ten son man kin 2006

Upload: magicianchemist

Post on 03-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/28/2019 Ten Son Man Kin 2006

    1/14

    Molecular Microbiology (2006) doi:10.1111/j.1365-2958.2006.05063.x

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd

    Blackwell Publishing LtdOxford, UKMMIMolecular Microbiology0950-382X 2006 The Authors; Journal compilation 2006 Blackwell Publishing Ltd? 2006??Review ArticleAntibiotics and the ribosomeT.Tenson and A. Mankin

    Accepted 5 January, 2006. For correspondence. *[email protected]; Tel. (+372) 7 374 844; Fax (+372) 7 374 900; orE-mail [email protected]; Tel. (+

    1) 312 413 1406; Fax (

    +

    1) 312 4139303

    MicroReview

    Antibiotics and the ribosome

    Tanel Tenson

    1

    * and Alexander Mankin

    2

    1

    Institute of Technology, University of Tartu, Nooruse 1,

    Tartu 50411, Estonia.2

    Center for Pharmaceutical Biotechnology m/c 870,

    University of Illinois, 900 S. Ashland Ave., Chicago, IL

    60607, USA.

    Summary

    The ribosome is one of the main antibiotic targets in

    the cell. Recent years brought important insights into

    the mode of interaction of antibiotics with the ribo-some and mechanisms of antibiotic action. Ribosome

    crystallography provided a detailed view of the inter-

    actions between antibiotics and rRNA. Advances in

    biochemical techniques let us better understand

    how the binding of small organic molecules can

    interfere with functions of an enzyme four orders of

    magnitude larger than the inhibitor. These and other

    achievements paved the way for the development of

    new ribosome-targeting antibiotics, some of which

    have already entered medical practice. The recent

    progress, problems and new directions of research of

    ribosome-targeting antibiotics are discussed in thisreview.

    Introduction

    A great variety of natural, semisynthetic or synthetic anti-

    biotics inhibit the proliferation of pathogenic bacteria by

    binding to their ribosomes and interfering with translation.

    For decades, protein synthesis inhibitors have been

    among the most successful, clinically useful antibiotics.

    Recent years brought significant progress in our under-

    standing of how drugs bind to the ribosome and interfere

    with translation. However, many critical aspects of the

    drug action, adverse effects and mechanisms of resis-

    tance remain obscure, even when it comes to the oldest

    and best-studied drugs. There are even bigger gaps in

    understanding the modes of action of newer classes of

    ribosome-targeting antibiotics.

    These and related topics were the subject of discussion

    at the meeting Antibiotics and Translation that took place

    in Tartu (Estonia) in June 2005. However, this brief review,

    which was inspired by the talks and discussions that took

    place at the meeting, is not a meeting report. Rather, our

    intention is to summarize the current state of the field of

    ribosomal antibiotics and to illuminate its major trends,

    advances, and problems using examples of several well-

    studied drugs as well as new compounds.

    Interaction of antibiotics with the ribosome

    The ribosome exceeds the size of an average antibiotic

    by four orders of magnitude and presents multiple sites

    for antibiotic binding. Before the advent of ribosome crys-

    tallography, the main ways of studying drugribosome

    interactions were based on characterizing resistance

    mutations in the genes of ribosomal components and

    mapping drug binding sites by cross-linking or footprinting

    (Spahn and Prescott, 1996). These techniques provided

    extremely valuable, although low-resolution, information.

    They showed that most of the sites of antibiotic action

    coincide with the well-recognized functional centres of the

    ribosome, such as the decoding centre and tRNA bindingsites in the small subunit and the peptidyl transferase

    centre, nascent peptide exit tunnel and/or GTPase-acti-

    vating region in the large subunit. However, the molecular

    and atomic details of the interaction of drugs with their

    target were beyond the resolution power of these meth-

    ods. The nuclear magnetic resonance (NMR) studies of

    aminoglycosides complexed to synthetic RNA (Fourmy

    et al

    ., 1996) provided the first detailed view of the interac-

    tion of protein synthesis inhibitors with the RNA target.

    More recent crystallographic studies of ribosomal sub-

    units complexed with a variety of antibiotics presented an

    extended panel of fascinating, detailed views of the ribo-

    somedrug interactions at atomic resolution (Brodersen

    et al

    ., 2000; Carter et al

    ., 2000; Ogle et al

    ., 2001; Schlun-

    zen et al

    ., 2001; Hansen et al

    ., 2002; 2003; Tu et al

    .,

    2005; Yonath, 2005). Compatible with the view of the

    ribosome as an RNA machine and in excellent agreement

    with genetic and biochemical data, NMR and crystallog-

    raphy showed that interactions of drugs with rRNA

    account for most of the contacts that antibiotics establish

    with their ribosomal target.

  • 7/28/2019 Ten Son Man Kin 2006

    2/14

    2

    T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    In the small ribosomal subunit, the best-studied site of

    drug action is the decoding centre, which is targeted by

    aminoglycosides (Fig. 1). The site comprises a segment

    of an extended rRNA helix 44 that resides on the interface

    surface of the small subunit (Ogle et al

    ., 2001). The gen-

    eral organization of this rRNA element is not influenced

    significantly by contacts with other ribosome elements.

    Therefore, its structure is preserved even when isolated

    from the ribosome. This made it possible to use study

    fairly small artificial molecular complexes to understand

    how aminoglycosides interact with their rRNA target

    (Lynch et al

    ., 2003). Common aminoglycoside antibiotics

    share a universal two-ring structure of neamine that

    includes the 2

    -deoxystreptamine moiety (ring II) (Fig. 1A

    and B). Depending on the type of ring II substitution (4,5

    or 4,6), the drugs fall into neomycin/paromomycin or kan-

    amycin/gentamicin families respectively. The NMR and

    crystallographic studies showed that the universal neam-

    ine two-ring structure forms a solid lock-and-key type of

    interaction with the target site (Ogle et al

    ., 2001; Vicens

    and Westhof, 2001; 2003; Lynch et al

    ., 2003; Francois

    et al

    ., 2005). The ring I is inser ted into helix 44 by stacking

    against G1491 (here and throughout the article, the num-

    bering of Escherichia coli

    rRNA is used) and by forming

    a pseudo-base pair by establishing two hydrogen bonds

    with the universally conserved A1408. This interaction

    helps to maintain adenines A1492 and A1493 in the

    bulged-out conformation that is similar to the conformation

    in the ribosome where codonanticodon interaction

    occurs in the A site (Fig. 1E and F). The conserved ring

    II establishes hydrogen bonds with G1494 and U1495 of

    16S rRNA. The universal interactions of rings I and II of

    the neamine moiety of aminoglycosides with the decodingcentre are supplemented by opportunistic (fuzzy) inter-

    actions that drug-specific rings III and IV of aminoglyco-

    sides establish with helix 44 of 16S rRNA. These fuzzy

    contacts fine-tune placement of the drug within its RNA

    target and define the drug specificity of aminoglycoside

    resistance mutations.

    A similar trend the universal interactions of the con-

    served pharmacophore with the core binding site, which

    is fine-tuned by interactions of drug-specific chemical

    groups with the ancillary nucleotides is also seen in the

    binding mode of another ancillary important class of

    drugs, macrolides (Schlunzen et al

    ., 2001; Hansen et al

    .,

    2002; Berisio et al

    ., 2003a,b; Tu et al

    ., 2005; Wilson et al

    .,

    2005). Macrolides consist of a 14- to 16-atom lactone ring

    decorated with one or several sugars and side-chains

    (Fig. 2). These drugs bind at the upper segment of the

    nascent peptide exit tunnel in the large ribosomal subunit

    near the peptidyl transferase centre. Crystallographic

    studies of macrolides complexed to the large ribosomal

    subunit of archaeon Haloarcula marismortui

    and of bac-

    terium Deinococcus radiodurans

    showed a conserved

    interaction of the macrolide core with the ribosome. The

    lactone ring lays flat against the RNA-based exit tunnel

    wall, and the universally present monosaccharide or dis-

    accharide extensions at position 5 of the lactone protrude

    towards the peptidyl transferase centre (Fig. 2) and inter-

    act with nucleotides A2058 and A2059. Methylation of

    A2058 (and sometimes A2059) by Erm-type methyltrans-

    ferases confers resistance to all classes of macrolides

    (Madsen et al

    ., 2005). These universal contacts are

    supplemented by additional idiosyncratic interactions that

    side-chains of individual macrolides establish with rRNA.

    These latter drug-specific contacts may contribute dra-

    matically to the drug affinity for its ribosomal binding site.

    For example, binding of tylosin is enhanced owing to the

    interaction of the mycinose sugar extending from the C14

    position of the drug with the loop of helix 35 in domain II

    of 23S rRNA (Fig. 2G) (Hansen et al

    ., 2002). Not surpris-

    ingly, methylation of nucleotides in the loop of helix 35 in

    domain II of 23S rRNA can confer resistance to tylosin

    (Liu and Douthwaite, 2002), and the tylosin producer,

    Streptomyces fradiae

    , carries a dedicated methylase thathelps it to avoid suicide by methylating G748 in the loop

    of helix 35 (Douthwaite et al

    ., 2004).

    Increased affinity owing to the interaction of a side-

    chain with the loop of helix 35 is apparently also exploited

    by ketolides. These drugs represent the newest genera-

    tion of macrolides, which lack 3-cladinose but carry a

    functionally important alkyl-aryl or quinollylallyl side-chain

    that can be attached at different positions of the core

    macrolide structure (Zhong and Shortridge, 2001)

    (Fig. 3). Biochemical (RNA footprinting) and genetic

    (resistance mutations) data obtained with E. coli

    ribo-

    somes agree with interaction of the ketolide side-chain with the loop of helix 35 (Hansen et al

    ., 1999; Xiong

    et al

    ., 1999). Surprisingly, however, a direct interaction

    between the drugs and the loop of helix 35 was not

    observed in either of the explored crystalline complexes

    of ketolides with archaeal (

    H. marismortui

    ) or bacterial

    (

    D. radiodurans

    ) ribosomes (Berisio et al

    ., 2003a; Schlun-

    zen et al

    ., 2003; Tu et al

    ., 2005). Furthermore, in the

    D. radiodurans

    and H. marismortui

    structures, the position

    of the ketolide side-chain is drastically different (Fig. 2H

    and I) (Tu et al

    ., 2005; Wilson et al

    ., 2005). This discrep-

    ancy raises important questions: Which of the conflicting

    data should be used for rational design of newer mac-

    rolides? Do any of these techniques show how drugs

    interact with the ribosome in live pathogenic bacteria?

    Some insights into these problems are provided by

    comparative footprinting of ketolides on ribosomes of

    different organisms (L. Xiong and A. Mankin, unpubl.

    results). The RNA probing did not show an interaction of

    telithromycin, a clinically important ketolide, with the loop

    of helix 35 in ribosomes of D. radiodurans

    or Halobacteria

    in perfect agreement with the available crystallographic

  • 7/28/2019 Ten Son Man Kin 2006

    3/14

    Antibiotics and the ribosome

    3

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    Fig. 1.

    Interaction of aminoglycosideantibiotics with 16S rRNA in the smallribosomal subunit.AC. Chemical structures of (A) neomycin (B)

    kanamycin and (C) streptomycin.DF. Orientation of 16S rRNA residues criticalfor monitoring codonanticodon interaction: (D)in the vacant 30S subunit; (E) during recogni-tion of accurate codonanticodon pairing; (F) in

    the complexes of vacant 30S subunit with

    paromomycin. Note change in orientation ofA1492 and A1493 in (E) and (F) compared with(D). According to Ogle et al

    . (2001).

  • 7/28/2019 Ten Son Man Kin 2006

    4/14

    4

    T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    structures. In contrast, the telithromycin side-chain effi-

    ciently protected position 752 in the loop of helix 35 from

    chemical modification in ribosomes of E. coli

    and Staphy-

    lococcus aureus

    . Thus, footprinting data confirmed the

    conclusion from crystallographic studies that the precise

    placement of the same drug molecule can vary in ribo-

    somes of different organisms. This inference is in line with

    the notion that is starting to emerge from analysing the

    spectra of resistance mutations in different organisms:

    interactions of a drug with ribosomes of different species

    can be rather specific, and they can be influenced not only

    by the often conserved nucleotides of a functional sitetargeted by an antibiotic but also by less-conserved

    peripheral rRNA residues. This new paradigm is appar-

    ently going to replace the overly convenient generalization

    that knowledge about the interaction of a drug with the

    ribosomes of one model organism can be safely extrapo-

    lated to understand how the drug binds to the ribosomes

    of any pathogen.

    Crystallography and conventional biochemical tech-

    niques use fairly artificial ribosomedrug complexes that

    certainly do not reproduce the complexity and sophistica-

    tion of the real cell environment. Further limitations are

    imposed by the fact that most of the in vitro

    techniques

    usually deal with the static, non-translating ribosome

    whereas the ribosome undergoes cycles of conforma-

    tional changes during protein synthesis in vivo

    . Therefore,

    monitoring interactions of antibiotics with the translating

    ribosome in the living cell may provide unexpected

    insights into the mode of drug binding and mechanisms

    of drug action. One such example is presented by recent

    studies of the action of oxazolidinones on ribosomes of

    S. aureus

    (Colca et al

    ., 2003). Oxazolidinones are a new

    Fig. 2.

    Interaction of macrolides with the large ribosomal subunit.AC. Chemical structures of a 14-atom lactone ring erythromycin (A),and a 16-atom lactone ring josamycin (B) and tylosin (C).D. Beginning of the nascent peptide exit tunnel (Schmeing et al

    .,2002). The peptidyl transferase active site is marked by nucleotides

    A2451 and C2452 (space-fill representation). Position of a Tyr-Phe-tRNA analogue (sticks representation) illustrates orientation of thenascent peptide; 3

    -terminal A76 of tyrosyl-tRNA is shown in cyan,Tyr and Phe are yellow. Nucleotides A2058 and A2059 located in the

    upper part of the tunnel and essential for binding of macrolides are

    shown in green. Amino acid residues of proteins L4 and L22 areshown in blue and yellow respectively.E. Binding of carbomycin (a drug similar to josamycin) (Hansen et al

    .,2002). Lactone ring is red and the mycaminose, mycarose and isobu-

    tyrate residues extending to the peptidyl transferase centre are black.F. Binding or erythromycin to the ribosome (Schlunzen et al

    ., 2001).The lactone ring is red, desosamine sugar is black and cladinose isin magenta.

    G. Binding of tylosin, a 16-member ring macrolide (Hansen et al

    .,2002). Lactone is red, the mycaminosemycarose chain is black andthe mycinose sugar is magenta. Nucleotides belonging to the loop ofhelix 35 are shown in gold.H and I. difference in orientation of the alkyl-aryl side-chain (green)

    of a ketolide telithromycin in the structure of the drug bound to thelarge ribosomal subunit of D. radiodurans

    (H) (Berisio et al

    ., 2003a) orH. marismortui

    (I) (Tu et al

    ., 2005).

  • 7/28/2019 Ten Son Man Kin 2006

    5/14

    Antibiotics and the ribosome

    5

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    class of protein synthesis inhibitors that interfere with

    translation by binding to the large ribosomal subunit

    (Lin et al

    ., 1997; Kloss et al

    ., 1999; Xiong et al

    ., 2000).

    Although mutational studies clearly pointed to peptidyl

    transferase as the primary drug target, in vitro

    biochemi-

    cal data provided confusing results. In part, this was

    owing to the fact that oxazolidinones are notoriously

    sticky drugs that are prone to non-specific binding(Lin et al

    ., 1997). Therefore, attempts to cross-link oxazo-

    lidinones to the ribosome in vitro

    often resulted in the

    labelling of a number of ribosomal proteins as well as

    multiple sites in rRNA (Matassova et al

    ., 1999; A. Mankin,

    unpublished). In dramatic contrast to the in vitro

    experi-

    ments, cross-linking of photo-active oxazolidinones in the

    living S. aureus

    cells resulted in very specific labelling of

    only one ribosomal protein (L27) and only one nucleotide

    (A2602) in rRNA of the entire ribosome (Colca et al

    .,

    2003). Thus, either unique intracellular ionic conditions or

    one of the functionally significant conformational states of

    the ribosome creates a highly specific site for oxazolidi-

    none binding. This site is obviously the main site of the

    drug action, as mutations of the surrounding positions

    render cells resistant to oxazolidinones (Prystowsky

    et al

    ., 2001; Tsiodras et al

    ., 2001; Meka et al

    ., 2004).

    Interestingly, cross-linking of the drug to the protein L27

    indicates that the N-terminus of the protein reaches deep

    into the peptidyl transferase centre (Maguire et al

    ., 2005)

    a concept that is not immediately obvious from the

    available crystallographic structures of the archaeal large

    ribosomal subunit (Harms et al

    ., 2001; Moore and Steitz,

    2002). Intriguingly, a short RNA molecule of the size of

    tRNA was seen cross-linked by oxazolidinone in vivo

    (Colca et al

    ., 2003). The nature of this RNA has not been

    identified, but it may provide important clues about the

    still-obscure mechanism of drug action. Therefore, it

    appears that in vivo

    studies of the drug binding and action

    may reveal very important aspects of the mechanisms ofinhibition of protein synthesis that are beyond the reach

    of most in vitro

    techniques.

    Mechanisms of action

    Even knowing the exact location of the drug binding site

    in the ribosome and understanding the mode of antibiotic

    binding is often not enough to understand how antibiotics

    inhibit translation. The early functional studies on the

    mechanisms of antibiotic action carried out in the 1960s

    to 1970s elucidated the most general mechanisms of

    action of the major antibiotic families. However, the lack

    of appropriate experimental techniques left many ques-

    tions unanswered. What makes some classes of riboso-

    mal antibiotics bactericidal while most protein synthesis

    inhibitors are bacteriostatic? Do drugs that bind to the

    same ribosomal site have the same mode of action? What

    is the relationship between the drug structure, its affinity

    for the ribosome and the mechanism of protein synthesis

    inhibition? Some of these questions can now be

    addressed owing to advances in ribosome crystallogra-

    Fig. 3.

    Chemical structures of new ribosome-targeting antibiotics: ketolides (telithromycin and cethromycin), linezolid and tygacil.

  • 7/28/2019 Ten Son Man Kin 2006

    6/14

    6

    T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    phy, fast kinetic assays, the availability of efficient cell-free

    protein synthesis systems and other new techniques.

    Ribosome crystallography provided a clear view of the

    mechanism of action of aminoglycoside drugs that act at

    the decoding centre (Ogle et al

    ., 2001). The neamine

    core-based aminoglycoside drugs affect the ribosome

    accuracy at the initial step of aminoacyl-tRNA selection.

    Binding of the universal aminoglycoside rings I and II to

    the decoding centre changes the orientation of two uni-

    versally conserved adenine residues at positions 1492

    and 1493 of 16S rRNA (Fig. 1F). A similar conformational

    switch occurs during mRNA decoding when these two

    adenines monitor the accuracy of codonanticodon base

    pairing at the first and second anticodon positions

    (Fig. 1E). Flipping out of the adenines 1492/1493 during

    decoding signals the establishment of an accurate codon

    anticodon pairing. Reorientation of A1492/A1493 induced

    by aminoglycosides explains the miscoding effect of these

    drugs: in the presence of a bound aminoglycoside, the

    decoding centre issues an approval signal, even on the

    binding of a non-cognate tRNA, thereby increasing thefrequency of amino acid misincorporation (Ogle et al

    .,

    2001; Ogle and Ramakrishnan, 2005).

    A rather special aminoglycoside is streptomycin. It is

    chemically related to aminoglycosides of the neamine

    type (Fig. 1C), and it produces a similar effect, miscoding.

    However, in contrast to deoxystreptamine-based ami-

    noglycosides, the structure of streptomycin is built around

    the streptamine core, and it binds at a different site in the

    small ribosomal subunit, where it bridges helices 18 and

    27 (Carter et al

    ., 2000). Aminoglycosides of both types

    the neamine type and streptomycin- induce misincorpora-

    tion of amino acids into a growing polypeptide chain byscrambling communication between the decoding and

    GTPase-associated centres of the ribosome. However,

    the molecular mechanisms of their action are different

    (Pape et al

    ., 2000; Ogle et al

    ., 2001; Gromadski and Rod-

    nina, 2004). An aminoacyl-tRNA enters the ribosome in

    the complex with EF-Tu and GTP. GTP hydrolysis is

    required for EF-Tu dissociation and aminoacyl-tRNA

    accommodation in the A site. In case of the cognate

    codonanticodon interaction the induction of GTP hydrol-

    ysis is faster than in case of non-cognate or near-cognate

    codons (Ogle and Ramakrishnan, 2005). This difference

    in the rate of activation of GTP hydrolysis contributes

    significantly to the accuracy of translation. Paramomycin,

    an aminoglycoside of the neamine type, increases the rate

    of GTP hydrolysis in case of the near-cognate tRNA to be

    closer to the cognate value (Papeet al

    ., 2000). In contrast,

    streptomycin, which reduces the conformational flexibility

    of the small ribosomal subunit, lowers the rate of GTPase

    activation for cognate tRNA but increases it for near-

    cognate tRNA, bringing them to a more similar value

    (Gromadski and Rodnina, 2004). Thus, paramomycin and

    streptomycin use different strategies to eliminate the dif-

    ference in the rate of GTPase activation for cognate and

    near-cognate tRNAs, resulting in reduced accuracy of

    translation.

    Significant progress has also been achieved in under-

    standing the mode of action of macrolides. The binding of

    macrolides in the exit tunnel close to the peptidyl trans-

    ferase centre inhibits translation by blocking progression

    of the nascent peptide through the exit tunnel (Schlunzen

    et al

    ., 2001; Hansen et al

    ., 2002). Different macrolides

    leave different amounts of space available for the newly

    synthesized peptide. In cell-free translation systems, bulk-

    ier macrolides that leave less free space between the

    peptidyl transferase active site and the site of drug binding

    begin to inhibit translation when the nascent peptide is still

    very short, while drugs that leave more free space inhibit

    translation only when the nascent peptide is already fairly

    long (Tenson et al

    ., 2003). For example, josamycin, which

    reaches very closely to the peptidyl transferase centre,

    effectively inhibits formation of dipeptides and tripeptides,

    whereas the smaller erythromycin molecule, which leavesmore space for the growing peptide chain, starts inhibiting

    translation only when the nascent peptide is at least five

    amino acids long (Tenson et al

    ., 2003; Lovmar et al

    .,

    2004). Interestingly, however, the crystal structures of the

    ribosomeerythromycin complexes indicate that nascent

    peptides that are only three or four amino acids long

    should already reach the erythromycin molecule bound at

    its ribosomal site (Schlunzen et al

    ., 2001; Hansen et al.,

    2002). Therefore, the longer nascent peptides that appar-

    ently can coexist with a bound macrolide molecule in the

    ribosomal tunnel have to recoil tightly, fold back or slip by

    the drug. This notion of a nascent peptide bypassing thebound macrolide molecule was emphasized by the recent

    crystallographic structures of ribosomeerythromycin

    complexes, which showed that the drug may not

    sufficiently obstruct the tunnel opening to completely

    block progression of the newly synthesized polypeptide

    (Tu et al., 2005).

    Hindering growth of the nascent peptide by the mac-

    rolide eventually results in dissociation of peptidyl-tRNA

    from the ribosome (Menninger and Otto, 1982). Accumu-

    lation of peptidyl-tRNA in cell cytoplasm and its slow recy-

    cling can be an important factor of the inhibitory action of

    this class of antibiotics (Lovmar et al., 2004). The differ-

    ence in the dissociation rate of a macrolide antibiotic from

    the ribosome as compared with the rate of peptidyl-tRNA

    drop-off might determine the mode of drug action (Lovmar

    et al., 2004). This conclusion is illustrated by the compar-

    ison of two macrolide antibiotics josamycin and erythro-

    mycin. Both of these drugs have comparable affinity for

    the ribosome (Kd of 5.5 nM and 10.8 nM for josamycin and

    erythromycin respectively). However, their dissociation

    rates differ by approximately two orders of magnitude.

  • 7/28/2019 Ten Son Man Kin 2006

    7/14

    Antibiotics and the ribosome 7

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    Because of this difference, josamycin dissociates from the

    ribosome more than 100-fold more slowly than peptidyl-

    tRNA, whereas the dissociation rate of erythromycin is

    roughly the same as that of peptidyl-tRNA. Essentially

    none of the ribosomes that are blocked by josamycin will

    be able to synthesize a polypeptide, while half of the

    ribosomes will lose erythromycin before the dissociation

    of peptidyl-tRNA, and polypeptide elongation can then

    resume. Once the nascent polypeptide becomes long

    enough to prevent rebinding of erythromycin (more than

    eight amino acids long; Tenson et al., 2003), the synthesis

    of the polypeptide will be completed, despite the presence

    of the drug in cell cytoplasm. Therefore, while all of the

    ribosomes that carry josamycin will be inactive in protein

    synthesis, only a fraction of erythromycin-bound ribo-

    somes will be unable to complete synthesis of the initiated

    polypeptide. This prediction was experimentally con-

    firmed: in a cell-free system, josamycin could completely

    inhibit synthesis of a dodecapeptide, whereas up to 40%

    of its control amount could be synthesized, even at satu-

    rating concentrations of erythromycin (Lovmar et al.,2004). This example shows how high-resolution kinetic

    data can illuminate important aspects of drug action. It is

    likely that extending detailed kinetic measurements to

    studies of other antibiotics will provide important clues

    about the mechanisms of action of various protein synthe-

    sis inhibitors.

    Cell response to protein synthesis inhibitors

    Binding a drug to the active centres of the ribosome blocks

    translation. This eventually leads to inhibition of cell

    growth and, sometimes, to cell death. What happensbetween these two events? How does the cell respond to

    the drug while it still has the capacity to do so?

    All antibiotics induce cell stress response. Heat- or cold-

    shock responses are usually activated when bacterial

    cells are treated with inhibitors of protein synthesis (Van-

    Bogelen and Neidhardt, 1990). In addition, several other

    stress response pathways might be induced by antibiotics

    (Shapiro and Baneyx, 2002; Fischer et al., 2004). It is

    difficult, however, to correlate the type of stress response

    with the mode of drug action. It is also unclear whether

    stress response increases cell survival upon antibiotic

    treatment. Investigation of drug effects on strains lacking

    individual stress response pathways and studying the

    general response of a bacterial cell to treatment with

    protein synthesis inhibitors may therefore open new direc-

    tions for developing better drugs (Luidalepp et al., 2005).

    The DNA microarray and new proteomic technologies

    make it possible to explore cellular response to antibiotics

    at the level of transcription, mRNA accumulation or protein

    production (Evers et al., 2001; Goh et al., 2002; Sabina

    et al., 2003; Hutter et al., 2004; Tsui et al., 2004). Antibi-

    otic treatment changes the expression level of many

    genes. One of the expected cell responses to inhibition of

    translation is an attempt to restore protein synthetic

    capacity by increasing ribosome production and activating

    transcription of rRNA and ribosomal protein genes. How-

    ever, while rRNA can be rapidly produced by transcription,

    corresponding accumulation of ribosomal proteins is

    diminished as translation is inhibited. Such an imbalance

    in the manufacturing of ribosomal components should

    inevitably lead to defects in ribosome assembly in

    response to antibiotic treatment. This effect has long been

    known for chloramphenicol (Dodd et al., 1991), and more

    recently, it has been described for several other inhibitors

    of protein synthesis: aminoglycosides, macrolides and

    others (Champney and Miller, 2002a,b; Mehta and

    Champney, 2003). In the case of macrolides, the drug

    binding to assembly precursors has been suggested as a

    cause of assembly inhibition, and it was proposed that

    interference with the ribosome biogenesis accounts for

    half of the macrolide inhibitory action (Champney and

    Burdine, 1996). It has to be seen, however, whetherantibiotic-induced assembly defects are caused directly

    by the drug binding to the assembly intermediates or are

    a mere consequence of a distorted ratio in the production

    of ribosomal components.

    Interestingly, even low concentrations of antibiotics pro-

    voke notable cell responses at the level of gene expres-

    sion (Goh et al., 2002). Transcription of 510% of cellular

    genes appears to be up- or downregulated in the pres-

    ence of subinhibitory concentrations of the ribosome-

    targeting drugs. This response is highly drug-specific and

    may be directly or indirectly linked to the mechanism of

    the drug action (Tsui et al., 2004). Therefore, it can beused as a hallmark of the mechanism of action of new

    antibiotics. Surprisingly, even drugs that are not consid-

    ered to be inhibitors of bacterial protein synthesis (aniso-

    mycin, cycloheximide) appear to elicit significant changes

    in the transcription of a number of genes, suggesting that

    either these drugs do, in fact, interact with the bacterial

    ribosome or that their interactions with other cellular tar-

    gets provoke alterations in gene expression.

    Separation of the fine effects of the drug on protein

    production from their gross effect on cell growth opens

    interesting venues for investigating detailed mechanisms

    of the drug action. Most of the current models of the

    mechanisms of antibiotic action operate with the effects

    of the drugs on bulk protein synthesis. However, a number

    of indications suggest that the extent of inhibition of trans-

    lation of different genes may vary and depend on the

    nature of mRNA, tRNA or the polypeptide (Tenson and

    Mankin, 1995; Lovett and Rogers, 1996; Tenson and

    Ehrenberg, 2002). Thus, josamycin inhibits incorporation

    of phenylalanine into dipeptide much more strongly than

    incorporation of valine (Lovmar et al., 2004), whereas

  • 7/28/2019 Ten Son Man Kin 2006

    8/14

    8 T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    erythromycin and chemically unrelated pactamycin

    severely inhibit poly(A)-dependent poly(Lys) synthesis but

    have little effect on poly(U)-dependent synthesis of

    poly(Phe) (Picking et al., 1991; Dinos et al., 2004). The

    proteome- and transcriptome-profiling approaches and

    recently developed defined efficient cell-free translation

    systems may be instrumental in unravelling the specific

    effects of antibiotics on the synthesis of individual proteins

    (Antoun et al., 2004; Freistroffer et al., 1997; Tenson et al.,

    2003).

    Antibiotic resistance

    Appearance and spread of antibiotic resistance closely

    follow the introduction of new drugs to medical practice.

    When drugs act on a protein enzyme, target site muta-

    tions are one of the most common resistance mecha-

    nisms. Multiplicity of rRNA genes in most microorganisms

    slows development of this type of resistance to the

    ribosome-targeting protein synthesis inhibitors (Cundliffe,

    1990). Nevertheless, other resistance mechanisms, oftenoriginating in drug-producing organisms, eventually ren-

    der pathogens resistant to the currently used drugs. In

    addition, several important pathogens (such as Mycobac-

    terium tuberculosisor Helicobacter pylori) have only one

    or two rRNA cistrons which facilitates development of drug

    resistance due to rRNA mutations. The problem of rising

    resistance is the main driving force for the pharmaceutical

    industry to continue the effort to modify existing drugs and

    to develop newer antibiotics, and it has been a subject of

    research in many laboratories (Shah, 2005).

    Substantial progress has been achieved in understand-

    ing the mechanisms of macrolide resistance. One of themost prominent mechanisms is modification of rRNA in

    the drug target site by Erm-type methyltransferases

    whose expression can be either constitutive or inducible

    by low concentrations of macrolides (Weisblum, 1995a;

    Liu and Douthwaite, 2002; Madsen et al., 2005). Erm

    enzyme attaches one or two methyl groups to the N6

    position of adenine 2058 in 23S rRNA. Methylation of

    A2058 sterically hinders the binding of macrolides to the

    ribosome (Fig. 2) (Schlunzen et al., 2001; Tu et al., 2005).

    The A2058 methylation also affects binding of two other

    important groups of antibiotics, lincosamides and strepto-

    gramins B, producing a so-called MLSB resistance (Weis-

    blum, 1995a). Importantly, different types of Erm genes

    produce an idiosyncratic species-specific resistance pat-

    tern (Douthwaite et al., 2005).

    Until recently, Erm studies were complicated by the

    lack of convenient techniques for monitoring A2058 meth-

    ylation. While dimethylation of A2058 can be easily

    detected by primer extension, monomethylation was diffi-

    cult to monitor (Hansen et al., 2001). A considerable step

    forward was brought about by combining efficient RNA

    fragmentation techniques with mass spectrometry. This

    approach, which has been used by S. Douthwaite and

    colleagues at the University of Southern Denmark,

    helped to establish a correlation between the site and

    mechanism of rRNA methylation and the resistance pat-

    tern in several pathogens (Madsen et al., 2005). Thus,

    whereas dimethylation of A2058 in 23S rRNA renders

    cells resistant to macrolides and ketolides, monomethyla-

    tion of A2058 in Streptococcus pneumoniae and Stre-

    ptococcus pyogenesconfers resistance to erythromycin,

    but cells remain sensitive to ketolides (telithromycin)

    (Douthwaite et al., 2005). In M. tuberculosis, the relaxed

    specificity of Erm(37) results in methylation of not only

    A2058 but also neighbouring A2057 and A2059, which

    accounts for an unusual resistance profile of this patho-

    gen (Madsen et al., 2005).

    An important but still poorly understood problem is the

    mechanism of inducible Erm resistance. Early studies by

    the B. Weisblum and D. Dubnau laboratories revealed the

    basic translation attenuation scheme, which involves stall-

    ing of the ribosome during translation of a leader cistronof a bicistronic ermC operon (Dubnau, 1984; Weisblum,

    1995b). Stalling, which occurs at low concentrations of

    inducing macrolides, results in rearrangement of the

    mRNA secondary structure and activation of expression

    of the downstream ermmethyltransferase cistron. Stalling

    critically depends on the sequence of the leader peptide

    and is mediated by specific interactions between the ribo-

    some, the drug and the nascent peptide. However, the

    details of the operation of this mechanism remain largely

    unknown. As Erm can interact with 23S rRNA only before

    or during ribosome assembly (Skinner et al., 1983;

    Champney et al., 2003), the ensuing resistance is basedon an intricate kinetic interplay between induction of erm

    translation owing to stalling of the sensitive ribosome on

    the leader peptide open reading frame (ORF), production

    of functional Erm by resistant or drug-free ribosomes, and

    accumulation of newly assembled resistant ribosomes.

    Better understanding of the rates of individual reactions

    and kinetic modelling of the entire network involved in

    induction may help to find ways of evading this major

    resistance mechanism. Little is known about the molecu-

    lar mechanisms of interactions leading to the ribosome

    stalling. It is clear that the structure of the drug can

    dramatically influence erm inducibility (Kamimiya and

    Weisblum, 1988). One of the acclaimed advantages of

    ketolides, the newest generation of macrolide antibiotics,

    is that these drugs do not induce inducible erm (Bryskier,

    2000). The lack of the cladinose sugar at position 3 of the

    lactone ring in ketolides is a likely reason why ketolides

    do not activate ermexpression. However, it remains to be

    elucidated whether ketolides, in contrast to other mac-

    rolides, are unable to cause the ribosome stalling on the

    leader ORF. Progress in this area will critically depend on

  • 7/28/2019 Ten Son Man Kin 2006

    9/14

    Antibiotics and the ribosome 9

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    our understanding of the conformation of the nascent

    peptide and its progression through the exit tunnel.

    Induction of erm resistance is likely related to another

    resistance mechanism that also involves non-covalent

    interaction between the ribosome and the nascent pep-

    tides. Production of specific small peptides can render

    cells resistant to a variety of macrolide antibiotics (Tenson

    and Mankin, 2001; Vimberg et al., 2004). The resistance

    peptides affect only the ribosome on which they were

    synthesized, a phenomenon common to several other

    functionally important nascent peptides (Tenson and Man-

    kin, 1995; Lovett and Rogers, 1996; Tenson and Ehren-

    berg, 2002; Cruz-Vera et al., 2005). Resistance conferred

    by short peptide expression is based on specific interac-

    tions between the nascent peptide and the macrolide, as

    different peptide sequences produce resistance against

    structurally different macrolides (Vimberg et al., 2004).

    The available experimental data seem to be compatible

    with a bottle brush model according to which resistance

    peptides eject the drug from the ribosome (M. Lovmar

    et al., in preparation). Yet again, the nature of molecularinteractions in the ribosome exit tunnel leading to drug

    expulsion is unknown.

    We already mentioned an emerging new paradigm in

    understanding drugribosome interactions the recogni-

    tion that ribosomes of different organisms may vary

    substantially in their structure. This view has a direct impli-

    cation in studies of antibiotic resistance. The general loca-

    tion of the site of antibiotic action is usually the same in

    different bacteria and, as a result, resistance mutations in

    different species cluster in the same rRNA segments

    (Vester and Douthwaite, 2001). However, specific effects

    of mutations are often organism-specific. The recent find-ings by the Boettger and Yonath laboratories indicated that

    the fitness cost of the A2058G mutation that confers MLSBresistance depends on the nature of the neighbouring

    polymorphic 20572611 base pair (Pfister et al., 2005).

    As a result, the same resistance mutation may have a

    substantially different cost of fitness in different organ-

    isms. An even more striking example of organism-specific

    resistance mutations comes from studies of oxazolidino-

    nes (reviewed in Meka and Gold, 2004). In clinical isolates

    of streptococci, enterococci and staphylococci, oxazolidi-

    none resistance is usually associated with a mutation at

    position 2576 of 23S rRNA. In the laboratory-selected

    resistant E. coli strains, resistance resulted from a

    mutation at position 2032. In Mycobacterium smegmatis,

    resistance was conferred by the G2447U mutation.

    In the archaeon Halobacterium halobium, the resistance

    resulted from several mutations in domain V of 23S rRNA,

    with U2500C providing the highest resistance (Kloss

    et al., 1999; Xiong et al., 2000; Prystowsky et al., 2001;

    Sander et al., 2002). Although all of these mutations clus-

    ter in the peptidyl transferase centre, thus revealing it as

    the primary site of oxazolidinone action, their diversity

    suggests either that the binding of the drug is somewhat

    different in ribosomes of different organisms or that the

    fitness cost of mutations varies significantly between dif-

    ferent species.

    The fitness cost of resistance mutations is an important

    factor that affects appearance, fixation and maintenance

    of the resistance. Acquisition of resistance is often asso-

    ciated with a certain loss of fitness that can be manifested,

    among other things, in hypersensitivity to several stress

    conditions. For example, Salmonellastrains that became

    resistant to fusidic acid owing to acquisition of the P413L

    mutation in the translation factor EF-G are hypersensitive

    to oxidative stress, UV light and several antibiotics

    (Macvanin et al., 2004; Macvanin and Hughes, 2005).

    A secondary mutation in the EF-G allele, T423I, partially

    restores cell fitness (Nagaev et al., 2001). Compounds

    targeting ribosomal sites where resistance mutations will

    have high cost of fitness and could not be compensated

    by the second site mutations should make good antibiot-

    ics. Approaches are being developed to identify such sitesin the ribosome (Laios et al., 2004).

    Adverse antibiotic effects mediated by inhibition of

    mitochondrial protein synthesis

    Most clinically useful inhibitors of protein synthesis act on

    functionally important sites of the ribosome, such as

    the decoding centre, peptidyl transferase or GTPase-

    activating region (Spahn and Prescott, 1996). These sites

    are highly conserved between ribosomes of different

    domains of life, and only minor differences in the struc-

    tures of bacterial and eukaryotic cytoplasmic ribosomesusually account for antibiotic selectivity.

    An even bigger problem is avoiding inhibition of mito-

    chondrial protein synthesis. Evolutionarily, bacterial

    ribosomes are much closer to mitochondrial than to

    cytoplasmic eukaryotic ribosomes. In spite of the much

    smaller size of mitochondrial rRNA as compared with that

    of bacteria, structural similarity of the functionally critical

    regions of the mitochondrial and bacterial rRNA allows

    some antibiotics to inhibit mitochondrial translation

    (Bottger et al., 2001). One example is chloramphenicol, a

    peptidyl transferase inhibitor that was widely used in the

    past (Spahn and Prescott, 1996). Inhibition of mitochon-

    drial protein synthesis by chloramphenicol (Denslow and

    OBrien, 1978) seems to parallel the myelosuppression

    induced by the drug, and it was the main reason why the

    use of this antibiotic was discontinued in most developed

    countries. The new class of ribosome-targeting antibiotics,

    oxazolidinones, that, similarly to chloramphenicol, act on

    the peptidyl transferase centre of the large ribosomal sub-

    unit (Bobkova et al., 2003) can cause myelosuppression

    upon prolonged administration (Kuter and Tillotson, 2001).

  • 7/28/2019 Ten Son Man Kin 2006

    10/14

    10 T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    Studies by Leach and co-workers at Pharmacia and

    Upjohn (now Pfizer) indicated that inhibition of mitochon-

    drial synthesis is a likely cause of oxazolidinone adverse

    effects and is mediated by the drug binding to the large

    subunit of mammalian mitochondrial ribosomes (Nagiec

    et al., 2005).

    Other drugs, such as aminoglycosides, that act on the

    small ribosomal subunit can also interfere with mitochon-

    drial protein synthesis. The unfortunate consequence

    is neural damage, resulting in hearing loss (Fischel-

    Ghodsian et al., 2004; Li et al., 2005). As toxic effects are

    seen only in a limited number of patients, it seems that

    the toxicity depends on the genetic variations between

    individuals. Some mutations in mitochondrial rRNA that

    appear to increase the similarity of the mitochondrial

    rRNA to its eubacterial homologue have been shown to

    correlate with the aminoglycoside-induced hearing loss

    (Fischel-Ghodsian et al., 2004; Li et al., 2005). Other

    mitochondrial mutations can potentially decrease the gen-

    eral performance of mitochondrial ribosomes so that even

    limited inhibition of their activity by antibiotics may bringthe level of protein synthesis to below the critical threshold

    (Jacobs, 2003). It would be important to establish a

    correlation between the polymorphism of human mito-

    chondrial rRNA and the sensitivity of mitochondrial

    protein synthesis to inhibitors of bacterial translation.

    Such a pharmacogenomic approach may help to optimize

    antibiotic regimens on the basis of the patients genetic

    background.

    New drugs

    In recent years, we saw several impressive breakthroughsin the development of new, clinically useful ribosomal anti-

    biotics (Sutcliffe, 2005). Significant progress has been

    achieved with macrolides and oxazolidinones. Develop-

    ment of newer macrolides was driven by the need for

    drugs that can inhibit the growth of pathogens that devel-

    oped macrolide resistance. Ketolides, which represent a

    prominent new class of macrolide antibiotics, can do

    exactly that (Zhong and Shortridge, 2001). Different

    ketolides, including telithromycin developed by Aventis

    and approved by the US Food and Drug Administration

    (FDA) in 2004, cethromycin from Abbott, bridged ketolides

    introduced by Enanta, and other ketolides have a cladi-

    nose sugar at position 3 of the lactone ring replaced by a

    ketone function (Fig. 3). In addition, ketolides usually

    carry a carbamate cycle at positions 11,12 of the lactone,

    and, most importantly, they all have an extended aromatic

    side-chain whose site of attachment differs in different

    drugs (Fig. 3). As mentioned earlier, the aromatic side-

    chain may establish new interactions in the ribosome,

    facilitating ketolide binding (even though it remains to be

    seen which contacts the aromatic side-chain establishes

    with ribosomes of real pathogens in vivo). Other interac-

    tions of ketolides with the ribosome may also be different

    from those of older macrolides (Garza-Ramos et al.,

    2001). Differences in the drugribosome contacts account

    for the higher affinity of ketolides for the ribosome and

    help these drugs to overcome, at least partly, the resis-

    tance caused by Erm methylation of rRNA (Douthwaite

    et al., 2005). In addition, the lack of 3-cladinose is thought

    to prevent ketolides from activating expression of inducible

    Erm genes (reviewed in Bryskier, 2000).

    A new generation of tetracyclines, glycylcyclines,

    entered the market in June 2005 with FDA approval of the

    first drug of this class, tigecycline (Tygacyl), developed by

    Weyth. Tigecycline (Fig. 3) carries a glycylamido substitu-

    tion at position 9 of the tetracycline ring structure and

    shows activity against many resistant strains. Most

    importantly, in contrast to the tetracyclines of previous

    generations, tigecycline is not affected by conventional

    tetracycline resistance mechanisms mediated by efflux or

    ribosomal protection proteins (Bauer et al., 2004).

    After more than 25 years during which all of the newantibiotics were a mere improvement of old antibiotic plat-

    forms, a principally new class of antibacterials was intro-

    duced in 2001 (reviewed in Bozdogan and Appelbaum,

    2004). Linezolid (Fig. 3), a protein synthesis inhibitor rep-

    resenting a new oxazolidinone class of antibiotics, was

    developed by Pharmacia and Upjohn. The drug shows

    excellent activity against many Gram-positive pathogens

    and was not affected by existing mechanisms of ribosomal

    resistance. Oxazolidinones bind at the A site of the ribo-

    somal peptidyl transferase centre and inhibit its activity

    (although in a not fully understood fashion) (Shinabarger

    et al., 1997; Kloss et al., 1999; Polacek et al., 2002; Bobk-ova et al., 2003; Colca et al., 2003; Ippolito et al., 2005).

    Several lines of evidence suggest that translation initiation

    may be the main target of the drug action (Swaney et al.,

    1998); however, other scenarios cannot be discarded

    (Thompson et al., 2002) and more work is needed to sort

    out the exact mechanism of translation inhibition by

    oxazolidinones. It is most regrettable therefore that most

    pharmaceutical companies developing oxazolidinones (or,

    for that matter, other ribosome-targeting antibiotics) do not

    appear to express much interest in the mechanisms of the

    drug action. The university laboratories will apparently

    need to pick up the slack to unravel the mode of action of

    this important class of antibacterials.

    Several other ribosome-targeting antibiotics are cur-

    rently at different stages of development and characteriza-

    tion. These include pleuromutilins (Schlunzen et al., 2004),

    TAN-1057 (Limburg et al., 2004), evernimicin (Belova

    et al., 2001) and several others. An example of structure-

    based drug design potential is represented by a series of

    compounds being developed by Rib-X, a start-up biotech-

    nology company that uses ribosome crystallography and

  • 7/28/2019 Ten Son Man Kin 2006

    11/14

    Antibiotics and the ribosome 11

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    computational modelling as the main guides for develop-

    ment of new drugs (Ippolito et al., 2005). The proximity of

    the binding sites of linezolid and sparsomycin (Hansen

    et al., 2003) in the large ribosomal subunit provided direc-

    tion for the design of Rx-01 series of oxazolidinone com-

    pounds whose affinity to the ribosome is improved due to

    additional interactions in the sparsomycin site (Lawrence

    and Sutcliffe, 2005). One of these compounds has entered

    phase I clinical trials at the end of 2005.

    Concluding remarks

    In this review, we briefly touched on several aspects that,

    in our opinion, represent the major trends in the develop-

    ment of the field of ribosome-targeting antibiotics. Inevita-

    bly, this view is skewed by our own research interests, and

    we realize that we might have left out some exciting lines

    of research. However, what we have tried to communicate

    is the recent rapid progress in the field that was brought

    about by high-resolution crystallographic structures of the

    ribosome and its complexes with antibiotics and wasfuelled by important developments in other techniques

    and approaches. New antibiotics, inhibitors of protein syn-

    thesis that are already used in medical practice as well as

    drugs in the pipelines of several pharmaceutical compa-

    nies will hopefully keep the field going for the foreseeable

    future. However, a general pullout of the Big Pharma from

    the antibiotic field and the continuous merging of pharma-

    ceutical companies into bigger conglomerates eliminated

    many important lines of antibiotic research, including stud-

    ies of ribosome-targeting drugs. Thus, the influence of

    academic science in the near future on the general direc-

    tion of ribosome antibiotic research is expected to grow.We believe that studies of the fundamental aspects of the

    drug action that are currently funded primarily by non-

    industrial sources will generate new important venues for

    drug development that will eventually be exploited by the

    pharmaceutical industry. Importantly, these studies will

    also advance our understanding of the important aspects

    of protein synthesis and will provide new tools for studies

    of translation.

    Acknowledgements

    We want to thank the participants of the meeting Antibioticsand Translation for stimulating discussions that inspired us

    for writing this review. Our special thanks are to Stephen

    Douthwaite, Joyce Sutcliffe, Karen Leach, Peter Moore,

    Martin Lovmar and Julian Davies for the comments on the

    manuscripts. We thank Shannon Foley for proofreading. The

    antibiotic-related work in the T. Tenson laboratory is sup-

    ported by The Wellcome Trust International Senior Fellowship

    (070210/Z/03/Z) and grant from Estonian Science Founda-

    tion (5311) and in the A. Mankin laboratory by NIH Grant U19

    AI56575.

    References

    Antoun, A., Pavlov, M.Y., Tenson, T., and Ehrenberg, M.M.

    (2004) Ribosome formation from subunits studied by

    stopped-flow and Rayleigh light scattering. Biol Proced

    Online6: 3554.

    Bauer, G., Berens, C., Projan, S.J., and Hillen, W. (2004)

    Comparison of tetracycline and tigecycline binding to ribo-

    somes mapped by dimethylsulphate and drug-directed

    Fe2+ cleavage of 16S rRNA. J Antimicrob Chemother53:592599.

    Belova, L., Tenson, T., Xiong, L., McNicholas, P.M., and

    Mankin, A.S. (2001) A novel site of antibiotic action in

    the ribosome: interaction of evernimicin with the large

    ribosomal subunit. Proc Natl Acad Sci USA 98: 3726

    3731.

    Berisio, R., Harms, J., Schluenzen, F., Zarivach, R., Hansen,

    H.A., Fucini, P., and Yonath, A. (2003a) Structural insight

    into the antibiotic action of telithromycin against resistant

    mutants. J Bacteriol185: 42764279.

    Berisio, R., Schluenzen, F., Harms, J., Bashan, A., Auer-

    bach, T., Baram, D., and Yonath, A. (2003b) Structural

    insight into the role of the ribosomal tunnel in cellular

    regulation. Nat Struct Biol10: 366370.Bobkova, E.V., Yan, Y.P., Jordan, D.B., Kurilla, M.G., and

    Pompliano, D.L. (2003) Catalytic properties of mutant 23S

    ribosomes resistant to oxazolidinones. J Biol Chem 278:

    98029807.

    Bottger, E.C., Springer, B., Prammananan, T., Kidan, Y., and

    Sander, P. (2001) Structural basis for selectivity and toxic-

    ity of ribosomal antibiotics. EMBO Rep2: 318323.

    Bozdogan, B., and Appelbaum, P.C. (2004) Oxazolidinones:

    activity, mode of action, and mechanism of resistance. Int

    J Antimicrob Agents23: 113119.

    Brodersen, D.E., Clemons, W.M., Jr, Carter, A.P., Morgan-

    Warren, R.J., Wimberly, B.T., and Ramakrishnan, V.

    (2000) The structural basis for the action of the antibiotics

    tetracycline, pactamycin, and hygromycin B on the 30Sribosomal subunit. Cell103: 11431154.

    Bryskier, A. (2000) Ketolides-telithromycin, an example of a

    new class of antibacterial agents. Clin Microbiol Infect6:

    661669.

    Carter, A.P., Clemons, W.M., Brodersen, D.E., Morgan-

    Warren, R.J., Wimberly, B.T., and Ramakrishnan, V.

    (2000) Functional insights from the structure of the 30S

    ribosomal subunit and its interactions with antibiotics.

    Nature407: 340348.

    Champney, W.S., and Burdine, R. (1996) 50S ribosomal

    subunit synthesis and translation are equivalent targets for

    erythromycin inhibition in Staphylococcus aureus. Antimi-

    crob Agents Chemother40: 13011303.

    Champney, W.S., and Miller, M. (2002a) Inhibition of 50Sribosomal subunit assembly in Haemophilus influenzae

    cells by azithromycin and erythromycin. Curr Microbiol44:

    418424.

    Champney, W.S., and Miller, M. (2002b) Linezolid is a spe-

    cific inhibitor of 50S ribosomal subunit formation in Staphy-

    lococcus aureuscells. Curr Microbiol44: 350356.

    Champney, W.S., Chittum, H.S., and Tober, C.L. (2003) A

    50S ribosomal subunit precursor particle is a substrate for

    the ErmC methyltransferase in Staphylococcus aureus

    cells. Curr Microbiol46: 453460.

  • 7/28/2019 Ten Son Man Kin 2006

    12/14

    12 T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    Colca, J.R., McDonald, W.G., Waldon, D.J., Thomasco, L.M.,

    Gadwood, R.C., Lund, E.T., et al. (2003) Cross-linking in

    the living cell locates the site of action of oxazolidinone

    antibiotics. J Biol Chem278: 2197221979.

    Cruz-Vera, L.R., Rajagopal, S., Squires, C., and Yanofsky, C.

    (2005) Features of ribosomepeptidyltRNA interactions

    essential for tryptophan induction of tna operon expres-

    sion. Mol Cell19: 333343.

    Cundliffe, E. (1990) Recognition sites for antibiotics within

    rRNA. In The Ribosome: Structure, Function and Evolu-

    tion. Hill, W. E., Dalberg, A., Garrett, R.A., Moore, P.B.,

    Schlessinger, D., and Warner, J.R. (eds). Washington, DC:

    American Society for Microbiology, pp. 479490.

    Denslow, N.D., and OBrien, T.W. (1978) Antibiotic suscep-

    tibility of the peptidyl transferase locus of bovine mitochon-

    drial ribosomes. Eur J Biochem91: 441448.

    Dinos, G., Wilson, D.N., Teraoka, Y., Szaflarski, W., Fucini,

    P., Kalpaxis, D., and Nierhaus, K.H. (2004) Dissecting the

    ribosomal inhibition mechanisms of edeine and pactamy-

    cin: the universally conserved residues G693 and C795

    regulate P-site RNA binding. Mol Cell13: 113124.

    Dodd, J., Kolb, J.M., and Nomura, M. (1991) Lack of com-

    plete cooperativity of ribosome assembly in vitro and its

    possible relevance to in vivo ribosome assembly and theregulation of ribosomal gene expression. Biochimie 73:

    757767.

    Douthwaite, S., Crain, P.F., Liu, M., and Poehlsgaard, J.

    (2004) The tylosin-resistance methyltransferase RlmA(II)

    (TlrB) modifies the N-1 position of 23S rRNA nucleotide

    G748. J Mol Biol337: 10731077.

    Douthwaite, S., Jalava, J., and Jakobsen, L. (2005) Ketolide

    resistance in Streptococcus pyogenescorrelates with the

    degree of rRNA dimethylation by Erm. Mol Microbiol 58:

    613622.

    Dubnau, D. (1984) Translational attenuation: the regulation

    of bacterial resistance to the macrolide-lincosamide-

    streptogramin B antibiotics. CRC Crit Rev Biochem 16:

    103132.Evers, S., Di Padova, K., Meyer, M., Langen, H., Fountou-

    lakis, M., Keck, W., and Gray, C.P. (2001) Mechanism-

    related changes in the gene transcription and protein syn-

    thesis patterns of Haemophilus influenzaeafter treatment

    with transcriptional and translational inhibitors. Proteomics

    1: 522544.

    Fischel-Ghodsian, N., Kopke, R.D., and Ge, X. (2004) Mito-

    chondrial dysfunction in hearing loss. Mitochondrion 4:

    675694.

    Fischer, H.P., Brunner, N.A., Wieland, B., Paquette, J.,

    Macko, L., Ziegelbauer, K., and Freiberg, C. (2004)

    Identification of antibiotic stress-inducible promoters: a

    systematic approach to novel pathway-specific reporter

    assays for antibacterial drug discovery. Genome Res14:9098.

    Fourmy, D., Recht, M.I., Blanchard, S.C., and Puglisi, J.D.

    (1996) Structure of the A site of Escherichia coli16S ribo-

    somal RNA complexed with an aminoglycoside antibiotic.

    Science274: 13671371.

    Francois, B., Russell, R.J., Murray, J.B., Aboul-ela, F.,

    Masquida, B., Vicens, Q., and Westhof, E. (2005) Crystal

    structures of complexes between aminoglycosides and

    decoding A site oligonucleotides: role of the number of

    rings and positive charges in the specific binding leading

    to miscoding. Nucleic Acids Res33: 56775690.

    Freistroffer, D.V., Pavlov, M.Y., MacDougall, J., Buckingham,

    R.H., and Ehrenberg, M. (1997) Release factor RF3 in E.

    coliaccelerates the dissociation of release factors RF1 and

    RF2 from the ribosome in a GTP-dependent manner.

    EMBO J16: 41264133.

    Garza-Ramos, G., Xiong, L., Zhong, P., and Mankin, A.

    (2001) Binding site of macrolide antibiotics on the ribo-

    some: new resistance mutation identifies a specific inter-

    action of ketolides with rRNA. J Bacteriol183: 68986907.

    Goh, E.B., Yim, G., Tsui, W., McClure, J., Surette, M.G., and

    Davies, J. (2002) Transcriptional modulation of bacterial

    gene expression by subinhibitory concentrations of antibi-

    otics. Proc Natl Acad Sci USA99: 1702517030.

    Gromadski, K.B., and Rodnina, M.V. (2004) Streptomycin

    interferes with conformational coupling between codon

    recognition and GTPase activation on the ribosome. Nat

    Struct Mol Biol11: 316322.

    Hansen, L.H., Mauvais, P., and Douthwaite, S. (1999) The

    macrolide-ketolide antibiotic binding site is formed by struc-

    tures in domains II and V of 23S ribosomal RNA. Mol

    Microbiol31: 623631.

    Hansen, L.H., Kirpekar, F., and Douthwaite, S. (2001) Rec-ognition of nucleotide G745 in 23S ribosomal RNA by the

    rrmA methyltransferase. J Mol Biol310: 10011010.

    Hansen, J.L., Ippolito, J.A., Ban, N., Nissen, P., Moore, P.B.,

    and Steitz, T.A. (2002) The structures of four macrolide

    antibiotics bound to the large ribosomal subunit. Mol Cell

    10: 117128.

    Hansen, J.L., Moore, P.B., and Steitz, T.A. (2003) Structures

    of five antibiotics bound at the peptidyl transferase center

    of the large ribosomal subunit. J Mol Biol330: 10611075.

    Harms, J., Schluenzen, F., Zarivach, R., Bashan, A., Gat, S.,

    Agmon, I., et al. (2001) High resolution structure of the

    large ribosomal subunit from a mesophilic eubacterium.

    Cell107: 679688.

    Hutter, B., Schaab, C., Albrecht, S., Borgmann, M., Brun-ner, N.A., Freiberg, C., et al. (2004) Prediction of mecha-

    nisms of action of antibacterial compounds by gene

    expression profiling. Antimicrob Agents Chemother 48:

    28382844.

    Ippolito, J., Kanyo, Z., Wimberly, B., Wang, D., Skripkin, E.,

    Devito, J., et al. (2005) Structural basis for the binding Rx-

    01, a novel oxazolidinone class, to bacterial ribosomes. In

    Program and Abstracts of the 45th Interscience Confer-

    ence on Antimicrobial Agents and Chemotherapy. Wash-

    ington, DC: American Society for Microbiology, abst. F-

    1254.

    Jacobs, H.T. (2003) Disorders of mitochondrial protein syn-

    thesis. Hum Mol Genet12 (Spec No 2): R293R301.

    Kamimiya, S., and Weisblum, B. (1988) Translational attenu-ation control of ermSF, an inducible resistance determinant

    encoding rRNA N-methyltransferase from Streptomyces

    fradiae. J Bacteriol170: 18001811.

    Kloss, P., Xiong, L., Shinabarger, D.L., and Mankin, A.S.

    (1999) Resistance mutations in 23S rRNA identify the site

    of action of the protein synthesis inhibitor linezolid in the

    ribosomal peptidyl transferase center. J Mol Biol294: 93

    101.

    Kuter, D.J., and Tillotson, G.S. (2001) Hematologic effects of

  • 7/28/2019 Ten Son Man Kin 2006

    13/14

    Antibiotics and the ribosome 13

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    antimicrobials: focus on the oxazolidinone linezolid. Phar-

    macotherapy21: 10101013.

    Laios, E., Waddington, M., Saraiya, A.A., Baker, K.A.,

    OConnor, E., Pamarathy, D., and Cunningham, P.R.

    (2004) Combinatorial genetic technology for the develop-

    ment of new anti-infectives. Arch Pathol Lab Med 128:

    13511359.

    Lawrence, L., and Sutcliffe, J. (2005) In vitro activity of

    designer oxazolidinones against respiratory pathogens. In

    Program and Abstracts of the 45th Interscience Confer-

    ence on Antimicrobial Agents and Chemotherapy.

    Washington, DC: American Society for Microbiology, abst.

    F-1256.

    Li, Z., Li, R., Chen, J., Liao, Z., Zhu, Y., Qian, Y., et al. (2005)

    Mutational analysis of the mitochondrial 12S rRNA gene in

    Chinese pediatric subjects with aminoglycoside-induced

    and non-syndromic hearing loss. Hum Genet117: 915.

    Limburg, E., Gahlmann, R., Kroll, H.P., and Beyer, D. (2004)

    Ribosomal alterations contribute to bacterial resistance

    against the dipeptide antibiotic TAN 1057. Antimicrob

    Agents Chemother48: 619622.

    Lin, A.H., Murray, R.W., Vidmar, T.J., and Marotti, K.R.

    (1997) The oxazolidinone eperezolid binds to the 50S ribo-

    somal subunit and competes with binding of chlorampheni-col and lincomycin. Antimicrob Agents Chemother 41:

    21272131.

    Liu, M., and Douthwaite, S. (2002) Resistance to the mac-

    rolide antibiotic tylosin is conferred by single methylations

    at 23S rRNA nucleotides G748 and A2058 acting in syn-

    ergy. Proc Natl Acad Sci USA99: 1465814663.

    Lovett, P.S., and Rogers, E.J. (1996) Ribosome regulation

    by the nascent peptide. Microbiol Rev60: 366385.

    Lovmar, M., Tenson, T., and Ehrenberg, M. (2004) Kinetics

    of macrolide action: the josamycin and erythromycin

    cases. J Biol Chem279: 5350653515.

    Luidalepp, H., Hallier, M., Felden, B., and Tenson, T. (2005)

    tmRNA decreases the bactericidal activity of aminoglyco-

    sides and the susceptibility to inhibitors of cell wall synthe-sis. RNA Biol2: 7074.

    Lynch, S.R., Gonzalez, R.L., and Puglisi, J.D. (2003) Com-

    parison of X-ray crystal structure of the 30S subunit

    antibiotic complex with NMR structure of decoding site

    oligonucleotideparomomycin complex. Structure (Camb)

    11: 4353.

    Macvanin, M., and Hughes, D. (2005) Hyper-susceptibility of

    a fusidic acid-resistant mutant of Salmonella to different

    classes of antibiotics. FEMS Microbiol Lett247: 215220.

    Macvanin, M., Ballagi, A., and Hughes, D. (2004) Fusidic

    acid-resistant mutants of Salmonella entericaserovar typh-

    imurium have low levels of heme and a reduced rate of

    respiration and are sensitive to oxidative stress. Antimicrob

    Agents Chemother48: 38773883.Madsen, C.T., Jakobsen, L., and Douthwaite, S. (2005)

    Mycobacterium smegmatis Erm(38) is a reluctant

    dimethyltransferase. Antimicrob Agents Chemother 49:

    38033809.

    Maguire, B.A., Beniaminov, A.D., Ramu, H., Mankin, A.S.,

    and Zimmermann, R.A. (2005) A protein component at the

    heart of an RNA machine: the importance of protein L27

    for the function of the bacterial ribosome. Mol Cell20: 427

    435.

    Matassova, N.B., Rodnina, M.V., Endermann, R., Kroll, H.P.,

    Pleiss, U., Wild, H., and Wintermeyer, W. (1999) Riboso-

    mal RNA is the target for oxazolidinones, a novel class of

    translational inhibitors. RNA5: 939946.

    Mehta, R., and Champney, W.S. (2003) Neomycin and paro-

    momycin inhibit 30S ribosomal subunit assembly in

    Staphylococcus aureus. Curr Microbiol47: 237243.

    Meka, V.G., and Gold, H.S. (2004) Antimicrobial resistance

    to linezolid. Clin Infect Dis39: 10101015.

    Meka, V.G., Pillai, S.K., Sakoulas, G., Wennersten, C., Ven-

    kataraman, L., DeGirolami, P.C., et al. (2004) Linezolid

    resistance in sequential Staphylococcus aureus isolates

    associated with a T2500A mutation in the 23S rRNA gene

    and loss of a single copy of rRNA. J Infect Dis190: 311

    317.

    Menninger, J.R., and Otto, D.P. (1982) Erythromycin, carbo-

    mycin, and spiramycin inhibit protein synthesis by stimulat-

    ing the dissociation of peptidyl-tRNA from ribosomes.

    Antimicrob Agents Chemother21: 811818.

    Moore, P.B., and Steitz, T.A. (2002) The involvement of RNA

    in ribosome function. Nature418: 229235.

    Nagaev, I., Bjorkman, J., Andersson, D.I., and Hughes, D.

    (2001) Biological cost and compensatory evolution in

    fusidic acid-resistant Staphylococcus aureus. Mol Micro-biol40: 433439.

    Nagiec, E.E., Wu, L., Swaney, S.M., Chosay, J.G., Ross,

    D.E., Brieland, J.K., and Leach, K.L. (2005) Oxazolidino-

    nes inhibit cellular proliferation via inhibition of mitochon-

    drial protein synthesis. Antimicrob Agents Chemother49:

    38963902.

    Ogle, J.M., and Ramakrishnan, V. (2005) Structural insights

    into translational fidelity. Annu Rev Biochem74: 129177.

    Ogle, J.M., Brodersen, D.E., Clemons, W.M., Jr, Tarry, M.J.,

    Carter, A.P., and Ramakrishnan, V. (2001) Recognition

    of cognate transfer RNA by the 30S ribosomal subunit.

    Science292: 897902.

    Pape, T., Wintermeyer, W., and Rodnina, M.V. (2000) Con-

    formational switch in the decoding region of 16S rRNAduring aminoacyl-tRNA selection on the ribosome. Nature

    Struct Biol7: 104107.

    Pfister, P., Corti, N., Hobbie, S., Bruell, C., Zarivach, R.,

    Yonath, A., and Bottger, E.C. (2005) 23S rRNA base pair

    20572611 determines ketolide susceptibility and fitness

    cost of the macrolide resistance mutation 2058A-->G. Proc

    Natl Acad Sci USA102: 51805185.

    Picking, W.D., Odom, O.W., Tsalkova, T., Serdyuk, I., and

    Hardesty, B. (1991) The conformation of nascent polyl-

    ysine and polyphenylalanine peptides on ribosomes. J Biol

    Chem266: 15341542.

    Polacek, N., Swaney, S., Shinabarger, D., and Mankin, A.S.

    (2002) SPARK a novel method to monitor ribosomal

    peptidyl transferase activity. Biochemistry 41: 1160211610.

    Prystowsky, J., Siddiqui, F., Chosay, J., Shinabarger, D.L.,

    Millichap, J., Peterson, L.R., and Noskin, G.A. (2001)

    Resistance to linezolid: characterization of mutations in

    rRNA and comparison of their occurrences in vancomycin-

    resistant enterococci. Antimicrob Agents Chemother 45:

    21542156.

    Sabina, J., Dover, N., Templeton, L.J., Smulski, D.R., Soll,

    D., and LaRossa, R.A. (2003) Interfering with different

  • 7/28/2019 Ten Son Man Kin 2006

    14/14

    14 T. Tenson and A. Mankin

    2006 The AuthorsJournal compilation 2006 Blackwell Publishing Ltd, Molecular Microbiology

    steps of protein synthesis explored by transcriptional pro-

    filing of Escherichia coliK-12. J Bacteriol185: 61586170.

    Sander, P., Belova, L., Kidan, Y.G., Pfister, P., Mankin, A.S.,

    and Bottger, E.C. (2002) Ribosomal and non-ribosomal

    resistance to oxazolidinones: species-specific idiosyncrasy

    of ribosomal alterations. Mol Microbiol46: 12951304.

    Schlunzen, F., Zarivach, R., Harms, J., Bashan, A., Tocilj, A.,

    Albrecht, R., et al. (2001) Structural basis for the interaction

    of antibiotics with the peptidyl transferase centre in eubac-

    teria. Nature413: 814821.

    Schlunzen, F., Harms, J.M., Franceschi, F., Hansen, H.A.,

    Bartels, H., Zarivach, R., and Yonath, A. (2003) Structural

    basis for the antibiotic activity of ketolides and azalides.

    Structure (Camb)11: 329338.

    Schlunzen, F., Pyetan, E., Fucini, P., Yonath, A., and Harms,

    J.M. (2004) Inhibition of peptide bond formation by pleuro-

    mutilins: the structure of the 50S ribosomal subunit from

    Deinococcus radiodurans in complex with tiamulin. Mol

    Microbiol54: 12871294.

    Schmeing, T.M., Seila, A.C., Hansen, J.L., Freeborn, B.,

    Soukup, J.K., Scaringe, S.A., et al. (2002) A pre-

    translocational intermediate in protein synthesis observed

    in crystals of enzymatically active 50S subunits. Nat Struct

    Biol9: 225230.Shah, P.M. (2005) The need for new therapeutic agents:

    what is the pipeline? Clin Microbiol Infect 11 (Suppl. 3):

    3642.

    Shapiro, E., and Baneyx, F. (2002) Stress-based identifica-

    tion and classification of antibacterial agents: second-

    generation Escherichia colireporter strains and optimiza-

    tion of detection. Antimicrob Agents Chemother46: 2490

    2497.

    Shinabarger, D.L., Marotti, K.R., Murray, R.W., Lin, A.H.,

    Melchior, E.P., Swaney, S.M., et al. (1997) Mechanism of

    action of oxazolidinones: effects of linezolid and eperezolid

    on translation reactions. Antimicrob Agents Chemother41:

    21322136.

    Skinner, R., Cundliffe, E., and Schmidt, F.J. (1983) Site ofaction of a ribosomal RNA methylase responsible for resis-

    tance to erythromycin and other antibiotics. J Biol Chem

    258: 1270212706.

    Spahn, C.M., and Prescott, C.D. (1996) Throwing a spanner

    in the works: antibiotics and the translation apparatus.

    J Mol Med74: 423439.

    Sutcliffe, J.A. (2005) Improving on nature: antibiotics that

    target the ribosome. Curr Opin Microbiol8: 534542.

    Swaney, S.M., Aoki, H., Ganoza, M.C., and Shinabarger,

    D.L. (1998) The oxazolidinone linezolid inhibits initiation

    of protein synthesis in bacteria. Antimicrob Agents

    Chemother42: 32513255.

    Tenson, T., and Ehrenberg, M. (2002) Regulatory nascent

    peptides in the ribosomal tunnel. Cell108: 591594.Tenson, T., and Mankin, A. (1995) Comparison of functional

    peptide encoded in the Escherichia coli 23S rRNA with

    other peptides involved in cis-regulation of translation.

    Biochem Cell Biol73: 10611070.

    Tenson, T., and Mankin, A.S. (2001) Short peptides confer-

    ring resistance to macrolide antibiotics. Peptides22: 1661

    1668.

    Tenson, T., Lovmar, M., and Ehrenberg, M. (2003) The mech-

    anism of action of macrolides, lincosamides and strepto-

    gramin B reveals the nascent peptide exit path in the

    ribosome. J Mol Biol330: 10051014.

    Thompson, J., OConnor, M., Mills, J.A., and Dahlberg, A.E.

    (2002) The protein synthesis inhibitors, oxazolidinones and

    chloramphenicol, cause extensive translational inaccuracy

    in vivo. J Mol Biol322: 273279.

    Tsiodras, S., Gold, H.S., Sakoulas, G., Eliopoulos, G.M.,

    Wennersten, C., Venkataraman, L., et al. (2001) Linezolid

    resistance in a clinical isolate of Staphylococcus aureus.

    Lancet358: 207208.

    Tsui, W.H., Yim, G., Wang, H.H., McClure, J.E., Surette,

    M.G., and Davies, J. (2004) Dual effects of MLS antibiotics:

    transcriptional modulation and interactions on the ribo-

    some. Chem Biol11: 13071316.

    Tu, D., Blaha, G., Moore, P.B., and Steitz, T.A. (2005)

    Structures of MLSBK antibiotics bound to mutated large

    ribosomal subunits provide a structural explanation for

    resistance. Cell121: 257270.

    VanBogelen, R.A., and Neidhardt, F.C. (1990) Ribosomes as

    sensors of heat and cold shock in Escherichia coli. Proc

    Natl Acad Sci USA87: 55895593.

    Vester, B., and Douthwaite, S. (2001) Macrolide resistance

    conferred by base substitutions in 23S rRNA. AntimicrobAgents Chemother45: 112.

    Vicens, Q., and Westhof, E. (2001) Crystal structure of paro-

    momycin docked into the eubacterial ribosomal decoding

    A site. Structure (Camb)9: 647658.

    Vicens, Q., and Westhof, E. (2003) Crystal structure of gene-

    ticin bound to a bacterial 16S ribosomal RNA A site oligo-

    nucleotide. J Mol Biol326: 11751188.

    Vimberg, V., Xiong, L., Bailey, M., Tenson, T., and Mankin,

    A. (2004) Peptide-mediated macrolide resistance reveals

    possible specific interactions in the nascent peptide exit

    tunnel. Mol Microbiol54: 376385.

    Weisblum, B. (1995a) Erythromycin resistance by ribosome

    modification. Antimicrob Agents Chemother39: 577585.

    Weisblum, B. (1995b) Insights into erythromycin action fromstudies of its activity as inducer of resistance. Antimicrob

    Agents Chemother39: 797805.

    Wilson, D.N., Harms, J.M., Nierhaus, K.H., Schlnzen, F.,

    and Fucini, P. (2005) Species-specific antibiotic-ribosome

    interactions: implications for drug development. Biol Chem

    386: 12391252.

    Xiong, L., Shah, S., Mauvais, P., and Mankin, A.S. (1999) A

    ketolide resistance mutation in domain II of 23S rRNA

    reveals the proximity of hairpin 35 to the peptidyl trans-

    ferase centre. Mol Microbiol31: 633639.

    Xiong, L., Kloss, P., Douthwaite, S., Andersen, N.M.,

    Swaney, S., Shinabarger, D.L., and Mankin, A.S. (2000)

    Oxazolidinone resistance mutations in 23S rRNA of

    Escherichia coli reveal the central region of domain V asthe primary site of drug action. J Bacteriol 182: 5325

    5331.

    Yonath, A. (2005) Antibiotics targeting ribosomes: resistance,

    selectivity, synergism and cellular regulation. Annu Rev

    Biochem74: 649679.

    Zhong, P., and Shortridge, V. (2001) The emerging new

    generation of antibiotic: ketolides. Curr Drug Targets Infect

    Disord1: 125131.