejuni, Campylobacter jejuni; B. pilosicoli, Brachyspira pilosicoli; S. oneidensis, Shewanella oneidensis; S. algae, Shewanella algae.A. lwoffii is a commensal organism found on human skin, the perineum, and the oropharynx. It has been associated with catheter-related bloodstream infections in immunocompromised patients and with bacteremia associated with community-acquired gastroenteritis and gastritis (20, 21). All the Acinetobacter genomospecies 9/A. lwoffii isolates we included in our study were fully susceptible to all antibiotics tested, including penicillins, imipenem, and meropenem. It is therefore likely that the blaOXA-134-like genes were not expressed (or were expressed at an insignificant level) in these hosts.In order to study the biochemical properties of OXA-134, cloning of the blaOXA-134 gene into the kanamycin-resistant plasmid pCR-BluntII-TOPO (Invitrogen, Life Technologies, Cergy-Pontoise, France) was performed using PCR products generated with primers PreOXA-134A (5′-GAAAAATGACCAAAATTTGATCG-3′) and PreOXA-134B (5′-TATTTGCATCATCCTTCAGC-3′) as described previously (16). Escherichia coli TOP10(pOXA-134) showed reduced susceptibility to imipenem and meropenem and resistance to most penicillins that was not inhibited by β-lactamase inhibitors (Table (Table11).

TABLE 1.

MICs of β-lactams for the different A. lwoffii isolates, E. coli TOP10 harboring recombinant plasmid pOXA-134, and the E. coli TOP10 reference strain
β-Lactam(s)aMIC (μg/ml) for:
A. lwoffii isolatesE. coli TOP10(pOXA-134)E. coli TOP10
Amoxicillin0.5-1>5124
Amoxicillin + CLA0.5-11284
Ticarcillin0.5-1>5124
Ticarcillin + CLA0.5-12564
Piperacillin1.5-481
Piperacillin + TZB1.5-481
Cephalothin4-882
Cefuroxime4-842
Ceftazidime1-40.120.06
Cefotaxime0.75-30.120.12
Cefepime0.25-10.120.06
Moxalactam1-40.120.06
Aztreonam4-80.120.12
Imipenem0.12-0.50.50.06
Meropenem0.12-0.50.50.06
Open in a separate windowaCLA, clavulanic acid at a fixed concentration of 4 μg/ml; TZB, tazobactam at a fixed concentration of 4 μg/ml.To further characterize the activity of OXA-134, the enzyme was purified from E. coli cultures containing recombinant plasmid pOXA-134 as described previously (17). After DNase treatment and ultracentrifugation at 40,000 × g for 1 h, the extract was loaded successively onto two Q-Sepharose columns with 20 mM diethanolamine (pH 8.5) and 20 mM diethanolamine (pH 9.5) buffers. The specific activity of the purified β-lactamase OXA-134, measured with 100 μM imipenem as the substrate, was 116 U·mg of protein−1, with a 20-fold purification factor. The kinetic measurements of the purified enzymes were carried out at 25°C in 50 mM sodium phosphate (pH 7.0), and Km and kcat values were determined as described previously (6). β-Lactamase OXA-134 showed a narrow-spectrum hydrolysis profile, including mostly penicillins (Table (Table2).2). The rates of imipenem and meropenem hydrolysis were low, whereas the MICs of both carbapenems for E. coli TOP10 expressing OXA-134 were increased by 3-fold (Table (Table1).1). Overall, the catalytic activities obtained for OXA-134 were similar to those for OXA-58 and OXA-40, taken as references for CHDL activity (17).

TABLE 2.

Kinetic parameters for purified β-lactamase OXA-134a
Substratekcat (s−1)Km (μM)kcat/Km ratio (s−1·mM−1)
Benzylpenicillin70501,400
Ampicillin150250600
Ticarcillin0.22001
Piperacillin30200150
Ceftazidime<0.01NDb
Cefotaxime<0.01ND
Cefepime<0.01ND
Cefoxitin<0.01ND
Aztreonam<0.01ND
Imipenem0.11010
Meropenem0.052500.2
Open in a separate windowaData are means of results from three independent experiments. Standard deviations were within 10% of the means.bND, no detectable hydrolysis (<0.01 s−1).In order to assess whether the blaOXA-134-like gene was naturally present in A. lwoffii, a blaOXA-134-specific PCR was performed using whole-cell DNA samples from a collection of 10 A. lwoffii isolates recovered from clinical specimens, including blood cultures, urine samples, cerebrospinal fluids, and central venous catheter tips, from Bicêtre and Cologne hospitals. PCR results showed that all strains possessed a blaOXA-134-like gene. Sequencing of the amplicons allowed the identification of six additional OXA-134 derivatives (named OXA-186 to OXA-191) (see www.lahey.org/Studies) differing by 3 to 18 amino acid substitutions (Fig. (Fig.1).1). Noteworthily, OXA-134 and OXA-186 each possessed 273 amino acids whereas the five other variants each possessed an additional 3-amino-acid stretch (Fig. (Fig.1).1). In three isolates, the blaOXA-134-like gene was disrupted by nucleotide substitutions located in the center of the gene and likely leading to interruption of the open reading frame (data not shown). In order to evaluate whether an OXA-134-like variant possessing additional amino acids may confer a different β-lactam resistance pattern, the blaOXA-187 gene chosen as a representative was cloned and expressed in the same manner as the blaOXA-134 gene. MICs of β-lactams for E. coli(pOXA-187) were similar to those for E. coli(pOXA-134) (data not shown), showing that those additional amino acids did not play any significant role in hydrolysis. The chromosomal locations of the blaOXA-134-like genes in these A. lwoffii isolates were confirmed by using the endonuclease I-CeuI technique, as described previously (10).In order to evaluate whether blaOXA-134-like genes might have disseminated among A. baumannii strains, a collection of 100 A. baumannii isolates (with variable susceptibilities to imipenem, including 50 carbapenem-resistant isolates) were screened by PCR. None of the screened A. baumannii isolates harbored a blaOXA-134-like gene.A. lwoffii was found to be a reservoir of a novel type of CHDL-encoding gene. Detection of that β-lactamase gene might be used as a tool for rapid and accurate identification of the A. lwoffii species.  相似文献   

16.
The in vitro activity of ACHN-490, a novel aminoglycoside (“neoglycoside”), was evaluated against 102 multidrug-resistant (MDR) Klebsiella pneumoniae strains, including a subset of 25 strains producing the KPC carbapenemase. MIC50 values for gentamicin, tobramycin, and amikacin were 8 μg/ml, 32 μg/ml, and 2 μg/ml, respectively; MIC90 values for the same antimicrobials were ≥64 μg/ml, ≥64 μg/ml, and 32 μg/ml, respectively. ACHN-490 showed an MIC50 of 0.5 μg/ml and an MIC90 of 1 μg/ml, which are significantly lower than those of comparator aminoglycosides. ACHN-490 represents a promising aminoglycoside for the treatment of MDR K. pneumoniae isolates, including those producing KPC β-lactamase.The spread of Klebsiella pneumoniae isolates producing extended-spectrum β-lactamases (ESBLs) represents a serious threat to our therapeutic armamentarium (21). These isolates are also frequently resistant to other classes of antibiotics, such as β-lactam/β-lactamase inhibitor combinations, quinolones, and aminoglycosides (8, 9), thereby limiting our choice to carbapenems for the treatment of serious infections (21).Unfortunately, there is growing concern regarding the emergence of carbapenem-resistant K. pneumoniae isolates (20). In particular, K. pneumoniae isolates producing KPC carbapenemases (KPC-Kp) are spreading at an alarming rate in North and South America, the Caribbean, Europe, Israel, and Asia (6, 7, 15, 17, 18). Like ESBL producers, KPC-Kp are often resistant to quinolones and aminoglycosides (6). Therefore, our therapeutic options against KPC-Kp are limited to tigecycline and colistin. However, tigecycline may not reach desired serum levels to treat bloodstream infections (19), leaving colistin as the “last choice” against infections caused by KPC-Kp (13). Unfortunately, colistin-resistant KPC-Kp isolates are also reported in the United States (1, 12). As a result of this therapeutic dilemma, new antimicrobial agents with potent activity against multidrug-resistant (MDR) K. pneumoniae need to be developed.Recently, there has been an increased interest in developing novel aminoglycosides. This new attention is due to (i) the potent bactericidal activity of aminoglycosides against a wide spectrum of aerobic gram-positive and gram-negative pathogens, (ii) the more gradual decline in susceptibility to aminoglycosides among gram-negative bacteria than that in susceptibility to other antimicrobials, and (iii) the ability of novel aminoglycosides to bypass common mechanisms of resistance that have gradually decreased the susceptibility to clinically used aminoglycosides (e.g., gentamicin, tobramycin, and amikacin) (11, 14, 16).ACHN-490 (Achaogen, San Francisco, CA) is a “neoglycoside,” a next-generation aminoglycoside, currently in early clinical development (FDA, http://clinicaltrials.gov/), which has never been reported previously in the literature. The chemical structure of ACHN-490 is presented in Fig. Fig.11.Open in a separate windowFIG. 1.Chemical structure of ACHN-490 [6′-(hydroxylethyl)-1-(haba)-sisomicin].In the present work, we analyzed the in vitro activity of ACHN-490 against a collection of 102 K. pneumoniae clinical isolates collected from January 2006 to October 2007 at the University of Pittsburgh Medical Center, and three Cleveland institutions, including University Hospitals Case Medical Center, the Cleveland Clinic, and the Louis Stokes Department of Veterans Affairs Medical Center.The 102 K. pneumoniae isolates were selected based on an MDR phenotype (i.e., resistance to ≥3 antibiotic classes). Twenty-five isolates were KPC-Kp and were part of a previous study in which the β-lactamase background and clonality were characterized (6). The remaining 77 MDR K. pneumoniae isolates were ESBL producers, according to the phenotypic results (see below).MICs were determined by a microdilution method using cation-adjusted Mueller-Hinton broth, according to the Clinical and Laboratory Standards Institute (CLSI) criteria (2). Specific panels containing the following antibiotics were customized by Trek Diagnostics (Cleveland, OH): cefotaxime, cefotaxime-clavulanate (constant concentration of 4 mg/liter), ceftazidime, ceftazidime-clavulanate (constant concentration of 4 mg/liter), piperacillin-tazobactam, imipenem, ciprofloxacin, tigecycline, gentamicin, tobramycin, amikacin, arbekacin, neomycin, and ACHN-490. The following ATCC control strains were used: Escherichia coli 25922, Pseudomonas aeruginosa 27853, and K. pneumoniae 700603. Susceptibility results were interpreted according to the guidelines recommended by CLSI (3). Tigecycline MICs were interpreted according to the U.S. FDA criteria (i.e., susceptible at an MIC of ≤2 μg/ml). According to the CLSI criteria, isolates were defined as ESBL producers when they showed a ≥3 twofold concentration decrease in MICs for ceftazidime or cefotaxime when tested in combination with clavulanate versus their MICs when tested alone (3).The 25 KPC-Kp isolates were analyzed by PCR for the presence of 16S rRNA methylase genes (i.e., armA, rmtA, rmtB, rmtC, rmtD, and npmA), using primers and conditions previously reported (4, 23). In addition, these strains were examined by PCR and sequencing for the presence of the most common aminoglycoside-modifying enzymes (AMEs) in gram-negative pathogens (22). In particular, the following genes were analyzed: aac(6′)-Ib, aac(6′)-Ic, aac(6′)-Id, ant(3")-Ia, ant(2")-Ia, aac(3)-Ia, aac(3)-Ib, aac(3)-IIc, aph(3′)-VIa, and aph(3′)-VIb, using primers previously reported (5, 10).As shown in Table Table1,1, MDR K. pneumoniae isolates were highly resistant to ceftazidime and piperacillin-tazobactam (each MIC90, >32 μg/ml). Two-thirds of the isolates were resistant to ciprofloxacin, whereas approximately 75% and 90% of strains were still susceptible to imipenem and tigecycline, respectively. Almost all KPC-Kp isolates were resistant to β-lactams and quinolones, whereas tigecycline frequently remained active in vitro (Table (Table1).1). All of these 25 isolates were colistin susceptible, as previously reported (6).

TABLE 1.

Susceptibility results of MDR K. pneumoniae isolates, including those producing KPC enzymes
AntibioticAll MDR K. pneumoniae isolates (n = 102)
KPC-producing K. pneumoniae isolates (n = 25)
MIC50 (μg/ml)MIC90 (μg/ml)S (%)aMIC50 (μg/ml)MIC90 (μg/ml)S (%)a
Ceftazidime>32>329.8>32>320.0
Imipenem0.5875.58>1612.0
Piperacillin-tazobactam>64>6438.2>64>640.0
Ciprofloxacin41626.5>8>88.0
Tigecyclineb,c1290.21296.0
Amikacin23278.4323248.0
Gentamicin8≥6425.581644.0
Tobramycin32≥6410.832≥648.0
Arbekacinc416816
Neomycinc232232
ACHN-490c,d0.510.51
Open in a separate windowaS, susceptibility according to CLSI criteria (3): ceftazidime (MIC, ≤8 μg/ml); imipenem (MIC, ≤4 μg/ml); piperacillin-tazobactam (MIC, ≤16 μg/ml); ciprofloxacin (MIC, ≤1 μg/ml); amikacin (MIC, ≤16 μg/ml); gentamicin (MIC, ≤4 μg/ml); tobramycin (MIC, ≤4 μg/ml).bTigecycline was interpreted according to U.S. FDA criteria (susceptibility, MIC ≤ 2 μg/ml).cCLSI criteria not available.dE. coli ATCC 25922 (MICs, 0.5 to 1 μg/ml); P. aeruginosa ATCC 27853 (MIC, 4 μg/ml); K. pneumoniae ATCC 700603 (MICs, 0.25 to 0.5 μg/ml).Figure Figure22 shows our analysis of aminoglycoside susceptibility. MDR K. pneumoniae isolates were highly resistant to gentamicin and tobramycin (less than 26% of strains were susceptible). In contrast, amikacin still maintained in vitro activity (78% of isolates were susceptible) with only five isolates being fully resistant (i.e., MICs of 64 μg/ml). The subgroup of KPC-Kp showed lower susceptibility rates for amikacin and tobramycin (48% and 8%, respectively) than did the entire group of MDR strains (Fig. (Fig.2).2). Notably, gentamicin was more active in vitro against KPC-Kp (44% of strains susceptible) than against the overall MDR isolate group.Open in a separate windowFIG. 2.MIC distributions of amikacin, gentamicin, tobramycin, and ACHN-490 against the overall collection of MDR K. pneumoniae isolates (n = 102) and the subgroup of KPC-producing strains (n = 25). S, susceptible; I, intermediate; R, resistant. Results were interpreted according to CLSI criteria (3). Dashed vertical line, susceptibility cutoff; solid vertical line, resistance cutoff.For both MDR and KPC-Kp strains, ACHN-490 showed MIC50 and MIC90 values (i.e., 0.5 and 1 μg/ml, respectively) that were significantly lower than those for comparator aminoglycosides. The ACHN-490 MICs for all strains were ≤4 μg/ml. In particular, the MIC90 of ACHN-490 was at least 5 twofold dilutions lower than that of amikacin, which is currently the aminoglycoside with the least resistance in our armamentarium (Fig. (Fig.22).To better understand the impact of these susceptibility data, we investigated the genetic background of KPC-Kp isolates in terms of their AMEs and methylases. All KPC-Kp strains were positive for aac(6′)-Ib and ant(3")-Ia (alternative name of aadA1) AME genes. Since neither of these AMEs modifies gentamicin, this explains the lower level of gentamicin resistance observed in the KPC-Kp strains. In contrast, the AAC(3)-II enzyme is common among Enterobacteriaceae and may be generating gentamicin resistance among the non-KPC-positive isolates (16). Two KPC-Kp strains (i.e., VA362 and VA373) were also positive for the ant(2")-Ia gene. Consistent with our MIC results (i.e., all strains with arbekacin MICs of <32 μg/ml) and the low prevalence in the clinical population, we did not find any methylase genes. An E. coli control strain in which the rmtA methylase gene was cloned had an MIC of >8 μg/ml for ACHN-490.In conclusion, ACHN-490 possesses potent in vitro activity against MDR K. pneumoniae isolates, including those producing KPC carbapenemase. ACHN-490 represents a promising alternative to tigecycline and colistin for the treatment of isolates resistant to quinolones, β-lactam/β-lactamase inhibitor combinations, carbapenems, and existing aminoglycosides.  相似文献   

17.
Biophysical methods to study the binding of oritavancin, a lipoglycopeptide, to serum protein are confounded by nonspecific drug adsorption to labware surfaces. We assessed oritavancin binding to serum from mouse, rat, dog, and human by a microbiological growth-based method under conditions that allow near-quantitative drug recovery. Protein binding was similar across species, ranging from 81.9% in human serum to 87.1% in dog serum. These estimates support the translation of oritavancin exposure from nonclinical studies to humans.Estimates of serum protein binding are essential to translate drug exposure from nonclinical species to humans during assessments of toxicology, pharmacokinetics, and pharmacodynamics since the free fraction dictates drug activity (3, 7, 17, 18). Recent evidence supports the concept of an “active fraction” that offers insight into the pharmacodynamic behavior of highly protein-bound drugs, such as daptomycin (20).Oritavancin is a late-stage investigational lipoglycopeptide under study for treatment of serious Gram-positive infections (6). Nonspecific binding of oritavancin to labware surfaces (1, 2) and to dialysis membranes has called into question the accuracy of previous oritavancin human serum binding estimates (85.7% to 89.9% [16]). We therefore used conditions that minimize nonspecific oritavancin binding (4, 5) to estimate its binding to serum by a single in vitro methodology for three nonclinical species (mouse, rat, and dog) and humans. Protein binding estimates were derived from serum-induced increases in oritavancin MICs (9, 10, 21). To control for any impact of serum components on bacterial growth and antibiotic activity, oritavancin activity in serum was compared to its activity in serum ultrafiltrate, which is devoid of albumin, the protein responsible for the majority of oritavancin serum binding (23). The method was benchmarked using daptomycin and ceftriaxone (8, 18, 22). (Part of this work was previously presented at the 19th European Congress of Clinical Microbiology and Infectious Diseases as a poster [12].)Pooled serum from humans, mice, and rats was from Equitech-Bio (Kerrville, TX); pooled serum from beagle dogs was from Bioreclamation (Liverpool, NY). Serum ultrafiltrate was prepared using Centricon Plus-50 ultrafilters (Millipore, Billerica, MA), whose molecular mass cutoff (50 kDa) excludes albumin. MICs against Staphylococcus aureus ATCC 29213 were determined by broth microdilution (4) using arithmetic drug dilutions in 95% serum or 95% serum ultrafiltrate, each supplemented with 5% cation-adjusted Mueller-Hinton broth (CAMHB). Serum protein binding for each drug was calculated using the following formula: % bound = (1 − [mean MIC in serum ultrafiltrate/mean MIC in serum]) × 100.The MICs for each condition, serum source, and test agent were precise (Table (Table1),1), with a mean coefficient of variation of 17%. MICs as determined under CLSI M7-A8 conditions (Table (Table1)1) (4) were within the quality control ranges (5).

TABLE 1.

Oritavancin, ceftriaxone, and daptomycin MICs against S. aureus ATCC 29213 in cation-adjusted Mueller-Hinton broth and 95% serum ultrafiltrate and 95% serum from human, mouse, rat, and dog
SpeciesMIC (μg/ml)
Ceftriaxoneb
Oritavancina
Daptomycinc
CAMHBUltrafiltrateSerumCAMHBdUltrafiltrateeSerumfCAMHBUltrafiltrateSerum
Humang
    Mean4.882.8838.80.0840.1400.7750.9750.5133.00
    SD0.8350.35411.00.0050.0380.3240.0460.1250.535
Mouseh
    Mean5.003.756.000.1050.0790.5380.9753.0012.5
    SD0.8160.5001.160.0300.0040.0520.0502.89
Ratg
    Mean3.503.885.880.0860.0550.3131.250.5381.56
    SD0.5350.3540.6410.0070.0050.0990.2670.0520.32
Dogg
    Mean5.251.091.380.0800.0610.4751.000.6382.50
    SD0.7070.5820.51800.0140.04600.1500.530
Open in a separate windowaArithmetic dilution steps of 0.5 μg/ml from 3 to 1 μg/ml, of 0.1 μg/ml from 1 to 0.3 μg/ml, of 0.05 μg/ml from 0.3 to 0.1 μg/ml, and of 0.01 μg/ml from 0.1 to 0.04 μg/ml were prepared in cation-adjusted Mueller-Hinton broth (CAMHB) containing 0.002% polysorbate-80.bArithmetic dilution steps of 10 μg/ml from 100 to 10 μg/ml and of 1 μg/ml from 10 to 1 μg/ml were prepared in cation-adjusted Mueller-Hinton broth.cArithmetic dilution steps of 5 μg/ml from 20 to 10 μg/ml, of 1 μg/ml from 10 to 2 μg/ml, of 0.5 μg/ml from 2 to 1 μg/ml, and of 0.1 μg/ml from 1 to 0.3 μg/ml were prepared in cation-adjusted Mueller-Hinton broth supplemented with 50 μg/ml CaCl2.dMICs determined by CLSI M7-A8 guidelines in cation-adjusted Mueller-Hinton broth, supplemented with 0.002% polysorbate-80 (oritavancin) or 50 μg/ml CaCl2 (daptomycin) (5).eMICs determined in 95% serum ultrafiltrate plus 5% cation-adjusted Mueller-Hinton broth.fMICs determined in 95% serum plus 5% cation-adjusted Mueller-Hinton broth.gMeans were derived from 8 replicates per condition per drug.hMeans were derived from 4 to 8 replicates per condition per drug.Increases in the oritavancin MICs in serum compared to its MICs in serum ultrafiltrate, by species, were similar across species (5.5- to 7.8-fold) (Table (Table2).2). Such shifts yielded similar mean values of oritavancin serum protein binding for the four species tested (81.9% to 87.1%) (Table (Table2).2). The 81.9% human serum protein binding estimate from this study falls within the 79% to 89.9% range of previously reported values from growth-based or biophysical approaches (summarized in Table Table3).3). Our finding supports the premise that growth-based methods can complement biophysical methods in the estimation of the free fraction of antibiotics.

TABLE 2.

Serum-induced increases in broth microdilution MICs against S. aureus ATCC 29213 and corresponding protein binding estimates for oritavancin, ceftriaxone, and daptomycin
Serum sourceOritavancin
Ceftriaxone
Daptomycin
Mean fold MIC increasea% BoundbMean fold MIC increase% BoundMean fold MIC increase% Bound
Human5.581.913.592.65.882.9
Mouse6.885.31.637.54.276.0
Rat5.782.41.534.02.965.6
Dog7.887.11.320.93.974.5
Open in a separate windowaRatio of the mean arithmetic MIC in 95% serum to the mean arithmetic MIC in 95% serum ultrafiltrate.bCalculated from mean MICs using the following formula: percent protein bound = [1 − (MIC in ultrafiltrate/MIC in serum)] × 100.

TABLE 3.

Oritavancin serum protein binding estimates for human, mouse, rat, and dog
SpeciesMatrixProtein bindinga (%)MethodOritavancin concn (μg/ml)Reference
HumanPlasma87.5Broth microdilutionVarious23
Plasma85.7-89.9DCCb adsorption1-9116
Albumin79 ± 0.2Cantilever nanosensorc0.2R. A. McKendry, unpublished
Serum81.9Broth microdilutionVariousThis study
MouseSerum85.3Broth microdilutionVariousThis study
RatPlasma>80Broth microdilutionVarious23
Serum82.4Broth microdilutionVariousThis study
DogSerum87.1Broth microdilutionVariousThis study
Open in a separate windowaStandard deviation value is provided where available.bDextran-coated charcoal.cSee reference 14.Oritavancin was found to bind rat serum at 82.4% in the present study; this concurs with the >80% binding to rat plasma using a broth microdilution approach (Table (Table3)3) (23). Oritavancin binding to serum of beagle dogs, a species which had not been evaluated prior to the present study despite its importance in nonclinical toxicology assessments, was estimated at 87.1% (Table (Table2).2). Our results showing a similar extent of oritavancin protein binding to human, mouse, rat, and dog serum should facilitate the translation of drug exposure between these species since the free fraction of oritavancin is likely to be equivalent across species, within the error of measurement of any single assay.Comparison of the assessment of the area under the bacterial kill curves (10) for oritavancin determined in the presence of serum and in the presence of serum ultrafiltrate yielded protein binding values of 67.4, 63.9, and 61.7% for human serum (at 0.5, 1, and 2 μg/ml oritavancin, respectively) and of 66.5, 68.3, and 68.8% for mouse serum (at 0.5, 1, and 2 μg/ml oritavancin, respectively) (12). While these estimates are lower than those derived from the analysis of arithmetic MIC shifts in human and mouse serum noted above, they may be explained at least in part by the rapid killing kinetics of oritavancin (11) that cannot be surmised from the MIC shift endpoints of broth microdilution assays.Ceftriaxone was highly bound to human serum (92.6%) (Table (Table2),2), in agreement with both Yuk et al. (22) and MIC shift assessments by Schmidt et al. (18) but substantially higher than the 76.8% binding estimate derived from in vitro microdialysis (18). Variability in ceftriaxone serum protein binding across species (15, 18) was also noted in the present study, with substantially lower binding estimates for serum from mouse, rat, and beagle dog (range, 20.9% to 37.5%) than for human serum. These differences may result from true species-specific binding affinity differences (15) or from methodological differences during the isolation or assay of serum from each species.Daptomycin binding to serum protein also varied across species in the present study, ranging from 65.6% (rat) to 82.9% (human) (Table (Table2).2). For human serum, this value falls between the values of 58% reported by Tsuji et al. (21) and 94% reported by Lee et al. (8). The implications of such variability are potentially important during the translation of nonclinical findings to humans, for example, in pharmacokinetic-pharmacodynamic target attainment studies to support susceptibility breakpoint proposals (13).While it is difficult to assess the accuracy of serum protein binding estimates from any single method, the precision of our cross-species comparative study, the concordance of single-species data from different methods, and the similarity of binding estimates across different species suggest that oritavancin is approximately 85% bound to serum protein and that differences in oritavancin protein binding across species are negligible. This conclusion is similar to one from studies of telavancin, another lipoglycopeptide, in which plasma protein binding was approximately 90% across tested species (19), although this value was substantially higher than the 62 to 70% estimates determined using a growth-based assay (21). The approximately 65% protein binding estimates from time kill-based assays with oritavancin (12) support the idea that the active fraction (20) of oritavancin, namely, its bioactive concentration in the presence of serum protein, is greater than the free fraction as predicted from biophysical approaches. Whether this conclusion applies to other lipoglycopeptides remains to be determined.  相似文献   

18.
The in vitro activity of azithromycin against 1,237 nontyphoidal Salmonella enterica isolates collected from Finnish patients between 2003 and 2008 was investigated. Only 24 (1.9%) of the isolates tested and 15 (5.1%) of the 294 isolates with reduced fluoroquinolone susceptibility had azithromycin MICs of ≥32 μg/ml. These data show that azithromycin has good in vitro activity against nontyphoidal S. enterica, and thus, it may be a good candidate for clinical treatment studies of salmonellosis.Salmonella is one of the most common causes of food-borne illnesses and a major cause of human infections all over the world (23). Salmonella infections are usually treated with fluoroquinolones or extended-spectrum cephalosporins. Unfortunately, excessive use of fluoroquinolones both in human and in veterinary medicine has led to increasing numbers of resistant isolates, including nontyphoidal strains of Salmonella enterica (18, 19, 28). In addition, the nonclassical quinolone resistance phenotype (the Qnr phenotype), showing reduced susceptibility to ciprofloxacin (MIC of ≥0.125 μg/ml) but susceptibility or only low-level resistance to nalidixic acid (MIC of ≤32 μg/ml), has become more common (6, 14, 19, 20, 22). Extended-spectrum β-lactamase (ESBL) producers have emerged in Enterobacteriaceae and in Salmonella, and there are reports of the coappearance of ESBL and qnr genes in the same transferable genetic elements (5, 10, 21, 27). These resistance problems may jeopardize the treatment of severe Salmonella infections. Thus, alternative antibiotics for the treatment of Salmonella infections are needed.Salmonella isolates are intrinsically resistant to erythromycin via active efflux (2) but naturally susceptible to azithromycin (29), which is a 15-membered erythromycin derivative. Resistance to macrolides is usually conferred by mutations in nucleotides A2058 and A2059 of the 23S rRNA, according to the Escherichia coli numbering (26). Also, the alteration of the 50S ribosomal subunit proteins L4 (rlpD) and L22 (rlpV) may lead to macrolide resistance (4).The purpose of the present study was to determine the in vitro activity of azithromycin against nontyphoidal Salmonella isolates collected between 2003 and 2008 from Finnish patients. Special attention was paid to isolates with reduced fluoroquinolone susceptibility or showing the Qnr phenotype. In addition, mutations in the 23S rRNA and in the L4 and L22 ribosomal proteins were investigated.(This work was presented in part at the 19th European Congress of Clinical Microbiology and Infectious Diseases [ECCMID], Helsinki, Finland, 2009 [P1043].)A total of 1,237 nontyphoidal Salmonella isolates (638 domestic and 599 foreign) collected from Finnish patients between 2003 and 2008 were included in this study. Starting in January each year, we collected the first 100 domestic and the first 100 foreign, i.e., collected from Finnish travelers returning from abroad, Salmonella isolates. The strains were serotyped at the National Salmonella Reference Centre in Finland.The MICs of antimicrobial agents for the isolates were determined by the agar dilution method according to the CLSI guidelines (8). Mueller-Hinton II agar (Becton Dickinson, Cockeysville, MD) was used as the culture medium. All 1,237 Salmonella isolates were tested for susceptibility to ciprofloxacin, nalidixic acid, and azithromycin (all from Sigma, Steinheim, Germany). We tested 809 selected isolates for erythromycin (Sigma) and 635 for telithromycin (Sanofi Aventis, Paris, France) susceptibility. Control strains for susceptibility testing were as described previously (14). On the basis of earlier publications (1, 16, 17), the MIC breakpoint chosen for reduced ciprofloxacin susceptibility was ≥0.125 μg/ml. For nalidixic acid, CLSI breakpoints were used (9). Based on the EUCAST recommendation for S. enterica serovar Typhi (www.eucast.org) and a previous publication (3), the epidemiological cutoff value chosen for azithromycin was ≥32 μg/ml. The susceptibility data were analyzed by using the WHONET 5.4 computer program.Pyrosequencing was used to detect macrolide resistance causing point mutations in the ribosomal target sites A2058 and A2059 (E. coli numbering) of the 23S rRNA in 22 isolates with azithromycin MICs of ≥32 μg/ml, 44 isolates belonging to the Qnr phenotype, and 73 isolates showing erythromycin MICs of ≥32 μg/ml. Pyrosequencing was performed with previously described primers and protocols (15) using a PSQ 96MA pyrosequencer.The 50S ribosomal proteins L4 and L22 were amplified, and mutations were screened in 24 isolates having azithromycin MICs of ≥32 μg/ml, of which 6 were of the Qnr phenotype. In addition, 13 isolates having only erythromycin MICs of ≥32 μg/ml and showing the Qnr phenotype were tested. DNA was prepared as described above. The specific primers used for amplification of the complete L4 and L22 genes rlpD and rlpV were Salm_L4_f (5′-TGAAGGCGTAAGGGGATAGCA-3′) and Salm_L4_r (5′-TCAGCAGA CGTTCTTCACGAA-3′) and Salm_L22_f (5′-GAAATAAGGTAG GAGGAAGAG-3′) and Salm_L22_r (5′-CCATTGCTAGTCTCCAGAGTC-3′). The PCR conditions were as follows: 94°C for 10 min, 94°C for 30 s, 56°C for 30 s, and 72°C for 60 s for 33 cycles. Any L4- or L22-positive results were confirmed by direct sequencing of both strands of amplicons using specific PCR primers as previously described (24). Amino acid sequences were then compared with the known rlpD and rlpV genes by a BLAST search through the European Bioinformatics Institute (http://www.ebi.ac.uk/Tools/blast/).Two different populations of S. enterica were detected regarding the azithromycin MIC distribution. The majority of the isolates had azithromycin MICs of 4 to 8 μg/ml, i.e., representing the wild-type population, whereas a minority of the isolates had MICs of 32 to ≥128 μg/ml, i.e., over the epidemiological cutoff value. Between 2003 and 2008, 24 (1.9%) of the 1,237 isolates had azithromycin MICs of ≥32 μg/ml (Fig. (Fig.1;1; Table Table1).1). Nine (1.4%) of the 638 domestic isolates and 15 (2.1%) of the 599 foreign isolates had azithromycin MICs over the epidemiological cutoff value.Open in a separate windowFIG. 1.Histograms of azithromycin (n = 1,237), erythromycin (n = 809), and telithromycin (n = 635) MICs for S. enterica isolates collected from Finnish travelers between 2003 and 2008. The vertical line represents the resistance breakpoint.

TABLE 1.

Twenty-four S. enterica isolates showing azithromycin MICs of ≥32 μg/ml
Isolate no.YrStrainOriginS. enterica serovarAZMa MIC (μg/ml)qnr geneb
12003s2099ThailandNewport32
22003s2018ThailandStanley64+
32003s2021ThailandRissen>128
42003s2085ThailandStanley64+
52003s2086ThailandStanley64+
62003s2181FinlandPoona32
72003s2137FinlandTyphimurium128
82003s2195FinlandVirchow>128
92004s2236ThailandStanley32+
102004s2265ThailandStanley128+
112004s2280ThailandStanley64+
122004s2389FinlandTyphimurium128
132004s2391FinlandTyphimurium128
142005s2439EgyptBlockley64
152005s2477EgyptBredeney>128
162005s2608FinlandBlockley64
172006s2635ThailandEmek64
182006s2768FinlandBlockley128
192006s2829FinlandSaintpaul64
202007s2868MalaysiaStanley128
212007s2948FinlandHvittingfoss128
222008s3141ThailandRissen64
232008s3082ThailandTyphimurium128
242008s3139ThailandEnteritidis128
Open in a separate windowaAZM, azithromycin.b+, present; −, absent.Two hundred ninety-four (23.8%) S. enterica isolates showed reduced fluoroquinolone susceptibility, and 53 (18.0%) of them showed the Qnr phenotype. Among the isolates with reduced fluoroquinolone susceptibility, 4 (5.2%) domestic and 11 (5.1%) foreign isolates had azithromycin MICs of ≥32 μg/ml. Among the Qnr phenotype isolates, six (11.3%) had azithromycin MICs of ≥32 μg/ml (Table (Table1).1). Azithromycin MICs of ≥32 μg/ml were detected among 12 different serovars, S. enterica serovar Stanley being the most common one, and all isolates with a Qnr phenotype belonged to S. enterica serovar Stanley (Table (Table11).Of the S. enterica isolates tested, 99.6% (806/809) had erythromycin MICs of ≥32 μg/ml. The erythromycin MICs varied between 8 and ≥128 μg/ml, and the vast majority of the isolates had erythromycin MICs of ≥64 μg/ml (Fig. (Fig.11).No mutations in A2058 or A2059 of the 23S rRNA gene were detected among any of the isolates tested. Sequencing of the 50S ribosomal proteins L4 and L22 revealed G235A and C379T mutations in the rlpD gene and G25A in the rlpV gene, respectively. A Glu79-Lys substitution in the rlpD gene was found in six isolates belonging to three different serovars (Table (Table2).2). An Arg127-Trp substitution in the rlpD gene was found in three S. enterica serovar Montevideo isolates, which also had an Asp9-Asn substitution outside the coding region (Table (Table22).

TABLE 2.

Mutations found in the 50S ribosomal proteins L4 and L22 and corresponding MICs
S. enterica serovarOriginYrMutation in L4 and L22
MICa
rlpDrlpVAZMERYTELCIPNAL
BlockleyEgypt2005G235A64>128>1280.5>512
BlockleyFinland2005G235A64>128>1280.25>512
BlockleyFinland2006G235A128>128>1280.25>512
SaintpaulFinland2006G235A64>128>1280.0316
TyphimuriumThailand2008G235A128>128>1280.0616
TyphimuriumFinland2004G235A12864NDb0.25512
MontevideoThailand2004C379TG25A464320.532
MontevideoThailand2007C379TG25A8>128320.516
MontevideoThailand2008C379TG25A8>12840.516
Open in a separate windowaAZM, azithromycin; ERY, erythromycin; TEL, telithromycin; CIP, ciprofloxacin; NAL, nalidixic acid.bND, not determined.Salmonella isolates are intrinsically resistant to erythromycin via the AcrAB efflux pump (2) but naturally susceptible to azithromycin (29). Azithromycin has shown good efficacy in the treatment of patients suffering from typhoid fever (7, 12, 13, 25, 30), and there has been speculation that azithromycin could be used for empirical therapy of traveler''s diarrhea (11, 31). The present study was performed to determine the in vitro activity of azithromycin toward nontyphoidal Salmonella isolates collected from Finnish patients between 2003 and 2008. Our results show that while nearly all (99.6%) of our S. enterica isolates had erythromycin MICs of ≥32 μg/ml, only 24 isolates had azithromycin MICs of ≥32 μg/ml. Azithromycin showed good in vitro efficacy also against S. enterica isolates with reduced fluoroquinolone susceptibility, although 11.3% of the Qnr phenotype isolates showed azithromycin MICs of ≥32 μg/ml.These data show that azithromycin has good in vitro activity against nontyphoidal S. enterica isolates. Although highly azithromycin-resistant isolates did occur, azithromycin was effective even against the isolates with reduced fluoroquinolone susceptibility, including those showing the Qnr phenotype. Based on these results, azithromycin is a good candidate for clinical treatment studies of salmonellosis.  相似文献   

19.
We evaluated the in vitro activity of fosfomycin against a total of 192 CTX-M β-lactamase-producing Escherichia coli strains isolated in 70 Japanese clinical settings. Most of the isolates (96.4%) were found to be susceptible to fosfomycin. On the other hand, some of the resistant isolates were confirmed to harbor the novel transferable fosfomycin resistance determinants named FosA3 and FosC2, which efficaciously inactivate fosfomycin through glutathione S-transferase activity.Clinical efficacy of an old antibiotic, fosfomycin, is being reassessed owing to its in vitro high potent activity against multidrug-resistant Gram-negative bacilli belonging to the family Enterobacteriaceae (5, 6, 8). In the present study, we investigated the prevalence of fosfomycin resistance among CTX-M extended-spectrum β-lactamase (ESBL)-producing Escherichia coli clinical isolates in Japan and clarified the molecular mechanisms underlying the fosfomycin resistance, with special focus on exogenous resistance determinants, like the FosATN protein (9).A total of 192 CTX-M ESBL-producing E. coli isolates, which were collected from 70 medical facilities throughout Japan between 2002 and 2007, were retrospectively subjected to fosfomycin susceptibility testing with the agar dilution method according to the CLSI guideline (4). The result is shown in Fig. Fig.1.1. Most of the strains (96.4%) investigated were susceptible to fosfomycin (MIC, ≤64 μg/ml), while seven isolates (3.6%) showed nonsusceptibility to fosfomycin (MIC, ≥128 μg/ml). It seems likely that CTX-M-producing E. coli isolates that have acquired fosfomycin resistance are infrequent in Japan, and these data suggest the probable clinical efficacy of fosfomycin for the treatment of infectious diseases, like urinary tract infections (UTIs), caused by CTX-M-producing E. coli to some extent.Open in a separate windowFIG. 1.Distribution of fosfomycin MICs for the 192 CTX-M-producing E. coli isolates.We evaluated the fosfomycin resistance mechanism of the 10 isolates and found reduced susceptibility to fosfomycin (MIC, ≥64 μg/ml) (Fig. (Fig.11 and Table Table1).1). The transmissibility of the fosfomycin resistance determinant in the 10 isolates was investigated, and it was found that the nature of fosfomycin resistance of three strains, 08-642, 06-607, and C316, was successfully transferred to a recipient E. coli strain. The cefotaxime resistance phenotype was cotransferred to a recipient strain with the fosfomycin resistance (Table (Table11).

TABLE 1.

Characteristics of E. coli strains used in the study
E. coli isolateCharacteristic(s) of murA, uhpA, uhpT, and/or glpT geneMIC (μg/ml) of:
FosfomycinCefotaxime
Clinical isolates
    08-555glpT with 981[227-bp deletion]1209>256>128
    08-642>25664
    06-607>25616
    05-244glpT with 328[14-bp duplication]343, uhpT with 1173[96-bp deletion]1270128>128
    05-690uhpA stop at amino acid 1446464
    03-271Failure in PCR amplification of uhpA and uhpT6432
    03-285128>128
    03-286glpT with 405[5-bp duplication]41164>128
    03-287glpT with 405[5-bp duplication]411>256128
    C316256128
CSH-2 conjugants
    CSH-2(p08-642)Conjugant of strain 08-642>256128
    CSH-2(p06-607)Conjugant of strain 08-607>25632
    CSH-2(pHPA)Conjugant of strain C316>25664
    CSH-2Resistant to rifampin and nalidixic acid1≤0.06
DH10B transformants
    DH10B(pK-fosA3)Contains KpnI fragment with fosA3 from p08-642>256≤0.06
    DH10B(pS-fosA3)Contains SacII fragment with fosA3 from p06-607>256≤0.06
    DH10B(pS-fosC2)Contains SacII fragment with fosC2 from pHPA>256≤0.06
    DH10B(pBCKS+)Resistant to streptomycin and chloramphenicol0.5≤0.06
Open in a separate windowThe DNA fragments containing fosfomycin resistance determinants were cloned from the conjugative plasmids of E. coli 08-642, 06-607, and C316 strains and partially sequenced (Table (Table1).1). The fosfomycin resistance determinants and their genetic neighboring regions are shown in Fig. Fig.2.2. The KpnI ca. 8-kb fragment cloned from the transferable plasmid of E. coli 08-642 and the SacII ca. 10-kb fragment from that of E. coli 06-607 included the same nucleotide region flanked by IS26 (Fig. (Fig.2).2). The deduced amino acid sequences of one open reading frame (named fosA3) showed 70% identities to those of FosATN, the Mn(II)- and K+-dependent glutathione (GSH) S-transferase from Tn2921 of Serratia marcescens (2, 3) and 59% identities to that of FosAPA from Pseudomonas aeruginosa (Fig. (Fig.3)3) (1, 11). The fosA3 gene is likely to be responsible for fosfomycin resistance in strains 08-642 and 06-607.Open in a separate windowFIG. 2.Genetic environment of transferable fosfomycin resistance determinants and their neighboring regions in E. coli strains 08-624, 06-607, and C316.Open in a separate windowFIG. 3.Predicted amino acid sequences of fosfomycin resistance determinants. *, amino acid residues conserved among the seven fosfomycin resistance determinants; colons and dots, amino acid substitutions that result in homologous amino acid residues. Proteins (GenBank accession no.): FosATN (AAA98399); FosAPA (AAT49669); FosA2 (ACC85616); FosA3 (AB522970); ORF1 (AAP50248); FosC (AAZ14834); FosC2 (AB522969).The 1.8-kb region containing orf1 to Δorf3 at the 3′ end of fosA3 had 78% nucleotide identity with a part of the chromosome sequence of Klebsiella pneumoniae strain 342 (Fig. (Fig.2)2) (7). Moreover, this 1.8-kb homology region on the chromosome of K. pneumoniae strain 342 was close to the fosA gene. FosA of K. pneumoniae strain 342 has 80% amino acid identity to the FosA3 found in the present study. Although the precise physiological function of chromosomally encoded FosA proteins of K. pneumoniae remains to be determined, it is speculated that these proteins are the origin of a plasmid-mediated fosfomycin-modifying enzyme like FosA3.Additionally, one open reading frame, named fosC2, was found in the fragment cloned from the conjugative plasmid of E. coli strain C316 (Table (Table1)1) (13). The amino acid sequence of FosC2 had 72%, 56%, and 51% identity to that of FosC found in Achromobacter xylosoxidans (GenBank accession no. DQ112222), FosATN, and FosAPA, respectively (Fig. (Fig.3).3). The fosC2 gene was the first gene cassette in a class 1 integron accompanied by dfrA17 and aadA5 (Fig. (Fig.22).No transfer of fosfomycin resistance determinants was observed in the seven E. coli strains showing reduced susceptibility to fosfomycin (MIC, ≥64 μg/ml) (Table (Table1).1). Next, already-known chromosomally derived genes glpT, uhpT, uhpA, and murA, which are involved in fosfomycin resistance, were investigated (Table (Table1)1) (10, 12). The primers used in the present study are listed in Table Table2.2. Several outcomes supposed to be involved in fosfomycin resistance were observed in six of the strains, but no remarkable change was detected in strain 03-285 among the investigated genes. Although the extent to which the fosfomycin resistance conferred by chromosomally encoded factors described above remains controversial, these factors would partially explain the fosfomycin resistance in the clinical isolates.

TABLE 2.

Primers used in the study
PrimerSequence
uhpT-forward5′-ATG CTG GCT TTC TTA AAC C-3′
uhpT-reverse5′-TTA TGC CAC TGT CAA CTG C-3′
uhpA-forward5′-ATC ACC GTT GCC CTT ATA GA-3′
uhpA-reverse5′-TCA CCA GCC ATC AAA CAT-3′
murA-forward5′-CTC CAG GGC GAA GTC ACA-3′
murA-reverse5′-GCC TTT CAC ACG CTC AAT A-3′
glpT-forward5′-ATG TTG AGT ATT TTT AAA CC-3′
glpT-reverse5′-TAG CCT CCG TTG CGT TTT TG-3′
Open in a separate windowFinally, we purified C-terminal histidine-tagged FosA3 and FosC2 with the combination of a pET29a vector and E. coli BL21(DE3)(pLysS) and determined the enzymatic characteristics through a bioassay. Assays were performed in 50 mM HEPES buffer (pH 7.8) containing 100 mM KCl, 0.05 mM MnCl2, 5 mM fosfomycin, 10 mM glutathione, and 10 μM purified protein in a final volume of 100 μl at 35°C for 30 min, and the reaction was quenched by adding methanol. Ten microliters of sample solution was added to a blank disc set on an agar plate inoculated with E. coli ATCC 25922, and the remaining antibacterial activity was measured as a growth inhibition zone. When a sample solution containing only fosfomycin and GSH was added, a 21-mm inhibitory zone was observed. When the same sample supplemented with FosA3 or FosC2 was added, no inhibitory zone was observed around the disc. No decrease in the growth inhibition zone was observed when the sample containing only fosfomycin and purified proteins were added. The consumption of GSH catalyzed by FosA3 and FosC2 was confirmed using Ellman''s reagent. These results indicated that FosA3 and FosC2 inactivated fosfomycin by exerting GSH S-transferase activity, very similar to FosATN and FosAPA (1, 3).In conclusion, we report here the prevalence of fosfomycin resistance among CTX-M-producing E. coli isolates in Japan, together with the emergence of two novel plasmid-borne fosfomycin-modifying enzymes, FosA3 and FosC2. The fosfomycin resistance rate in CTX-M-producing E. coli is still low (3.6%) in Japan, but the fosfomycin resistance genes were already indwelling in the transferable plasmid of ESBL-producing clinical isolates. Continuous monitoring will be necessary to prevent further dissemination of fosfomycin resistance genes, together with prudent use of fosfomycin in clinical settings.  相似文献   

20.
Bacterial resistance presents a difficult issue for fluoroquinolone treatment of bacterial infections. In previous work, we reported that 8-methoxy-quinazoline-2,4-diones are active against quinolone-resistant mutants of Escherichia coli. Here, we demonstrate the activity of a representative 8-methoxy-quinazoline-2,4-dione against quinolone-resistant gyrases. Furthermore, 8-methoxy-quinazoline-2,4-dione and other diones are shown to inhibit Staphylococcus aureus gyrase and topoisomerase IV with similar degrees of efficacy, suggesting that the diones might act as dual-targeting agents against S. aureus.Antibiotic resistance is among the most difficult problems we currently face during the treatment of bacterial infections (10, 32). The fluoroquinolones are among the antibacterials affected by resistance, which can severely limit their clinical use (1, 6, 29). Thus, there is an urgent need to develop antimicrobial agents that are effective against drug-resistant pathogens. Two properties of quinolone-like compounds are likely to be useful for finding effective derivatives: (i) activity against quinolone-resistant mutants already present (12) and (ii) equal effectiveness against the two quinolone targets, DNA gyrase and topoisomerase IV (Topo IV) (dual-targeting agents; dual-targeting agents are expected to slow the emergence of drug-resistant mutants [4, 19, 23, 26, 31]). The presence of an 8-methoxy group (5, 22, 33) and changing the quinolone core structure to either quinazoline-2,4-dione (2, 8, 18) or pyrido[1,2-c]pyrimidine-1,3-dione (UI7) improve the activity of fluoroquinolone-like compounds against fluoroquinolone-resistant bacteria. We recently identified 8-methoxy-quinazoline-2,4-diones that show little increase in MIC due to gyrA or gyrB quinolone resistance mutations of Escherichia coli (12). To establish that this activity against mutant bacteria is due to improved activity against gyrase, we examined the activity of a representative 8-methoxy-quinazoline-2,4-dione (UING5-207; 8-methoxy 2,4-dione) with purified gyrase. We also assessed its effect on the activity of purified Topo IV to determine whether the in vivo results were likely due to a target switch. Furthermore, to better understand how dione structure influences target selection, we compared the effectiveness of diones for inhibition of catalytic activities of Staphylococcus aureus and E. coli topoisomerases.Based on MIC values determined in previous work (12), we selected three mutant gyrases, GyrA S83W gyrase, GyrA G81C gyrase, and GyrA A67S gyrase, as examples exhibiting high, moderate, and low levels of quinolone resistance in vivo. Mutations were introduced into the E. coli gyrA gene using the overlap extension PCR technique (17), and subunits of E. coli and S. aureus gyrases and Topo IVs were expressed and purified. The active enzymes were reconstituted as described previously (13-16, 28). In vitro activities of the 8-methoxy 2,4-dione were compared with those of a cognate 8-methyl-quinazoline-2,4-dione (UIJR1-048; 8-methyl 2,4-dione), 5-methoxypyrido[1,2-c]pyrimidine-1,3-dione (UIJR1-100; 5-methoxy 1,3-dione), 8-methoxy fluoroquinolone (UING5-249), and ciprofloxacin, a clinically important fluoroquinolone. Synthesis of the diones and 8-methoxy fluoroquinolone was achieved using methods described previously (7, 12, 30); their structures are shown in Fig. Fig.1.1. A DNA supercoiling assay (for example, see Fig. Fig.2)2) was employed to examine the effects of these compounds on the catalytic activities of wild-type and mutant gyrases; a decatenation assay was used with Topo IV (Table (Table1).1). The abilities of these compounds to poison the E. coli topoisomerases were assessed using a DNA cleavage assay (Table (Table2).2). These assays were conducted as described previously (24). The activity against mutant enzymes was expressed as the ratio of either the 50% inhibitory concentration (IC50) or the CC3 value (the concentration required to triple the level of DNA cleavage from the background level in the absence of drug) for a mutant gyrase relative to that of wild-type gyrase for catalytic inhibition and poisoning assays, respectively. The CC3 value was used to minimize bias due to multiple cleavage events in a single DNA molecule.Open in a separate windowFIG. 1.Structures of the diones and fluoroquinolones used in this study. The structures of 8-methoxy 2,4-dione (UING5-207), 8-methyl 2,4-dione (UIJR1-048), 5-methoxy 1,3-dione (UIJR1-100), 8-methoxy fluoroquinolone (UING5-249), and ciprofloxacin are shown.Open in a separate windowFIG. 2.Representative results of supercoiling assays. The supercoiling assay was conducted using wild-type gyrase and GyrA S83W gyrase to examine the effect of either ciprofloxacin (A) or 8-methoxy 2,4-dione (B). wt, wild-type gyrase; S83W, GyrA S83W gyrase.

TABLE 1.

Inhibition of the catalytic activities of E. coli gyrase and Topo IV
CompoundMedian IC50 (μM) ± ADb
Wild-type gyraseGyrA S83W gyraseGyrA G81C gyraseGyrA A67S gyraseTopo IV
Ciprofloxacin0.45 ± 0.004101 ± 1.9 (224)a28 ± 7.0 (62)1.0 ± 0.15 (2.2)15.9 ± 1.9
8-Methoxy fluoroquinolone0.16 ± 0.011.4 ± 0.1 (8.8)1.2 ± 0.1 (7.5)0.12 ± 0.02 (0.75)1.7 ± 0.1
8-Methoxy 2,4-dione2.8 ± 0.15.9 ± 0.9 (2.1)4.3 ± 0.4 (1.5)2.4 ± 0.1 (0.86)16 ± 0.4
8-Methyl 2,4-dione0.95 ± 0.153.8 ± 0.6 (4.0)1.7 ± 0.2 (1.8)1.2 ± 0.3 (1.3)6.0 ± 0.2
5-Methoxy 1,3-dione11 ± 1.267 ± 4.5 (6.1)31 ± 2.9 (2.8)19 ± 5.0 (1.7)163 ± 7.0
Open in a separate windowaThe ratio of the IC50 for a mutant gyrase to that for the wild-type gyrase is shown in parentheses.bAD, absolute deviation.

TABLE 2.

Poisoning of E. coli gyrase and Topo IV
CompoundMedian CC3 (μM) ± AD
Wild-type gyraseGyrA S83W gyraseGyrA G81C gyraseGyrA A67S gyraseTopo IV
Ciprofloxacin0.097 ± 0.00339 ± 5.0 (402)a3.2 ± 0.7 (33)0.28 ± 0.05 (2.9)14.2 ± 3.3
8-Methoxy fluoroquinolone0.023 ± 0.0010.15 ± 0.01 (6.5)0.30 ± 0.02 (13)0.035 ± 0.004 (1.5)1.4 ± 0.2
8-Methoxy 2,4-dione0.52 ± 0.041.1 ± 0.22 (2.1)1.2 ± 0.04 (2.3)0.88 ± 0.1 (1.7)11.1 ± 1.9
8-Methyl 2,4-dione0.13 ± 0.020.73 ± 0.02 (5.6)0.41 ± 0.02 (3.1)0.33 ± 0.05 (2.5)3.0 ± 0.2
5-Methoxy 1,3-dione3.8 ± 0.213 ± 2.2 (3.4)8.9 ± 1.6 (2.3)5.8 ± 1.0 (1.5)108 ± 8.0
Open in a separate windowaThe ratio of the CC3 value for a mutant gyrase to that for the wild-type gyrase is shown in parentheses.Both assays showed that each methoxy-substituted compound, as well as the 8-methyl 2,4-dione, exhibited greater activity against mutant gyrase relative to wild-type gyrase than did ciprofloxacin (Tables (Tables11 and and2).2). This increased relative activity correlated well with that observed in vivo (Fig. (Fig.3)3) (12). Methoxy-substituted diones and the 8-methyl 2,4-dione were more effective against mutant gyrases relative to wild-type gyrase than the 8-methoxy fluoroquinolone, although the absolute IC50s for the 8-methoxy fluoroquinolone and the 8-methyl 2,4-dione were lower than those for the methoxy-substituted diones. The 8-methoxy 2,4-dione exhibited the highest activity against mutant gyrases relative to wild-type gyrase. The absolute IC50 of the 5-methoxy 1,3-dione was the highest among the compounds tested (Table (Table1).1). However, it was more effective than ciprofloxacin against GyrA S83W gyrase. The IC50 of ciprofloxacin against either GyrA S83W gyrase or GyrA G81C gyrase was higher than that against Topo IV (Table (Table1);1); consequently, the three methoxy-substituted compounds and the 8-methyl 2,4-dione were more effective against mutant gyrases, including GyrA S83W gyrase, than Topo IV, the secondary target of these compounds (Tables (Tables11 and and2).2). These results suggest that the elevated activity of the 8-methoxy 2,4-dione against gyrase mutants relative to wild-type E. coli (12) was due to its effectiveness against mutant gyrase and not to a target switch from gyrase to Topo IV. In addition, these studies are the first to report 1,3-dione activity against mutant gyrases. Further studies are necessary to identify the mutations that confer resistance to these diones on wild-type and quinolone-resistant bacteria. It will be interesting to investigate how the presence of a quinolone resistance-producing substitution, such as S83W in the GyrA protein, would affect the frequency and types of dione resistance mutations.Open in a separate windowFIG. 3.Comparison of in vitro and in vivo activities against mutant gyrase. Activities of the 8-methoxy 2,4-dione (A) and its cognate fluoroquinolone (B) against mutants measured in vivo (MIC values are from reference 12), in the supercoiling assay (Table (Table1),1), and in the DNA cleavage assay (Table (Table2)2) are expressed as the ratio of the MIC value, the IC50, or the CC3 value for each mutant gyrase to that of wild-type gyrase and are shown as solid, open, and striped bars, respectively. E. coli strains or purified gyrases containing the indicated mutation were used in the assays. As reported in reference 12, the absolute MIC values for the wild-type E. coli strain were 2.5 and 0.004 μg of the 8-methoxy 2,4-dione and the 8-methoxy fluoroquinolone/ml, respectively.Each quinolone class drug typically has either gyrase or Topo IV, but not both, as the primary target in vivo. DNA gyrase is often the primary target of quinolones in Gram-negative bacteria (20, 21), whereas Topo IV is frequently the primary target in Gram-positive organisms (9). However, the preferred target can be switched by changes in quinolone structure (11, 25). It is unclear what determines the primary target of a quinolone. Some newer fluoroquinolones, such as moxifloxacin and other 8-methoxy fluoroquinolones, target gyrase and Topo IV of some organisms with nearly equipotent activity, and these dual-targeting fluoroquinolones seem to reduce the emergence of drug-resistant mutants (4, 19, 23, 26, 31). The quinolone sensitivity of gyrase and Topo IV, estimated by in vitro catalytic assays, is likely to be among the key factors that determine the primary target of a quinolone in vivo. For example, the inhibitory effects of ciprofloxacin on the supercoiling activity of gyrase and the decatenating activity of Topo IV showed that purified E. coli gyrase, the primary target of ciprofloxacin in E. coli, was more sensitive to ciprofloxacin than is purified E. coli Topo IV, whereas purified S. aureus Topo IV, the primary target of ciprofloxacin in S. aureus, was more sensitive to ciprofloxacin than purified S. aureus gyrase (Table (Table3;3; the high-salt conditions likely to be relevant with S. aureus topoisomerases [3, 16] precluded the use of CC3 as a comparator in Table Table33).

TABLE 3.

Inhibition of the catalytic activities of E. coli and S. aureus topoisomerases
CompoundMedian IC50 (μM) ± AD for E. coli
SelectivityaMedian IC50 (μM) ± AD for S. aureus
Selectivitya
GyraseTopo IVGyraseTopo IV
Ciprofloxacin0.45 ± 0.00415.9 ± 1.935.331.5 ± 2.58.8 ± 1.43.6
8-Methoxy fluoroquinolone0.16 ± 0.011.7 ± 0.110.61.1 ± 0.10.5 ± 0.032.2
8-Methoxy 2,4-dione2.8 ± 0.116 ± 0.45.72.3 ± 0.15.5 ± 0.22.4
8-Methyl 2,4-dione0.95 ± 0.156.0 ± 0.26.31.6 ± 0.30.88 ± 0.071.8
5-Methoxy 1,3-dione11 ± 1.2163 ± 714.836.5 ± 0.548 ± 81.3
Open in a separate windowaSelectivity is defined as the ratio of the IC50 for the secondary target to that for the primary target. A selectivity value of 1 indicates perfect dual targeting.Target selection by diones has not been extensively studied. Only one report describes the differential interaction of an 8-methyl 2,4-dione with Streptococcus pneumoniae gyrase and Topo IV (27). Thus, characterization and comparison of the effectiveness of assorted dione structures against the catalytic activities of S. aureus and E. coli topoisomerases in vitro should provide useful insight into target selection by a dione in vivo. These studies will also further our understanding of how the structure of a dione (e.g., 8-methyl versus 8-methoxy or 1,3- versus 2,4-) might influence target selection. Although absolute potency varied, the 8-methoxy fluoroquinolone and three diones inhibited the activities of S. aureus gyrase and S. aureus Topo IV with similar degrees of efficacy (Table (Table3).3). The 8-methoxy fluoroquinolone and the 8-methyl 2,4-dione were slightly more effective against S. aureus Topo IV, while the methoxy-substituted diones were slightly more effective against S. aureus gyrase. Thus, the diones, as well as the 8-methoxy fluoroquinolone, might act as dual-targeting drugs against S. aureus while preferentially targeting gyrase over Topo IV with E. coli. Additional structure-function studies are required to determine whether dual targeting of S. aureus is inherently general to diones or if specific substituents in the core structures are required (e.g., the 8-methoxy group on fluoroquinolones).In conclusion, an 8-methoxy 2,4-dione, an 8-methyl 2,4-dione, and a 5-methoxy 1,3-dione exhibited greater in vitro activities against quinolone-resistant mutant gyrases relative to wild-type gyrase than did ciprofloxacin or an 8-methoxy fluoroquinolone. E. coli Topo IV was less sensitive to these diones than any of the quinolone-resistant mutant gyrases; thus, the activities of these quinolone class antimicrobial agents against gyrase mutants relative to wild-type E. coli were due to their effectiveness against mutant gyrases and not to a target switch from gyrase to Topo IV. In addition, the diones used in this study inhibited the catalytic activities of S. aureus gyrase and S. aureus Topo IV at similar effectiveness levels, indicating that some diones might function as dual-targeting agents against S. aureus.  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