Mechanism of quinolone action and resistance

Katie J Aldred, Robert J Kerns, Neil Osheroff, Katie J Aldred, Robert J Kerns, Neil Osheroff

Abstract

Quinolones are one of the most commonly prescribed classes of antibacterials in the world and are used to treat a variety of bacterial infections in humans. Because of the wide use (and overuse) of these drugs, the number of quinolone-resistant bacterial strains has been growing steadily since the 1990s. As is the case with other antibacterial agents, the rise in quinolone resistance threatens the clinical utility of this important drug class. Quinolones act by converting their targets, gyrase and topoisomerase IV, into toxic enzymes that fragment the bacterial chromosome. This review describes the development of the quinolones as antibacterials, the structure and function of gyrase and topoisomerase IV, and the mechanistic basis for quinolone action against their enzyme targets. It will then discuss the following three mechanisms that decrease the sensitivity of bacterial cells to quinolones. Target-mediated resistance is the most common and clinically significant form of resistance. It is caused by specific mutations in gyrase and topoisomerase IV that weaken interactions between quinolones and these enzymes. Plasmid-mediated resistance results from extrachromosomal elements that encode proteins that disrupt quinolone-enzyme interactions, alter drug metabolism, or increase quinolone efflux. Chromosome-mediated resistance results from the underexpression of porins or the overexpression of cellular efflux pumps, both of which decrease cellular concentrations of quinolones. Finally, this review will discuss recent advancements in our understanding of how quinolones interact with gyrase and topoisomerase IV and how mutations in these enzymes cause resistance. These last findings suggest approaches to designing new drugs that display improved activity against resistant strains.

Figures

Figure 1
Figure 1
Quinolone structures. Nalidixic acid and oxolinic acid were the first-generation quinolones that were used most often in the clinic. Norfloxacin, ciprofloxacin, and ofloxacin are the most relevant second-generation quinolones. Levofloxacin (the levorotary isomer of ofloxacin), sparfloxacin, and moxifloxacin are newer-generation quinolones.
Figure 2
Figure 2
Domain structures of type II topoisomerases. Gyrase and topoisomerase IV are heterotetrameric enzymes consisting of two A subunits and two B subunits. The A subunits (blue and red; GyrA in gyrase and ParC and GrlA in Gram-negative and Gram-positive topoisomerase IV, respectively) contain the active site tyrosine residue that covalently attaches to the newly generated 5′-termini of DNA during the cleavage reaction. The C-terminal domains (CTDs; red) of the A subunits are variable and allow gyrase to introduce negative supercoils into DNA. The B subunits (green; GyrB in gyrase and ParE and GrlB in Gram-negative and Gram-positive topoisomerase IV, respectively) contain the ATPase and TOPRIM domains, the latter of which binds the catalytic divalent metal ions essential for enzyme activity. Human topoisomerase IIα is homologous to the bacterial type II enzymes. However, during the course of evolution, the A and B subunits fused into a single polypeptide chain. Therefore, eukaryotic type II topoisomerases function as homodimers. A representation of the three-dimensional structure of type II topoisomerases is shown at the bottom.
Figure 3
Figure 3
Crystal structure of a moxifloxacin-stabilized Acinetobacter baumannii topoisomerase IV–DNA cleavage complex. The catalytic core of the enzyme is shown. Moxifloxacin is colored red; the topoisomerase IV A and B subunits are colored blue and green, respectively, and DNA is colored yellow. The top panel is a top view of the cleavage complex showing two quinolone molecules intercalating 4 bp apart at the sites of DNA cleavage. The bottom panel is a front view (rotated by 90° from the top view) of the cleavage complex. Protein Data Bank accession 2XKK was visualized using Discovery Studio 3.5 Visualizer (Accelrys Software Inc.). Adapted from ref (11).
Figure 4
Figure 4
Bacterial type II topoisomerases are essential but potentially toxic enzymes. The balance between enzyme-mediated DNA cleavage and religation is critical for cell survival. If the level of gyrase-mediated DNA cleavage decreases, rates of DNA replication slow and impair cell growth (left). If the level of topoisomerase IV-mediated DNA cleavage decreases, cells are not able to untangle daughter chromosomes and ultimately die of mitotic failure (left). If the level of gyrase- or topoisomerase IV-mediated DNA cleavage becomes too high (right), the actions of DNA tracking systems can convert these transient complexes to permanent double-stranded breaks. The resulting DNA breaks initiate the SOS response and other DNA repair pathways and can lead to cell death.
Figure 5
Figure 5
Quinolone–topoisomerase binding is facilitated through a water–metal ion bridge. The top panel shows the crystal structure of a moxifloxacin-stabilized A. baumannii topoisomerase IV–DNA cleavage complex. Moxifloxacin is colored black, and the noncatalytic Mg2+ ion that is chelated by the C3/C4 keto acid of the quinolone and participates in the bridge interaction is colored green. The four water molecules that fill out the coordination sphere of the Mg2+ ion are colored blue. The backbone of selected portions of the protein amino acid chain is colored yellow. The side chains of the serine and acidic residues that form hydrogen bonds with the water molecules in the water–metal ion bridge are colored red. In A. baumannii topoisomerase IV, these residues are Ser84 and Glu88, respectively. For clarity, DNA has been omitted from the picture. Protein Data Bank accession 2XKK was visualized using Discovery Studio 3.5 Visualizer (Accelrys Software Inc.). Adapted from ref (11). The middle panel shows a simplified diagram of the water–metal ion bridge. Only interactions with the protein (and not DNA) are shown. Ciprofloxacin (a representative quinolone) is colored black, and the noncatalytic Mg2+, water molecules, and coordinating serine and acidic residues are colored as described above. Blue dashed lines indicate the octahedral coordination sphere of the divalent metal ion interacting with four water molecules and the C3/C4 keto acid of the quinolone. The red dashed lines represent hydrogen bonds between the serine side chain hydroxyl group or the acidic residue side chain carboxyl group and the water molecules. Adapted from ref (13). The bottom panel shows a sequence alignment of the A subunits highlighting the conserved serine and acidic residues (red) that coordinate the water–metal ion bridge. Sequences of A. baumannii (Ab), Bacillus anthracis (Ba), E. coli (Ec), Staphylococcus aureus (Sa), and S. pneumoniae (Sp) gyrase (GyrA) and topoisomerase IV (ParC/GrlA) are shown. The homologous regions of human topoisomerase IIα (hTIIα) and IIβ (hTIIβ), which lack the residues necessary to coordinate the water–metal ion bridge interaction, are shown for comparison.
Figure 6
Figure 6
Mechanisms of quinolone resistance. (1) Target-mediated resistance. Mutations in gyrase and topoisomerase IV weaken quinolone–enzyme interactions. (2) Plasmid-mediated resistance. (2a) Qnr proteins (yellow) decrease topoisomerase–DNA binding and protect enzyme–DNA complexes from quinolones. (2b) Aac(6′)-Ib-cr is an aminoglycoside acetyltransferase that acetylates the free nitrogen on the C7 ring of ciprofloxacin and norfloxacin, decreasing their effectiveness. (2c) Plasmid-encoded efflux pumps decrease the concentration of quinolones in the cell. (3) Chromosome-mediated resistance. (3a) Underexpression of porins in Gram-negative species decreases drug uptake. (3b) Overexpression of chromosome-encoded efflux pumps decreases drug retention in the cell.
Figure 7
Figure 7
Roles of substituents and core elements of quinolones and quinazolinediones that mediate drug activity against bacterial and human type II topoisomerases. Results are based on studies with B. anthracis topoisomerase IV and human topoisomerase IIα. For quinolones, the binding of clinically relevant drugs to topoisomerase IV is mediated primarily through the water–metal ion bridge. The binding of quinolones that overcome resistance is mediated primarily by the substituent at C7. Binding of quinolones to human topoisomerase IIα also is mediated by the C7 substituent. The group at C8 affects the ability of quinolones to act against the human type II enzyme but is not required for drug binding. For quinazolinediones, interactions between drugs and topoisomerase IV (wild-type and resistant) are mediated through the C7 substituent. The effects of the C7 and C8 substituents on quinazolinedione activity against topoisomerase IIα are the same as described for the quinolones. The N3 amino group plays a role in the binding of quinazolinediones to the human enzyme. Adapted from ref (53).

References

    1. Emmerson A. M.; Jones A. M. (2003) The quinolones: Decades of development and use. J. Antimicrob. Chemother. 51(Suppl. 1), 13–20.
    1. Mitscher L. A. (2005) Bacterial topoisomerase inhibitors: Quinolone and pyridone antibacterial agents. Chem. Rev. 105, 559–592.
    1. Linder J. A.; Huang E. S.; Steinman M. A.; Gonzales R.; Stafford R. S. (2005) Fluoroquinolone prescribing in the United States: 1995 to 2002. Am. J. Med. 118, 259–268.
    1. Andriole V. T. (2005) The quinolones: Past, present, and future. Clin. Infect. Dis. 41(Suppl. 2), S113–S119.
    1. Drlica K.; Hiasa H.; Kerns R.; Malik M.; Mustaev A.; Zhao X. (2009) Quinolones: Action and resistance updated. Curr. Top. Med. Chem. 9, 981–998.
    1. Dalhoff A. (2012) Resistance surveillance studies: A multifaceted problem—the fluoroquinolone example. Infection 40, 239–262.
    1. Hooper D. C. (1999) Mode of action of fluoroquinolones. Drugs 58(Suppl. 2), 6–10.
    1. Hooper D. C. (2001) Mechanisms of action of antimicrobials: Focus on fluoroquinolones. Clin. Infect. Dis. 32(Suppl. 1), S9–S15.
    1. Anderson V. E.; Osheroff N. (2001) Type II topoisomerases as targets for quinolone antibacterials: Turning Dr. Jekyll into Mr. Hyde. Curr. Pharm. Des. 7, 337–353.
    1. Drlica K.; Malik M.; Kerns R. J.; Zhao X. (2008) Quinolone-mediated bacterial death. Antimicrob. Agents Chemother. 52, 385–392.
    1. Wohlkonig A.; Chan P. F.; Fosberry A. P.; Homes P.; Huang J.; Kranz M.; Leydon V. R.; Miles T. J.; Pearson N. D.; Perera R. L.; Shillings A. J.; Gwynn M. N.; Bax B. D. (2010) Structural basis of quinolone inhibition of type IIA topoisomerases and target-mediated resistance. Nat. Struct. Mol. Biol. 17, 1152–1153.
    1. Aldred K. J.; McPherson S. A.; Wang P.; Kerns R. J.; Graves D. E.; Turnbough C. L. Jr.; Osheroff N. (2012) Drug interactions with Bacillus anthracis topoisomerase IV: Biochemical basis for quinolone action and resistance. Biochemistry 51, 370–381.
    1. Aldred K. J.; McPherson S. A.; Turnbough C. L. Jr.; Kerns R. J.; Osheroff N. (2013) Topoisomerase IV-quinolone interactions are mediated through a water-metal ion bridge: Mechanistic basis of quinolone resistance. Nucleic Acids Res. 41, 4628–4639.
    1. Hooper D. C. (2001) Emerging mechanisms of fluoroquinolone resistance. Emerging Infect. Dis. 7, 337–341.
    1. Guan X.; Xue X.; Liu Y.; Wang J.; Wang Y.; Wang K.; Jiang H.; Zhang L.; Yang B.; Wang N.; Pan L. (2013) Plasmid-mediated quinolone resistance-current knowledge and future perspectives. J. Int. Med. Res. 41, 20–30.
    1. Lesher G. Y.; Froelich E. J.; Gruett M. D.; Bailey J. H.; Brundage R. P. (1962) 1,8-Naphthyridine derivatives. A new class of chemotherapeutic agents. J. Med. Pharm. Chem. 91, 1063–1065.
    1. Stein G. E. (1988) The 4-quinolone antibiotics: Past, present, and future. Pharmacotherapy 8, 301–314.
    1. Anderson V. R.; Perry C. M. (2008) Levofloxacin: A review of its use as a high-dose, short-course treatment for bacterial infection. Drugs 68, 535–565.
    1. Noel G. J. (2009) A review of levofloxacin for the treatment of bacterial infections. Clin. Med.: Ther. 1, 433–458.
    1. World Health Organization (2013) Global Tuberculosis Control (accessed February 25, 2014).
    1. Levine C.; Hiasa H.; Marians K. J. (1998) DNA gyrase and topoisomerase IV: Biochemical activities, physiological roles during chromosome replication, and drug sensitivities. Biochim. Biophys. Acta 1400, 29–43.
    1. Champoux J. J. (2001) DNA topoisomerases: Structure, function, and mechanism. Annu. Rev. Biochem. 70, 369–413.
    1. Forterre P.; Gribaldo S.; Gadelle D.; Serre M. C. (2007) Origin and evolution of DNA topoisomerases. Biochimie 89, 427–446.
    1. Forterre P.; Gadelle D. (2009) Phylogenomics of DNA topoisomerases: Their origin and putative roles in the emergence of modern organisms. Nucleic Acids Res. 37, 679–692.
    1. Pommier Y.; Leo E.; Zhang H.; Marchand C. (2010) DNA topoisomerases and their poisoning by anticancer and antibacterial drugs. Chem. Biol. 17, 421–433.
    1. Gentry A. C., and Osheroff N. (2013) DNA topoisomerases: Type II. In Encyclopedia of Biological Chemistry, pp 163–168, Elsevier Inc., Amsterdam.
    1. Schmidt B. H.; Burgin A. B.; Deweese J. E.; Osheroff N.; Berger J. M. (2010) A novel and unified two-metal mechanism for DNA cleavage by type II and IA topoisomerases. Nature 465, 641–644.
    1. Deweese J. E.; Osheroff N. (2010) The use of divalent metal ions by type II topoisomerases. Metallomics 2, 450–459.
    1. Pitts S. L.; Liou G. F.; Mitchenall L. A.; Burgin A. B.; Maxwell A.; Neuman K. C.; Osheroff N. (2011) Use of divalent metal ions in the DNA cleavage reaction of topoisomerase IV. Nucleic Acids Res. 39, 4808–4817.
    1. Zechiedrich E. L.; Khodursky A. B.; Bachellier S.; Schneider R.; Chen D.; Lilley D. M.; Cozzarelli N. R. (2000) Roles of topoisomerases in maintaining steady-state DNA supercoiling in Escherichia coli. J. Biol. Chem. 275, 8103–8113.
    1. Deibler R. W.; Rahmati S.; Zechiedrich E. L. (2001) Topoisomerase IV, alone, unknots DNA in E. coli. Genes Dev. 15, 748–761.
    1. Corbett K. D.; Shultzaberger R. K.; Berger J. M. (2004) The C-terminal domain of DNA gyrase A adopts a DNA-bending β-pinwheel fold. Proc. Natl. Acad. Sci. U.S.A. 101, 7293–7298.
    1. Tretter E. M.; Berger J. M. (2012) Mechanisms for defining supercoiling set point of DNA gyrase orthologs: I. A nonconserved acidic C-terminal tail modulates Escherichia coli gyrase activity. J. Biol. Chem. 287, 18636–18644.
    1. Tretter E. M.; Berger J. M. (2012) Mechanisms for defining supercoiling set point of DNA gyrase orthologs: II. The shape of the GyrA subunit C-terminal domain (CTD) is not a sole determinant for controlling supercoiling efficiency. J. Biol. Chem. 287, 18645–18654.
    1. Deweese J. E.; Osheroff N. (2009) The DNA cleavage reaction of topoisomerase II: Wolf in sheep’s clothing. Nucleic Acids Res. 37, 738–749.
    1. Deweese J. E.; Osheroff M. A.; Osheroff N. (2009) DNA topology and topoisomerases: Teaching a “knotty” subject. Biochem. Mol. Biol. Educ. 37, 2–10.
    1. Kreuzer K. N.; Cozzarelli N. R. (1979) Escherichia coli mutants thermosensitive for deoxyribonucleic acid gyrase subunit A: Effects on deoxyribonucleic acid replication, transcription, and bacteriophage growth. J. Bacteriol. 140, 424–435.
    1. Laponogov I.; Sohi M. K.; Veselkov D. A.; Pan X. S.; Sawhney R.; Thompson A. W.; McAuley K. E.; Fisher L. M.; Sanderson M. R. (2009) Structural insight into the quinolone-DNA cleavage complex of type IIA topoisomerases. Nat. Struct. Mol. Biol. 16, 667–669.
    1. Laponogov I.; Pan X. S.; Veselkov D. A.; McAuley K. E.; Fisher L. M.; Sanderson M. R. (2010) Structural basis of gate-DNA breakage and resealing by type II topoisomerases. PLoS One 5, e11338.
    1. Bax B. D.; Chan P. F.; Eggleston D. S.; Fosberry A.; Gentry D. R.; Gorrec F.; Giordano I.; Hann M. M.; Hennessy A.; Hibbs M.; Huang J.; Jones E.; Jones J.; Brown K. K.; Lewis C. J.; May E. W.; Saunders M. R.; Singh O.; Spitzfaden C. E.; Shen C.; Shillings A.; Theobald A. J.; Wohlkonig A.; Pearson N. D.; Gwynn M. N. (2010) Type IIA topoisomerase inhibition by a new class of antibacterial agents. Nature 466, 935–940.
    1. Sugino A.; Peebles C. L.; Kreuzer K. N.; Cozzarelli N. R. (1977) Mechanism of action of nalidixic acid: Purification of Escherichia coli nalA gene product and its relationship to DNA gyrase and a novel nicking-closing enzyme. Proc. Natl. Acad. Sci. U.S.A. 74, 4767–4771.
    1. Gellert M.; Mizuuchi K.; O’Dea M. H.; Itoh T.; Tomizawa J. (1977) Nalidixic acid resistance: A second genetic character involved in DNA gyrase activity. Proc. Natl. Acad. Sci. U.S.A. 74, 4772–4776.
    1. Kato J.; Nishimura Y.; Imamura R.; Niki H.; Hiraga S.; Suzuki H. (1990) New topoisomerase essential for chromosome segregation in E. coli. Cell 63, 393–404.
    1. Khodursky A. B.; Zechiedrich E. L.; Cozzarelli N. R. (1995) Topoisomerase IV is a target of quinolones in Escherichia coli. Proc. Natl. Acad. Sci. U.S.A. 92, 11801–11805.
    1. Aedo S.; Tse-Dinh Y. C. (2012) Isolation and quantitation of topoisomerase complexes accumulated on Escherichia coli chromosomal DNA. Antimicrob. Agents Chemother. 56, 5458–5464.
    1. Pan X.-S.; Ambler J.; Mehtar S.; Fisher L. M. (1996) Involvement of topoisomerase IV and DNA gyrase as ciprofloxacin targets in Streptococcus pneumoniae. Antimicrob. Agents Chemother. 40, 2321–2326.
    1. Fournier B.; Zhao X.; Lu T.; Drlica K.; Hooper D. C. (2000) Selective targeting of topoisomerase IV and DNA gyrase in Staphylococcus aureus: Different patterns of quinolone-induced inhibition of DNA synthesis. Antimicrob. Agents Chemother. 44, 2160–2165.
    1. Pan X.-S.; Fisher L. M. (1997) Targeting of DNA gyrase in Streptococcus pneumoniae by sparfloxacin: Selective targeting of gyrase or topoisomerase IV by quinolones. Antimicrob. Agents Chemother. 41, 471–474.
    1. Pan X.-S.; Fisher L. M. (1998) DNA gyrase and topoisomerase IV are dual targets of clinafloxacin action in Streptococcus pneumoniae. Antimicrob. Agents Chemother. 42, 2810–2816.
    1. Price L. B.; Vogler A.; Pearson T.; Busch J. D.; Schupp J. M.; Keim P. (2003) In vitro selection and characterization of Bacillus anthracis mutants with high-level resistance to ciprofloxacin. Antimicrob. Agents Chemother. 47, 2362–2365.
    1. Morgan-Linnell S. K.; Becnel Boyd L.; Steffen D.; Zechiedrich L. (2009) Mechanisms accounting for fluoroquinolone resistance in Escherichia coli clinical isolates. Antimicrob. Agents Chemother. 53, 235–241.
    1. Sissi C.; Perdona E.; Domenici E.; Feriani A.; Howells A. J.; Maxwell A.; Palumbo M. (2001) Ciprofloxacin affects conformational equilibria of DNA gyrase A in the presence of magnesium ions. J. Mol. Biol. 311, 195–203.
    1. Aldred K. J.; Schwanz H. A.; Li G.; McPherson S. A.; Turnbough C. L. Jr.; Kerns R. J.; Osheroff N. (2013) Overcoming target-mediated quinolone resistance in topoisomerase IV by introducing metal-ion-independent drug-enzyme interactions. ACS Chem. Biol. 8, 2660–2668.
    1. Drlica K.; Zhao X. (1997) DNA gyrase, topoisomerase IV, and the 4-quinolones. Microbiol. Mol. Biol. Rev. 61, 377–392.
    1. Li Z.; Deguchi T.; Yasuda M.; Kawamura T.; Kanematsu E.; Nishino Y.; Ishihara S.; Kawada Y. (1998) Alteration in the GyrA subunit of DNA gyrase and the ParC subunit of DNA topoisomerase IV in quinolone-resistant clinical isolates of Staphylococcus epidermidis. Antimicrob. Agents Chemother. 42, 3293–3295.
    1. Pan X. S.; Gould K. A.; Fisher L. M. (2009) Probing the differential interactions of quinazolinedione PD 0305970 and quinolones with gyrase and topoisomerase IV. Antimicrob. Agents Chemother. 53, 3822–3831.
    1. Anderson V. E.; Zaniewski R. P.; Kaczmarek F. S.; Gootz T. D.; Osheroff N. (2000) Action of quinolones against Staphylococcus aureus topoisomerase IV: Basis for DNA cleavage enhancement. Biochemistry 39, 2726–2732.
    1. Pan X. S.; Yague G.; Fisher L. M. (2001) Quinolone resistance mutations in Streptococcus pneumoniae GyrA and ParC proteins: Mechanistic insights into quinolone action from enzymatic analysis, intracellular levels, and phenotypes of wild-type and mutant proteins. Antimicrob. Agents Chemother. 45, 3140–3147.
    1. Yague G.; Morris J. E.; Pan X. S.; Gould K. A.; Fisher L. M. (2002) Cleavable-complex formation by wild-type and quinolone-resistant Streptococcus pneumoniae type II topoisomerases mediated by gemifloxacin and other fluoroquinolones. Antimicrob. Agents Chemother. 46, 413–419.
    1. Pfeiffer E. S.; Hiasa H. (2007) Determination of the primary target of a quinolone drug and the effect of quinolone resistance-conferring mutations by measuring quinolone sensitivity based on its mode of action. Antimicrob. Agents Chemother. 51, 3410–3412.
    1. Oppegard L. M.; Streck K. R.; Rosen J. D.; Schwanz H. A.; Drlica K.; Kerns R. J.; Hiasa H. (2010) Comparison of in vitro activities of fluoroquinolone-like 2,4- and 1,3-diones. Antimicrob. Agents Chemother. 54, 3011–3014.
    1. Willmott C. J.; Maxwell A. (1993) A single point mutation in the DNA gyrase A protein greatly reduces binding of fluoroquinolones to the gyrase-DNA complex. Antimicrob. Agents Chemother. 37, 126–127.
    1. Anderson V. E.; Gootz T. D.; Osheroff N. (1998) Topoisomerase IV catalysis and the mechanism of quinolone action. J. Biol. Chem. 273, 17879–17885.
    1. Anderson V. E.; Zaniewski R. P.; Kaczmarek F. S.; Gootz T. D.; Osheroff N. (1999) Quinolones inhibit DNA religation mediated by Staphylococcus aureus topoisomerase IV: Changes in drug mechanism across evolutionary boundaries. J. Biol. Chem. 274, 35927–35932.
    1. Barnard F. M.; Maxwell A. (2001) Interaction between DNA gyrase and quinolones: Effects of alanine mutations at GyrA subunit residues Ser(83) and Asp(87). Antimicrob. Agents Chemother. 45, 1994–2000.
    1. Hiasa H. (2002) The Glu-84 of the ParC subunit plays critical roles in both topoisomerase IV-quinolone and topoisomerase IV-DNA interactions. Biochemistry 41, 11779–11785.
    1. Hiramatsu K.; Igarashi M.; Morimoto Y.; Baba T.; Umekita M.; Akamatsu Y. (2012) Curing bacteria of antibiotic resistance: Reverse antibiotics, a novel class of antibiotics in nature. Int. J. Antimicrob. Agents 39, 478–485.
    1. Martinez-Martinez L.; Pascual A.; Jacoby G. A. (1998) Quinolone resistance from a transferable plasmid. Lancet 351, 797–799.
    1. Martinez-Freijo P.; Fluit A. C.; Schmitz F. J.; Grek V. S.; Verhoef J.; Jones M. E. (1998) Class I integrons in Gram-negative isolates from different European hospitals and association with decreased susceptibility to multiple antibiotic compounds. J. Antimicrob. Chemother. 42, 689–696.
    1. Wang M.; Tran J. H.; Jacoby G. A.; Zhang Y.; Wang F.; Hooper D. C. (2003) Plasmid-mediated quinolone resistance in clinical isolates of Escherichia coli from Shanghai, China. Antimicrob. Agents Chemother. 47, 2242–2248.
    1. Robicsek A.; Jacoby G. A.; Hooper D. C. (2006) The worldwide emergence of plasmid-mediated quinolone resistance. Lancet Infect. Dis. 6, 629–640.
    1. Poirel L.; Cattoir V.; Nordmann P. (2008) Is plasmid-mediated quinolone resistance a clinically significant problem?. Clin. Microbiol. Infect. 14, 295–297.
    1. Strahilevitz J.; Jacoby G. A.; Hooper D. C.; Robicsek A. (2009) Plasmid-mediated quinolone resistance: A multifaceted threat. Clin. Microbiol. Rev. 22, 664–689.
    1. Rodriguez-Martinez J. M.; Cano M. E.; Velasco C.; Martinez-Martinez L.; Pascual A. (2011) Plasmid-mediated quinolone resistance: An update. J. Infect. Chemother. 17, 149–182.
    1. Carattoli A. (2013) Plasmids and the spread of resistance. Int. J. Med. Microbiol. 303, 298–304.
    1. Tran J. H.; Jacoby G. A. (2002) Mechanism of plasmid-mediated quinolone resistance. Proc. Natl. Acad. Sci. U.S.A. 99, 5638–5642.
    1. Xiong X.; Bromley E. H.; Oelschlaeger P.; Woolfson D. N.; Spencer J. (2011) Structural insights into quinolone antibiotic resistance mediated by pentapeptide repeat proteins: Conserved surface loops direct the activity of a Qnr protein from a Gram-negative bacterium. Nucleic Acids Res. 39, 3917–3927.
    1. Sun H. I.; Jeong da U.; Lee J. H.; Wu X.; Park K. S.; Lee J. J.; Jeong B. C.; Lee S. H. (2010) A novel family (QnrAS) of plasmid-mediated quinolone resistance determinant. Int. J. Antimicrob. Agents 36, 578–579.
    1. Lahey Clinic. qnr Numbering and Sequence (Jacoby G. A., Ed.) (accessed January 9, 2014).
    1. Tran J. H.; Jacoby G. A.; Hooper D. C. (2005) Interaction of the plasmid-encoded quinolone resistance protein Qnr with Escherichia coli DNA gyrase. Antimicrob. Agents Chemother. 49, 118–125.
    1. Tran J. H.; Jacoby G. A.; Hooper D. C. (2005) Interaction of the plasmid-encoded quinolone resistance protein QnrA with Escherichia coli topoisomerase IV. Antimicrob. Agents Chemother. 49, 3050–3052.
    1. Robicsek A.; Strahilevitz J.; Jacoby G. A.; Macielag M.; Abbanat D.; Park C. H.; Bush K.; Hooper D. C. (2006) Fluoroquinolone-modifying enzyme: A new adaptation of a common aminoglycoside acetyltransferase. Nat. Med. 12, 83–88.
    1. Guillard T.; Cambau E.; Chau F.; Massias L.; de Champs C.; Fantin B. (2013) Ciprofloxacin treatment failure in a murine model of pyelonephritis due to an AAC(6′)-Ib-cr-producing Escherichia coli strain susceptible to ciprofloxacin in vitro. Antimicrob. Agents Chemother. 57, 5830–5835.
    1. Yamane K.; Wachino J.; Suzuki S.; Kimura K.; Shibata N.; Kato H.; Shibayama K.; Konda T.; Arakawa Y. (2007) New plasmid-mediated fluoroquinolone efflux pump, QepA, found in an Escherichia coli clinical isolate. Antimicrob. Agents Chemother. 51, 3354–3360.
    1. Cattoir V.; Poirel L.; Nordmann P. (2008) Plasmid-mediated quinolone resistance pump QepA2 in an Escherichia coli isolate from France. Antimicrob. Agents Chemother. 52, 3801–3804.
    1. Hansen L. H.; Sorensen S. J.; Jorgensen H. S.; Jensen L. B. (2005) The prevalence of the OqxAB multidrug efflux pump amongst olaquindox-resistant Escherichia coli in pigs. Microb. Drug Resist. 11, 378–382.
    1. Kim H. B.; Wang M.; Park C. H.; Kim E. C.; Jacoby G. A.; Hooper D. C. (2009) oqxAB encoding a multidrug efflux pump in human clinical isolates of Enterobacteriaceae. Antimicrob. Agents Chemother. 53, 3582–3584.
    1. Martinez-Martinez L.; Pascual A.; Garcia I.; Tran J.; Jacoby G. A. (2003) Interaction of plasmid and host quinolone resistance. J. Antimicrob. Chemother. 51, 1037–1039.
    1. Jacoby G. A. (2005) Mechanisms of resistance to quinolones. Clin. Infect. Dis. 41(Suppl. 2), S120–S126.
    1. Poole K. (2007) Efflux pumps as antimicrobial resistance mechanisms. Ann. Med. 39, 162–176.
    1. Goldman J. D.; White D. G.; Levy S. B. (1996) Multiple antibiotic resistance (mar) locus protects Escherichia coli from rapid cell killing by fluoroquinolones. Antimicrob. Agents Chemother. 40, 1266–1269.
    1. Singh R.; Swick M. C.; Ledesma K. R.; Yang Z.; Hu M.; Zechiedrich L.; Tam V. H. (2012) Temporal interplay between efflux pumps and target mutations in development of antibiotic resistance in Escherichia coli. Antimicrob. Agents Chemother. 56, 1680–1685.
    1. Tran T. P.; Ellsworth E. L.; Sanchez J. P.; Watson B. M.; Stier M. A.; Showalter H. D.; Domagala J. M.; Shapiro M. A.; Joannides E. T.; Gracheck S. J.; Nguyen D. Q.; Bird P.; Yip J.; Sharadendu A.; Ha C.; Ramezani S.; Wu X.; Singh R. (2007) Structure-activity relationships of 3-aminoquinazolinediones, a new class of bacterial type-2 topoisomerase (DNA gyrase and topo IV) inhibitors. Bioorg. Med. Chem. Lett. 17, 1312–1320.
    1. German N.; Malik M.; Rosen J. D.; Drlica K.; Kerns R. J. (2008) Use of gyrase resistance mutants to guide selection of 8-methoxy-quinazoline-2,4-diones. Antimicrob. Agents Chemother. 52, 3915–3921.
    1. Malik M.; Marks K. R.; Mustaev A.; Zhao X.; Chavda K.; Kerns R. J.; Drlica K. (2011) Fluoroquinolone and quinazolinedione activities against wild-type and gyrase mutant strains of Mycobacterium smegmatis. Antimicrob. Agents Chemother. 55, 2335–2343.
    1. Mustaev A.; Malik M.; Zhao X.; Kurepina N.; Luan G.; Oppegard L. M.; Hiasa H.; Marks K. R.; Kerns R. J.; Berger J. M.; Drlica K. (2014) Fluoroquinolone-gyrase-DNA complexes: Two modes of drug binding. J. Biol. Chem. DOI: 10.1074/jbc.M113.529164.

Source: PubMed

3
Abonnere