2,945
Views
173
CrossRef citations to date
0
Altmetric
Original Articles

Antibiotic Resistance in Bacteria and Its Future for Novel Antibiotic Development

&
Pages 1060-1075 | Published online: 22 May 2014

  • 1) Cohen, M., Changing patterns of infectious disease. Nature, 406, 762–767 (2000).
  • 2) Fauch, A. S., Infectious diseases: considerations for the 21st century. Clin. Infect. Dis., 32, 675–685 (2001).
  • 3) Nathan, C., Antibiotics at the crossroads. Nature, 431, 899–902 (2004).
  • 4) Leeb, M., A shot in the arm. Nature, 431, 892–893 (2004).
  • 5) Projan, S. J., New (and not so new) antibacterial targets: from where and when will the novel drugs come? Curr. Opin. Pharmacol., 2, 513–522 (2002).
  • 6) Walsh, C., “Antibiotics, Action, Origins, Resistance,” ASM Press, Washington, DC, pp. 3–49 (2003).
  • 7) Mascaretti, O. A., “Bacteria Versus Antibacterial Agents, an Integrated Approach,” ASM Press, Washington, DC, pp. 97–105 (2003).
  • 8) Walsh, C., Where will new antibiotics come from? Nat. Rev. Microbiol., 1, 65–70 (2003).
  • 9) Nikaido, H., Prevention of drug access to bacterial targes: permeability barriers and active efflux. Science, 264, 382–388 (1994).
  • 10) Spratt, B. G., and Cromie, K. D., Penicillin-binding proteins of gram-negative bacteria. Rev. Infect. Dis., 10, 699–711 (1988).
  • 11) Williams, D. H., The glycopeptide story: how to kill the deadly ‘superbugs.’ Nat. Prod. Rep., 13, 469–477 (1996).
  • 12) Ban, N., Nissen, P., Hansen, J., Moore, P. B., and Steitz, T. A., The complete atomic structure of the large ribosomal subunit at 2.4 Å resolution. Science, 289, 905–920 (2000).
  • 13) Wimberly, B. T., Brodersen, D. E., Clemons, W. M., Jr., Morgan-Warren, R. J., Carter, A. P., Vonrhein, C., Hartsch, T., and Ramakrishnan, V., Structure of the 30S ribosomal subunit. Nature, 407, 327–339 (2000).
  • 14) Schluenzen, F., Tocilj, A., Zarivach, R., Harms, J., Gluehmann, M., Janell, D., Bashan, A., Bartels, H., Agmon, I., Franceschi, F., and Yonath, A., Structure of functionally activated small ribosomal subunit at 3.3 Å resolution. Cell, 102, 615–623 (2000).
  • 15) Yusupov, M. M., Yusupova, G. Z., Baucom, A., Lieberman, K., Earnest, T. N., Cate, J. H. D., and Noller, H. F., Crystal structure of the ribosome at 5.5 Å resolution. Science, 292, 883–896 (2001).
  • 16) Carter, A. P., Clemons, W. M., Brodersen, D. E., Morgan-Warren, R. J., Wimberly, B. T., and Ramakrishnan, V., Functional insights from the structure of the 30S ribosomal subunit and its interactions with antibiotics. Nature, 407, 340–348 (2000).
  • 17) Brodersen, D. E., Clemons, W. M., Carter, A. P., Morgan-Warren, R. J., Wimberly, B. T., and Ramakrishnan, V., The structural basis for the action of the antibiotics tetracycline, pactamycin, and hygromycin B on the 30S ribosomal subunit. Cell, 103, 1143–1154 (2000).
  • 18) Schlunzen, F., Zarivach, R., Harms, J., Bashan, A., Tocilj, A., Albrecht, R., Yonath, A., and Franceschi, F., Structural basis for the interaction of antibiotics with the peptidyl transferase centre in eubacteria. Nature, 413, 814–821 (2001).
  • 19) Sherratt, D. J., Bacterial chromosome dynamics. Science, 301, 780–785 (2003).
  • 20) Maxwell, A., DNA gyrase as a drug target. Trends Microbiol., 5, 102–109 (1997).
  • 21) Lesher, G. Y., Forelich, E. D., Gruet, M. D., Bailey, J. H., and Brundage, R. P., 1,8-Naphthyridone derivatives. A new class of chemotherapeutic agents. J. Med. Pharm. Chem., 5, 1063–1068 (1962).
  • 22) Mascaretti, O. A., “Bacteria Versus Antibacterial Agents, an Integrated Approach,” ASM Press, Washington, DC, pp. 289–328 (2003).
  • 23) Pan, X., and Fisher, L. M., Targeting of DNA gyrase in Streptococcus pneumoniae by sparfloxacin: selective targeting of gyrase or topoisomerase IV by quinolones. Antimicrob. Agents Chemother., 41, 471–474 (1997).
  • 24) Ng, E. Y., Trucksis, M., and Hooper, D. C., Quinolone resistance mutations in topoisomerase IV: relationship to the flqA locus and genetic evidence that topoisomerase IV is the primary target and DNA gyrase is the secondary target of fluoroquinolones in Staphylococcus aureus. Antimicrob. Agents Chemother., 40, 1881–1888 (1996).
  • 25) Murakami, K. S., and Darst, S. A., Bacterial RNA polymerases: the whole story. Curr. Opin. Struct. Biol., 13, 31–39 (2003).
  • 26) Spratt, B. G., Resistance to antibiotics mediated by target alterations. Science, 264, 388–393 (1994).
  • 27) Campbell, E. A., Korzheva, N., Mustaev, A., Murakami, K., Nair, S., Goldfarb, A., and Darst, S. A., Structural mechanism for rifampicin inhibition of bacterial RNA polymerase. Cell, 104, 901–912 (2001).
  • 28) Stryer, L., “Biochemistry” 4th ed., W. H. Freeman and Company, New York, pp. 719–721 (1995).
  • 29) Hancock, R. E. W., and Chapple, D. S., Peptide antibiotics. Antimicrob. Agents Chemother., 43, 1317–1323 (1999).
  • 30) Baba, T., and Schneewind, O., Instruments of microbial warfare: bacteriocin synthesis, toxicity and immunity. Trends Microbiol., 6, 66–71 (1998).
  • 31) Breukink, E., Wiedemann, I., van Kraaij, C., Kuuipers, O. P., Sahl, H.-G., and de Kruijff, B., Use of the cell wall precursor lipid II by a pore-forming peptide antibiotic. Science, 286, 2361–2364 (1999).
  • 32) Wiedemann, I., Breukink, E., van Kraaij, C., Kuipers, O. P., Bierbaum, G., de Kruijff, B., and Sahl, H.-G., Specific binding of nisin to the peptidoglycan precursor lipid II combines pore formation and inhibition of cell wall biosynthesis for potent antibiotic activity. J. Biol. Chem., 276, 1772–1779 (2001).
  • 33) Hsu, S.-T. D., Breukink, E., Tishenko, E., Lutters, M. A. G., de Kuijuff, B., Kaptein, R., Bonvin, A. M. J. J., and van Nuland, N. A. J., The nisin-lipid II complex reveals a pyrophosphate cage that provides a blueprint for novel antibiotics. Nat. Struct. Mol. Biol., 11, 963–967 (2004).
  • 34) Palumbi, S. R., Humans as the world’s greatest evolutionary force. Science, 293, 1786–1790 (2001).
  • 35) Davies, J., Inactivation of antibiotics and the dissemination of resistance genes. Science, 264, 375–382 (1994).
  • 36) Walsh, C., Molecular mechanisms that confer antibacterial drug resistance. Nature, 406, 775–781 (2000).
  • 37) Livermore, D., Can better prescribing turn the tide of resistance? Nat. Rev. Microbiol., 2, 73–78 (2004).
  • 38) Travis, J., Reviving the antibiotic miracle? Science, 264, 360–362 (1994).
  • 39) Waksman, S. A., Reily, H. C., and Schatz, A., Strain specificity and production of antibiotic substances. V. Strain resistance of bacteria to antibiotic substances, especially to streptomycin. Proc. Natl. Acad. Sci. USA, 31, 157–164 (1945).
  • 40) Murray, E., Vancomycin resistant enterococci. Am. J. Med., 102, 284–293 (1997).
  • 41) Neu, H. C., The crisis in antibiotic resistance. Science, 257, 1064–1073 (1992).
  • 42) Coates, A., Hu, Y., Bax, R., and Page, C., The future challenges facing the development of new antimicrobial drugs. Nat. Rev. Drug Discov., 1, 895–910 (2002).
  • 43) Sievert, D. M., Boulton, M. L., Stoltman, G., Johnson, D., Stobierske, M. G., Downes, F. P., Somsel, P. A., Rudrik, J. T., Brown, W., Hafeez, W., Lundstrom, T., Flanagan, E., Johnson, R., and Mitchell, J., Staphylococcus aureus resistant to vancomycin: United States, 2002. Morbid. Mortal. Wkly Rep., 51, 565–567 (2002).
  • 44) Falkow, S., and Kennedy, D., Antibiotics, animals, and people—again! Science, 291, 397 (2001).
  • 45) Witte, W., Medical consequences of antibiotic use in agriculture. Science, 279, 996–997 (1998).
  • 46) Gustafson, R. H., and Bowen, R. E., Antibiotic use in animal agriculture. J. Appl. Microbiol., 83, 531–541 (1997).
  • 47) Wehener, H. C., Antibiotics in animal feed and their role in resistance development. Curr. Opin. Microbiol., 6, 439–445 (2002).
  • 48) Casewell, M., Friis, C., Marco, E., McMullin, P., and Phillips, I., The European ban on growth-promoting antibiotics and emerging consequences for human and animal health. J. Antimicrob. Chemother., 52, 159–161 (2003).
  • 49) Tsiodras, S., Gold, H. S., Sakoulas, G., Eliopoulos, G. M., Wennersten, C., Venkataraman, L., Moellering, R. C., and Ferraro, M. J., Linezolid resistance in a clinical isolate of Staphylococcus aureus. Lancet, 358, 207–208 (2001).
  • 50) Gonzales, R. D., Schreckenberger, P. C., Graham, M. B., Kelkar, S., DenBesten, K., and Quinn, J. P., Infections due to vancomycin-resistant Enterococcus faecium resistant to linezolid. Lancet, 357, 1179 (2001).
  • 51) Nakae, T., Role of membrane permeability in determining antibiotic resistance in Pseudomonas aeruginosa. Microbiol. Immunol., 39, 221–229 (1995).
  • 52) McMurry, L., Petrucci, R. E., Jr., and Levy, S. B., Active efflux of tetracycline encoded by four genetically different tetracycline resistance determinants in Escherichia coli. Proc. Natl. Acad. Sci. USA, 77, 3974–3977 (1980).
  • 53) Yoneyama, H., and Nakae, T., Cloning of the protein D2 gene of Pseudomonas aeruginosa and its functional expression in the imipenem-resistant host. FEBS Lett., 283, 177–179 (1991).
  • 54) Poole, K., Krebes, K., McNally, C., and Neshat, S., Multiple antibiotic resistance in Pseudomonas aeruginosa: evidence for involvement of an efflux operon. J. Bacteriol., 175, 7363–7372 (1993).
  • 55) Morshed, S. R. M., Lei, Y., Yoneyama, H., and Nakae, T., Expression of genes associated with antibiotic extrusion in Pseudomonas aeruginosa. Biochem. Biophys. Res. Commun., 210, 356–362 (1995).
  • 56) Yoneyama, H., Ocaktan, A., Tsuda, M., and Nakae, T., The role of mex-gene products in antibiotic extrusion in Pseudomonas aeruginosa. Biochem. Biophys. Res. Commun., 233, 611–618 (1997).
  • 57) Paulsen, I. T., Multidrug efflux pumps and resistance: regulation and evolution. Curr. Opin. Microbiol., 6, 446–451 (2003).
  • 58) Bolhuis, H., Molenaar, D., Poelarends, G., van Veen, H. W., Poolman, B., Driessen, A. J. M., and Konings, W. N., Proton motive force-driven and ATP-dependent drug extrusion systems in multidrug-resistant Lactococcus lactis. J. Bacteriol., 176, 6957–6964 (1994).
  • 59) Neyfakh, A. A., Mystery of multidrug transporters: the answer can be simple. Mol. Microbiol., 44, 1123–1130 (2002).
  • 60) Murakami, S., Nakashima, R., Yamashita, E., and Yamaguchi, A., Crystal structure of bacterial multidrug efflux transporter AcrB. Nature, 419, 587–593 (2002).
  • 61) Akama, H., Matsuura, T., Kashiwagi, S., Yoneyama, H., Narita, S., Tsukihara, T., Nakagawa, A., and Nakae, T., Crystal structure of the membrane fusion protein, MexA, of the multidrug transporter in Pseudomonas aeruginosa. J. Biol. Chem., 279, 25939–25942 (2004).
  • 62) Higgins, M. K., Bokma, E., Koronakis, E., Hughes, C., and Koronakis, V., Structure of the periplasmic component of a bacterial drug efflux pump. Proc. Natl. Acad. Sci. USA, 101, 9994–9999 (2004).
  • 63) Koronakis, V., Sharff, A., Koronakis, E., Luisi, B., and Hughes, C., Crystal structure of the bacterial membrane protein TolC central to multidrug efflux and protein export. Nature, 405, 914–919 (2000).
  • 64) Akama, H., Kanemaki, M., Yoshimura, M., Tsukihara, T., Kashiwagi, T., Yoneyama, H., Narita, S., Nakagawa, A., and Nakae, T., Crystal structure of the drug discharge outer membrane protein, OprM, of Pseudomonas aeruginosa: dual modes of membrane anchoring and occluded cavity end. J. Biol. Chem., 279, 52816–52819 (2004).
  • 65) Fernandez-Recio, J., Walas, F., Federici, L., Pratap, J. V., Bavro, V. N., Miguel, R. N., Mizuguchi, K., and Luisi, B., A model of a transmembrane drug-efflux pump from gram-negative bacteria. FEBS Lett., 578, 5–9 (2004).
  • 66) Narita, S., Eda, S., Yoshihara, E., and Nakae, T., Linkage of the efflux-pump expression level with substrate extrusion rate in the MexAB-OprM efflux pump of Pseudomonas aeruginosa. Biochem. Biophys. Res. Commun., 308, 922–926 (2003).
  • 67) Murakami, S., and Yamaguchi, A., Multidrug-exporting secondary transporters. Curr. Opin. Struct. Biol., 13, 443–452 (2003).
  • 68) Yu, E. W., Aires, J. R., and Nikaido, H., AcrB multidrug efflux pump of Escherichia coli: composite substrate-binding cavity of exceptional flexibility generates its extremely wide substrate specificity. J. Bacteriol., 185, 5657–5664 (2003).
  • 69) Yu, E. W., McDermott, G., Zgurskaya, H. I., Nikaido, H., and Koshland, D. E., Jr., Structural basis of multiple drug-binding capacity of the AcrB multidrug efflux pump. Science, 300, 976–980 (2003).
  • 70) Fisher, J., Belasco, J. G., Khosla, S., and Knowles, J. R., β-lactamase proceeds via an acyl-enzyme intermediate: interaction of the Escherichia coli RTEM enzyme with cefoxitin. Biochemistry, 19, 2895–2901 (1980).
  • 71) Knox, J. R., Moews, P. C., and Frere, J. M., Molecular evolution of bacterial beta-lactam resistance. Chem. Biol., 3, 937–947 (1996).
  • 72) Knox, J. R., Extended-spectrum and inhibitor-resistant TEM-type beta-lactamases: mutations, specificity, and three-dimensional structure. Antimicrob. Agents Chemother., 39, 2593–2601 (1995).
  • 73) Kotra, L. P., Haddad, J., and Mobashery, S., Aminoglycosides: perspectives on mechanisms of action and resistance and strategies to counter resistance. Antimicrob. Agents Chemother., 44, 3249–3256 (2000).
  • 74) Wright, G. E., Aminoglycoside-modifying enzymes. Curr. Opin. Microbiol., 2, 499–503 (1999).
  • 75) Rodriguez-Fonseca, C., Amis, R., and Carrett, R. A., Fine structure of the peptidyl transferase center on 23S-like rRNAs deduced from chemical probing of antibiotic-ribosome complexes. J. Mol. Biol., 247, 224–235 (1995).
  • 76) Moazed, D., and Noller, H. F., Chloramphenicol, erythromycin, carbamycin and vernamycin B protect overlapping sites in the peptidyl transferase region of 23S ribosomal RNA. Biochemie, 69, 879–884 (1987).
  • 77) Murray, I. A., and Shaw, W. V., O-Acetyltransferases for chloramphenicol and other natural products. Antimicrob. Agents Chemother., 41, 1–6 (1997).
  • 78) Abraham, E. P., and Chain, E., An enzyme from bacteria able to destroy penicillin. Nature, 146, 837 (1940).
  • 79) Bussiere, D., Muchmore, S. W., Dealwis, C. G., Schluckebier, G., Nienaber, V. L., Edalji, R. P., Walter, K. A., Ladror, U. S., Holzman, T. F., and Abad-Zapatero, C., Crystal structure of ErmC′, an rRNA methyltransferase which mediates antibiotic resistance in bacteria. Biochemistry, 37, 7103–7112 (1998).
  • 80) Weisblum, B., Erythromycin resistance by ribosome modification. Antimicrob. Agents Chemother., 39, 577–585 (1995).
  • 81) Skinner, R., Cundliffe, E., and Schmidt, F. J., Site of action of a ribosomal RNA methylase responsible for resistance to erythromycin and other antibiotics. J. Biol. Chem., 258, 12702–12705 (1983).
  • 82) Cetinkaya, Y., Falk, P., and Mayhall, C. G., Vancomycin-resistant enterococci. Clin. Microbiol. Rev., 13, 686–707 (2000).
  • 83) Arias, C. A., Couvalin, P., and Reynolds, P. E., vanC cluster of vancomycin-resistant Enterococcus gallinarum BM4174. Antimicrob. Agents Chemother., 44, 1660–1666 (2000).
  • 84) Bugg, T. D., Wright, G. D., Dutka-Malen, S., Arthur, M., Couvalin, P., and Walsh, C. T., Molecular basis for vancomycin resistance in Enterococcus faecium BM4147: biosynthesis of a depsipeptide peptidoglycan precursor by vancomycin resistance proteins VanH and VanA. Biochemistry, 30, 10408–10415 (1991).
  • 85) Park, I. S., Lin, C. H., and Walsh, C. T., Bacterial resistance to vancomycin: overproduction, purification, and characterization of VanC2 from Enterococcus casseliflavus as a D-Ala-D-Ser ligase. Proc. Natl. Acad. Sci. USA, 94, 10040–10044 (1997).
  • 86) Nagai, K., Davies, T. A., Jacobs, M. R., and Appelbaum, P. C., Effects of amino acid alterations in penicillin-binding proteins (PBPs) 1a, 2b, and 2x on PBP affinities of penicillin, ampicillin, amoxicillin, cefditoren, cefuroxime, cefprozil, and cefaclor in 18 clinical isolates of penicillin-susceptible, -intermediate, and -resistant pneumococci. Antimicrob. Agents Chemother., 46, 1273–1280 (2002).
  • 87) Hakenbeck, R., Grebe, T., Zahner, D., and Stock, J. B., Beta-lactam resistance in Streptococcus pneumoniae: penicillin-binding proteins and non-penicillin-binding proteins. Mol. Microbiol., 33, 673–678 (1999).
  • 88) Zighelboim, S., and Tomasz, A., Penicillin-binding proteins of multiply antibiotic-resistant South African strains of Streptococcus pneumoniae. Antimicrob. Agents Chemother., 17, 434–442 (1980).
  • 89) Hakenbeck, R., Ellerbrock, H., Briese, T., Handwerger, S., and Tomasz, A., Penicillin-binding proteins of penicillin-susceptible and -resistant pneumococci: immunological relatedness of altered proteins and changes in peptides carrying the beta-lactam binding site. Antimicrob. Agents Chemother., 30, 553–558 (1986).
  • 90) Hakenbeck, R., Mosaic genes and their role in penicillin-resistant Streptococcus pneumoniae. Electrophoresis, 19, 597–601 (1998).
  • 91) Dowson, C. G., Coffey, T. J., Kell, C., and Whiley, R. A., Evolution of penicillin resistance in Streptococcus pneumoniae: the role of Streptococcus mitis in the formation of a low affinity PBP2B in S. pneumoniae. Mol. Microbiol., 9, 635–643 (1993).
  • 92) Kuroda, M., Ohta, T., Uchiyama, I., Baba, T., Yuzawa, H., Kobayashi, I., Cui, L., Oguchi, A., Aoki, K., Nagai, Y., Lian, J., Ito, T., Kanamori, M., Matsumura, H., Maruyama, A., Murakami, H., Hosoyama, A., Mizutani-Ui, Y., Takahashi, N. K., Sawano, T., Inoue, R., Kaito, C., Sekimizu, K., Hirakawa, H., Kuhara, S., Goto, S., Yabuzaki, J., Kanehisa, M., Yamashita, A., Oshima, K., Furuya, K., Yoshino, C., Shiba, T., Hattori, M., Ogasawara, N., Hayashi, H., and Hiramatsu, K., Whole genome sequencing of methicillin-resistant Staphylococcus aureus. Lancet, 357, 1225–1240 (2001).
  • 93) Ryffel, C., Tesch, W., Birch-Machin, I., Reynolds, P. E., Barberis-Maino, L., Kayser, F. H., and Berger-Bachi, B., Sequence comparison of mecA genes isolated from methicillin-resistant Staphylococcus aureus and Staphylococcus epidermidis. Gene, 94, 137–138 (1990).
  • 94) Ubukata, K., Nonoguchi, R., Song, M. D., Matsuhashi, M., and Konno, M., Homology of mecA gene in methicillin-resistant Staphylococcus haemolyticus and Staphylococcus simulans to that of Staphylococcus aureus. Antimicrob. Agents Chemother., 34, 170–172 (1990).
  • 95) Hall, B. G., Predicting the evolution of antibiotic resistance genes. Nat. Rev. Microbiol., 2, 430–435 (2004).
  • 96) McDevitt, D., and Rosenberg, M., Exploiting genomics to discover new antibiotics. Trends Microbiol., 9, 611–617 (2001).
  • 97) Rosamond, J., and Allsop, A., Harnessing the power of the genome in the sesarch for new antibiotics. Science, 287, 1973–1976 (2000).
  • 98) Allsop, A., New antibiotic discovery, novel screens, novel targets and impact of microbial genomics. Curr. Opin. Microbiol., 1, 530–534 (1998).
  • 99) Alksne, L. E., and Projan, S. J., Bacterial virulence as a target for antimicrobial chemotherapy. Curr. Opin. Biotechnol., 11, 625–636 (2000).
  • 100) Mahan, M. J., Slauch, J. M., and Mekalanos, J. J., Selection of bacterial virulence genes that are specifically induced in host tissues. Science, 259, 686–688 (1993).
  • 101) Chiang, S. L., Mekalanos, J. J., and Holden, D. W., In vivo genetic analysis of bacterial virulence. Annu. Rev. Microbiol., 53, 129–154 (1999).
  • 102) Valdivia, R. H., and Falkow, S., Fluorescence-based isolation of bacterial genes expressed within host cells. Science, 277, 2007–2011 (1997).
  • 103) Hensel, M., Shea, J. E., Gleeson, C., Jones, M. D., Dalton, E., and Holden, D. W., Simultaneous identification of bacterial virulence genes by negative selection. Science, 269, 400–403 (1995).
  • 104) Rodriguez, E., and McDaniel, R., Combinatorial biosynthesis of antimicrobials and other natural products. Curr. Opin. Microbiol., 4, 526–534 (2001).
  • 105) McDaniel, R., Thamchaipenet, A., Gustafsson, C., Fu, H., Betlach, M., and Ashley, G., Multiple genetic modifications of the erythromycin polyketide synthase to produce a library of novel “unnatural” natural products. Proc. Natl. Acad. Sci. USA, 96, 1846–1851 (1999).
  • 106) Sutcliffe, J. A., The search for new antibiotics targeting the 50S ribozyme. ASM News, 70, 513–519 (2004).
  • 107) Hansen, J. L., Ippolito, J. A., Ban, N., Nissen, P., Moore, P. B., and Steitz, T. A., The structure of four macrolide antibiotics bound to the large ribosomal subunit. Mol. Cell, 10, 117–128 (2002).
  • 108) Knowles, D. J. C., Foloppe, N. F., Matassova, N. B., and Murchie, A. I. H., The bacterial ribosome, a promising focus for structure-based drug design. Curr. Opin. Pharmacol., 2, 501–506 (2002).
  • 109) Yuan, Z., Trias, J., and White, R. J., Deformylase as a novel antibacterial target. Drug Discovery Today, 6, 954–961 (2001).
  • 110) Tao, J., and Schimmel, P., Inhibitors of amino acyl-tRNA synthetases as novel anti-infectives. Expert. Opin. Invest. Drugs, 9, 1767–1775 (2000).
  • 111) Payne, D. J., Warren, P. V., Holmes, D. J., and Lonsdale, J. Y., Bacterial fatty acid biosynthesis: a genomics-driven target for antibacterial drug discovery. Drug Discovery Today, 6, 537–544 (2001).
  • 112) Rohdich, F., Kis, K., Bacher, A., and Eisenreich, W., The non-mevalonate pathway of isoprenoids: genes, enzymes and intermediates. Curr. Opin. Chem. Biol., 5, 535–540 (2001).
  • 113) Clements, J. M., Coignard, F., Johnson, I., Chandler, S., Palan, S., Waller, A., Wijkmans, J., and Hunter, M. G., Antibacterial activities and characterization of novel inhibitors of LpxC. Antimicrob. Agents Chemother., 46, 1793–1799 (2002).
  • 114) Renau, T. E., Leger, R., Yen, R., She, M. W., Flamme, E. M., Sangalang, J., Gannon, C. L., Chamberland, S., Lomovskaya, O., and Lee, V. J., Peptidomimetics of efflux pump inhibitors potentiate the activity of levofloxacin in Pseudomonas aeruginosa. Bioorg. Med. Chem. Lett., 12, 763–766 (2002).
  • 115) Flatman, R. H., Howells, A. J., Heide, L., Fiedler, H.-P., and Maxwell, A., Simocyclinone D8, an inhibitor of DNA gyrase with a novel mode of action. Antimicrob. Agents Chemother., 49, 1093–1100 (2005).
  • 116) Alksne, L. E., Burgio, P., Hu, W., Feld, B., Singh, M. P., Tuckman, M., Petersen, P. J., Labthavikul, P., McGlynn, M., Barbieri, L., McDonald, L., Bradford, P., Dushin, R. G., Rothstein, D., and Projan, S. J., Identification and analysis of bacterial secretion inhibitors utilizing a SecA-LacZ reporter fusion system. Antimicrob. Agents Chemother., 44, 1418–1427 (2000).
  • 117) Mazmanian, S. K., Liu, G., Jensen, E. R., Lenoy, E., and Schneewind, O., Staphylococcus aureus sortase mutants defective in the display of surface proteins and in the pathogenesis of animal infections. Proc. Natl. Acad. Sci. USA, 97, 5510–5515 (2000).
  • 118) Darwin, K. H., Ehrt, S., Gutierrez-Ramos, J.-C., Weich, N., and Nathan, C. F., The proteasome of Mycobacterium tuberculosis is required for resistance to nitric oxide. Science, 302, 1963–1966 (2003).
  • 119) McKinney, T. K., Sharma, V. K., Craig, W. A., and Archer, G. L., Persistence of Mycobacterium tuberculosis in macrophages and mice requires the glyoxylate shunt enzyme isocitrate lyase. Nature, 406, 735–738 (2000).
  • 120) Barett, J. F., and Hoch, J. A., Two-component signal transduction as a target for microbial anti-infective therapy. Antimicrob. Agents Chemother., 42, 1529–1536 (1998).
  • 121) Mayville, P., Ji, G., Beavis, R., Yang, H., Goger, M., Novick, R. P., and Muir, T. W., Structure-activity analysis of synthetic autoinducing thiolactone peptides from Stphylococcus aureus responsible for virulence. Proc. Natl. Acad. Sci. USA, 96, 1218–1223 (1999).
  • 122) Mazel, D., Pochet, S., and Marliere, P., Genetic characterization of polypeptide deformylase, a distinctive enzyme of eubacterial translation. EMBO J., 13, 914–923 (1994).
  • 123) Mazel, D., Coic, E., Blanchard, S., Saurin, W., and Marliere, P., A survey of polypeptide deformylase function throughout the eubacterial lineage. J. Mol. Biol., 266, 939–949 (1997).
  • 124) Adams, J. M., and Capecchi, M. R., N-formylmethionyl-sRNA as the initiator of protein synthesis. Proc. Natl. Acad. Sci. USA, 55, 147–155 (1966).
  • 125) Lucchini, G., and Bianchetti, R., Initiation of protein synthesis in isolated mitochondria and chloroplasts. Biochim. Biophys. Acta, 608, 54–61 (1980).
  • 126) Guillon, J. M., Mechulam, Y., Schmitter, J. M., Blanquet, S., and Fayat, G., Disruption of the gene for Met-tRNA(fMet) formyltransferase severely impairs growth of Escherichia coli. J. Bacteriol., 174, 4294–4301 (1992).
  • 127) Meinnel, T., Mechulam, Y., and Blanquet, S., Methionine as translation start signal: a review of the enzymes of the pathway in Escherichia coli. Biochimie, 75, 1061–1075 (1993).
  • 128) Apfel, C. M., Locher, H., Evers, S., Takacs, B., Hubschwerlen, C., Pirson, W., Page, M. G. P., and Keck, W., Peptide deformylase as an antibacterial drug target: target validation and resistance development. Antimicrob. Agents Chemother., 45, 1058–1064 (2001).
  • 129) Clements, J. M., Beckett, R. P., Brown, A., Catlin, G., Lobell, M., Palan, S., Thomas, W., Whittaker, M., Wood, S., Salama, S., Baker, P. J., Rodgers, H. F., Barynin, V., Rice, D. W., and Hunter, M. G., Antibiotic activity and characterization of BB-3497, a novel peptide deformylase inhibitor. Antimicrob. Agents Chemother., 45, 563–570 (2001).
  • 130) Margolis, P. S., Hackbarth, C. J., Young, D. C., Chen, W. W. D., Yuan, Z., White, R., and Trias, J., Peptide deformylase in Staphylococcus aureus: resistance to inhibition is mediated by mutations in the formyltransferase gene. Antimicrob. Agents Chemother., 44, 1825–1831 (2000).
  • 131) Rohmer, M., Knani, M., Simonin, P., Sutter, B., and Sahm, H., Isoprenoid biosynthesis in bacteria: a novel pathway for the early steps leading to isopentenyl diphosphate. Biochem. J., 295, 517–524 (1993).
  • 132) Kazuyama, T., Shimizu, T., Takahashi, S., and Seto, H., Fosmidomycin, a specific inhibitor of 1-deoxy-D-xylulose 5-phosphate reductoisomerase in the nonmevalonate pathway for terpenoid biosynthesis. Tetrahedron Lett., 39, 7913–7916 (1998).
  • 133) Hedl, M., Sutherlin, A., Wilding, E. I., Mazaulla, M., McDevitt, D., Lane, P., Burgner, J. W., 2nd, Lehnbeuter, K. R., Stauffacher, C. V., Gwynn, M. N., and Rodwell, V. W., Enteroccus faecalis acetoacetyl-coenzyme A thiolase/3-hydroxy-3-methylglutaryl-coenzyme A reductase, a dual-function protein of isopentenyl diphosphate biosynthesis. J. Bacteriol., 184, 2116–2122 (2002).
  • 134) Wildling, E. I., Brown, J. R., Bryant, A. P., Chalker, A. F., Holmes, D. J., Ingraham, K. A., Iordanescu, S., So, S., Rosenberg, M., and Gwynn, M. N., Identification, evolution, and essentiality of the mevalonate pathway for isopentenyl diphosphate biosynthesis in gram-positive cocci. J. Bacteriol., 182, 4319–4327 (2000).
  • 135) Chirala, S. S., Huang, W. Y., Jayakumar, A., Sakai, K., and Wakil, S. J., Animal fatty acid synthetase: functional mapping and cloning and expression of the domain I constituent activities. Proc. Natl. Acad. Sci. USA, 94, 5588–5593 (1997).
  • 136) Heath, R. J., White, S. W., and Rock, C. O., Lipid biosynthesis as a target for antibacterial agents. Prog. Lipid Res., 40, 467–497 (2001).
  • 137) McMurry, L. M., Oethinger, M., and Levy, S. B., Triclosan targets lipid synthesis. Nature, 394, 531–532 (1998).
  • 138) Rozwarski, D. A., Grant, G. A., Barton, D. H., Jacobs, W. R., and Sacchettini, J. C., Modification of the NADH of the isoniazid target (InhA) from Mycobacterium tuberculosis. Science, 279, 98–102 (1998).
  • 139) Qiu, X., Janson, C. A., Court, R. I., Smyth, M. G., Payne, D. J., and Abdel-Meguid, S. S., Molecular basis of triclosan activity involves a flipping loop in the active site. Protein Sci., 8, 2529–2532 (2001).
  • 140) Moche, M., Schneider, G., Edwards, P., Dehesh, K., and Lindqvist, Y., Structure of the complex between the antibiotic cerulenin and its target, beta-ketoacyl carrier protein synthase. J. Biol. Chem., 274, 6031–6034 (1999).
  • 141) Qiu, X., Janson, C. A., Smith, W. W., Head, M., Lonsdale, J., and Konstantinidis, A. K., Refined structures of beta-ketoacyl carrier protein synthase III. J. Mol. Biol., 307, 341–356 (2001).
  • 142) Strauss, E. J., and Falkow, S., Microbial pathogenesis: genomics and beyond. Science, 276, 707–712 (1997).
  • 143) Falkow, S., Molecular Koch’s postulates applied to microbial pathogenicity. Rev. Infect. Dis., 10, S274–276 (1988).
  • 144) Cheung, A. L., Eberhardt, K. J., Chung, E., Yeaman, M. R., Sullam, P. M., Ramos, M., and Bayer, A. S., Diminished virulence of a sar−/agr− mutant of Staphylococcus aureus in the rabbit model of endocarditis. J. Clin. Invest., 94, 1815–1822 (1994).
  • 145) Throup, J. P., Koretke, K. K., Bryant, A. P., Ingraham, K. A., Chalker, A. F., Ge, Y., Marta, A., Wallis, N. G., Brown, J. R., Holmes, D. J., Rosenberg, M., and Burnham, M. K., A genomic analysis of two-component signal transduction in Streptococcus pnermoniae. Mol. Microbiol., 35, 566–576 (2000).
  • 146) Throup, J. P., Zappacosta, F., Lunsford, R. D., Annan, R. S., Carr, S. A., Lonsdale, J. T., Bryant, A. P., McDevitt, D., Rosenberg, M., and Burnham, M. K., The srhSR gene pair from Staphylococcus aureus: genomic and proteomic approaches to the identification and characterization of gene function. Biochemistry, 40, 10392–10401 (2001).
  • 147) Tang, H. B., DiMango, E., Bryan, R., Gambello, M., Iglewski, B. H., Goldberg, J. B., and Prince, A., Contribution of specific Pseudomonas aeruginosa virulence factors to pathogenesis of pneumonia in a neonatal mouse model of infection. Infect. Immun., 64, 37–43 (1996).
  • 148) Uchiya, K., Barbieri, M. A., Funato, K., Shah, A. H., Stahl, P. D., and Groisman, E. A., A Salmonella virulence protein that inhibits cellular trafficking. EMBO J., 18, 3924–3933 (1999).
  • 149) Menard, R., Sansonetti, P. J., and Parsot, C., Nonpolar mutagenesis of the ipa genes defines IpaB, IpaC, and IpaD as effectors of Shigella flexneri entry into epithelial cells. J. Bacteriol., 175, 5899–5906 (1993).
  • 150) Gaillot, O., Pellegrini, E., Bregenholt, S., and Berche, P., The ClpP serine protease is essential for the intracellular parasitism and virulence of Listeria monocytogenes. Mol. Microbiol., 35, 1286–1294 (2000).
  • 151) Fischetti, V. A., Pancholi, V., and Schneewind, O., Conservation of a hexapeptide sequence in the anchor region of surface proteins from gram-positive cocci. Mol. Microbiol., 4, 1603–1605 (1990).
  • 152) Schneewind, O., Mihaylove-Petkov, D., and Model, P., Cell wall sorting signals in surface proteins of gram-positive bacteria. EMBO J., 12, 4803–4811 (1993).
  • 153) Mazmanian, S. K., Liu, G., Ton-That, H., and Schneewind, O., Staphylococcus aureus sortase, an enzyme that anchors surface proteins to the cell wall. Science, 285, 760–763 (1999).
  • 154) Ilangovan, U., Ton-That, H., Isahara, J., Schneewind, O., and Clubb, R. T., Structure of sortase, the transpeptidase that anchors proteins to the cell wall of Staphylococcus aureus. Proc. Natl. Acad. Sci. USA, 98, 6056–6061 (2001).
  • 155) Couvalin, P., and Davies, J., Antimicrobials: time to act! Editorial overview. Curr. Opin. Microbiol., 6, 425–426 (2003).
  • 156) Schlaes, D. M., Gerding, D. N., John, J. F., Jr., Craig, W. A., Bornstein, D. L., Duncan, R. A., Eckman, M. R., Farrer, W. E., Greene, W. H., Lorian, V., Levy, S., McGowan, J. E. Jr., Paul, S. M., Ruskin, J., Tenover, F. C., and Watanakunakorn, C., Guidelines for the prevention of antimicrobial resistance in hospitals. Clin. Infect. Dis., 25, 583–599 (1997).

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.