You are on page 1of 8

European Journal of Medicinal Chemistry xxx (2013) 1e8

Contents lists available at SciVerse ScienceDirect

European Journal of Medicinal Chemistry


journal homepage: http://www.elsevier.com/locate/ejmech

Mini-review

Antibacterial action of quinolones: From target to network


Guyue Cheng, Haihong Hao, Menghong Dai, Zhenli Liu, Zonghui Yuan*
National Reference Laboratory of Veterinary Drug Residues (HZAU) and MOA Key Laboratory for the Detection of Veterinary Drug Residues in Foods,
Huazhong Agricultural University, Wuhan, Hubei 430070, China

a r t i c l e i n f o a b s t r a c t

Article history: Quinolones are widely used broad-spectrum antibacterials with incomplete elucidated mechanism of
Received 25 December 2012 action. Here, molecular basis for the antibacterial action of quinolones, from target to network, is fully
Received in revised form discussed and updated. Quinolones trap DNA gyrase or topoisomerase IV to form reversible drug-
23 January 2013
enzyme-DNA cleavage complexes, resulting in bacteriostasis. Cell death arises from chromosome
Accepted 26 January 2013
Available online xxx
fragmentation in protein synthesis-dependent or -independent pathways according to distinguished
quinolone structures. In the former pathway, irreversible oxidative DNA damage caused by reactive
oxygen species kills bacteria eventually. Toxineantitoxin mazEF is triggered as an additional lethal action.
Keywords:
Primary target
Bacteria survive and develop resistance by SOS and other stress responses. Enlarged knowledges of
Bacteriostatic action quinolone actions and bacterial response will provide new targets for drug design and approaches to
Lethal action prevent bacterial resistance.
Crystal structure Crown Copyright Ó 2013 Published by Elsevier Masson SAS. All rights reserved.
Bacterial response

1. Introduction Now, it is quite clear that QLs interfere with chromosomal to-
pology by targeting bacterial type IIA topoisomerases, DNA gyrase
Quinolones (QLs) are synthetic antimicrobials based on the and topo IV, trapping these enzymes at the DNA cleavage stage and
4-oxo-1,4-dihydroquinolone skeleton (Fig. 1). The first-generation preventing strand rejoining. As a result, the DNA replication ma-
of QLs, such as nalidixic and oxolinic acids, acts against Gram- chinery becomes arrested at the blocked replication forks, leading
negative bacteria and is used to treat urinary tract infections. to inhibition of DNA synthesis that immediately causes bacterio-
The second-generation of QLs, such as norfloxacin and ciproflox- stasis [5]. Till now, several crystal structures have been resolved to
acin, is introduced a fluorine atom at position 6 and a bulky exhibit accurate structures of the drug-enzyme-DNA ternary
piperidine at position 7, broadening the antimicrobial spectrum to complexes, but data gained from these crystal structures demon-
Pseudomonas species and some Gram-positive organisms, e.g. strate some contradictions that need to be explained and unified.
Staphylococcus aureus. The third generation of QLs, with sub- QL-induced cell death is associated with the formation of double-
stitutions at the 7 as well as at the 8 position, enhances the ac- stranded DNA breaks (DSBs), resulting in chromosome fragmen-
tivity against Gram-positive bacteria. For instance, levofloxacin tation and surge of reactive oxygen species (ROS) [5,6]. QLs differ
and moxifloxacin are active against Streptococcus pneumoniae and among various derivatives in rate and extent of killing, in the effect
S. aureus that are pathogens responsible for respiratory tract in- of protein synthesis inhibitors on QL lethality and in the need for
fections, acute otitis and meningitis [1]. Moreover, moxifloxacin is aerobic metabolism to kill cells [7]. However, the molecular
active against Mycobacterium tuberculosis which lacks topoisom- mechanisms of these differences have not been clearly elucidated.
erase IV (topo IV) [2]. The fourth-generation of QLs is developed It is also known that chromosome-encoded toxin MazF which
with enhanced potency and a broader spectrum including causes programmed cell death (PCD) is involved in efficient QL
anaerobic bacteria. Gemifloxacin is currently one of the most killing [8]. To fight against the drug action, SOS regulon and other
potent fluoroquinolones against community-acquired pneumonia bacterial response networks are triggered in responses to QLs,
and acute bronchitis [3]. Trovafloxacin is used against intra- providing strategies for bacterial survival and development of
abdominal and pelvic infections [4]. resistance.
In this review, a diverse body of knowledge was drawn into the
mode of action of QLs from target to network levels. This mainly
* Corresponding author. Tel.: þ86 27 87287186; fax: þ86 27 87672232. includes the following details. First, targeting of QLs on DNA gyrase
E-mail addresses: yuan5802@mail.hzau.edu.cn, yuanzongh@126.com (Z. Yuan). or topo IV resulting in bacteriostatic actions was briefly described

0223-5234/$ e see front matter Crown Copyright Ó 2013 Published by Elsevier Masson SAS. All rights reserved.
http://dx.doi.org/10.1016/j.ejmech.2013.01.057

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
2 G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8

regulation of DNA superhelicity, bacterial replication, transcription


initiation and elongation. Gyrase generates a pair of single-
stranded breaks (SSBs) in the region of DNA wrapped around the
enzyme. One DNA gate is opened through which another stretch of
DNA can pass. After that, the gate closes and hydrolysis of ATP re-
sets gyrase for another round.
In 1990, Kato et al. [18] discovered a homolog of gyrase called
topoisomerase IV (topo IV). Like gyrase, topo IV is composed of two
ParC subunits and two ParE subunits [19]. Soon after the discovery of
topo IV, it became clear that gyrase was not the only intracellular
target of QLs [20,21]. In Escherichia coli, mutations mapped in parE
[22] or near parC [23] were identified conferring a high level of
resistance. Since QL resistance alleles in parE or parC did not confer
resistance by themselves, topo IV must be a secondary target in E. coli
[22e25]. This is also the same case for Neisseria gonorrhoeae [26].
However, things are rather different for some Gram-positive
strains. Clinical isolates of S. aureus that were resistant to a mod-
erate level of ciprofloxacin contained a mutation in parC (grlA)
while isolates that were resistant to a high concentration of cip-
rofloxacin exhibited an additional mutation in gyrA [27]. Purified
topo IV was inhibited by QLs from a variety of assays, including
measurement of DNA cleavage, decatenation and relaxation activity
[25,28,29]. Gyrase and topo IV from both S. aureus and E. coli were
purified to study the fluoroquinolone sensitivity [30]. The super-
coiling activity of S. aureus gyrase was at least 500-fold less sensi-
tive to ciprofloxacin than that of E. coli gyrase and about 6-fold less
sensitive than the decatenating activity of S. aureus topo IV. These
observations suggest that topo IV, rather than gyrase, is the primary
target of ciprofloxacin in S. aureus. The same phenomenon occurs in
S. pneumoniae [31,32].
In summary, it is now quite clear that bacteria contain two
topoisomerase targets for QLs. In Gram-negative bacteria the pri-
mary target is gyrase, while in Gram-positive bacteria the primary
target is generally topo IV. Since the two enzymes have different
functions, it is likely that bacteria differ in their response to QLs
Fig. 1. The chemical structures of QLs. according to which enzyme is the primary target.

3. Bacteriostatic actions of QLs


and crystal structures of drugeenzymeeDNA ternary complexes
were updated. Then, different chromosome fragmentation path- 3.1. Formation of drug-enzyme-DNA complex
ways and subsequent events leading to cell death were depicted in
details. Last but not least, the bacterial responses induced by QL Formation of a QLeenzymeeDNA complex that contains broken
treatment were summarized. Understanding the mechanisms DNA is the central event in the interaction between the QLs and
underlying QL actions and the corresponding bacterial responses gyrase or topo IV. Footprinting experiments suggested that the QLs
could provide new targets for drug design, enhance drug efficiency allowed gyrase to proceed to a conformational change in which
and prevent bacterial resistance. additional DNA was wrapped around gyrase [33]. The QLe
enzymeeDNA complex formation is readily reversible. Chromo-
2. Primary targets of QLs: DNA gyrase and topoisomerase IV some fragmentation was eliminated after the drug removed or by
mild heat treatment (60  C) [34]. Addition of a protein denaturant
The first QL, nalidixic acid, was discovered with antimicrobial such as sodium dodecyl sulfate (SDS) to the ternary complexes
properties in the by-product distillates in the manufacture of the would release DNA ends [14,15,29,34]. Nevertheless, the drugs
anti-malarial chloroquine [9], then ten times more potent QL, cannot cause generation of free DNA ends inside bacterial cells
oxolinic acid, was synthesized five years later [10]. Goss et al. [11,12] since intact nucleoids containing negatively supercoiled DNA could
first showed that nalidixic acid was a selective, immediate and be isolated from oxolinic acid treated bacteria in a cell lysis pro-
reversible inhibitor of bacterial DNA synthesis. Soon after the dis- cedure without SDS [34,35].
covery of bacterial gyrase in 1976 [13], both Gellert and Cozzarelli Comparison of the nucleotide sequences at the cleavage sites
[14,15] demonstrated that the supercoiling activity of the purified reveals a loose consensus sequence [36e39]. Not all of the cleavage
gyrase, extracted from wild-type cells but not from resistant nalA sites identified on a given DNA molecule were cleaved when oxo-
mutants, was inhibited by nalidixic acid and oxolinic acids. In 1978, linic acid is added to cells [34]. Cleavage is especially frequent at a
Higgins et al. [16] confirmed that subunit A of DNA gyrase (GyrA) small number of specific sites called toposites on the chromosome
was the product of the gene controlling sensitivity to nalidixic acid. [40]. A large number of weaker sites also exist [41]. It is speculated
These results suggest that DNA gyrase is the primary target of QLs. that the small number of strong interaction sites are used by gyrase
Gyrase, belonging to type IIA topoisomerases, functions as a to maintain superhelical tension in the chromosome as a whole and
heterotetramer composed of two GyrA subunits and two GyrB the weak and widely dispersed sites allow gyrase to provide local
subunits [17]. Gyrase’s enzymatic activity is essential for the swiveling needed for transcription and replication.

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8 3

3.2. Crystal structures of cleaved complexes which has a basic substituent at the C7 position [46], providing a
structural rationalization for the effect of a ParE/GyrB mutation on
To date, a few of crystal structures of complexes between type QL sensitivity. Nearly at the same time, a structure of S. pneumoniae
IIA topoisomerases and DNA have been resolved. Crystal structures topo IV with DNA and levofloxacin was reported [47] (Fig. 2B),
of a complex consisting of the DNA-binding core of Saccharomyces presenting a similar fluoroquinolone orientation as in QL-bound
cerevisiae Topoisomerase II and a gate-DNA segment [42] and a noncatalytic magnesium model [45], but without the QL-bound
cleaved complex composed of the DNA-binding core of Mg2þ ion.
S. pneumoniae topo IV with broken DNA and either clinafloxacin or In this QL-bound noncatalytic magnesium model, QL orientation
moxifloxacin [43] have been comprehensively elucidated in the is likely to be conserved, as it explains the known involvement of
review written by Drlica et al. [44]. However, the effects of several Mg2þ in QL action, QL structureeactivity relationships and resis-
amino acid substitutions that alter the antimicrobial activity and tance mutations across orthologs [45]. Since the C-7 rings are still far
drug structure variations cannot be explained by these structures. from GyrA helix IV in this model, it is not easy to explain the influ-
Recently, a crystal structure of moxifloxacin in complex with ence of the C-7 ring structure of fluoroquinolones on the increased
Acinetobacter baumannii topo IV has been reported, demonstrating MIC of GyrA G89C (Mycobacterium smegmatis numbering, equiva-
the wedge-shaped QL stacking between base pairs at the DNA lent to G81C in E. coli numbering) [48] or G81D (E. coli numbering)
cleavage site and binding conserved residues in the DNA cleavage variant [49]. C-7 ring structure may have effects on binding that are
domain through chelation of a noncatalytic magnesium ion [45] not reflected by the static nature of the X-ray structures, since QL
(Fig. 2A). In this QL-bound noncatalytic magnesium model, two binding is known to involve at least two steps, one that occurs before
moxifloxacin molecules intercalate four base pairs apart at the two DNA cleavage and one that occurs after [50,51].
sites of DNA cleavage. The QL interaction with DNA is largely
through van der Waals, pep stacking interactions, as reported in 3.3. Replication fork arrest and bacteriostasis
the crystal structure with S. pneumoniae topo IV [43], but this
higher-resolution structure reveals a different orientation and Trapping topoisomerases on DNA can lead to a rapid inhibition
binding mode for moxifloxacin, with a Mg2þ mediating the inter- of DNA replication [11], since trapping gyrase in a QLegyraseeDNA
action between QL and the protein. complex ahead of the replication fork would block fork movement.
The QL-bound noncatalytic Mg2þ, some 14  A from Mg2þ at the Cleaved complexes also form with topo IV [29], but the rate at
active site, is modeled with an octahedral coordination sphere with which replication is inhibited is 50e100 times slower than with
two oxygens from the QL and four water molecules [45]. Two of the gyrase in E. coli [25]. The difference is generally explained by topo
Mg2þ coordinating water molecules make hydrogen bonds with IV functioning behind replication forks [52], while gyrase is prob-
ParC Ser84 and Glu88, thus providing a structural explanation for ably ahead of them [52,53]. In Gram-positive bacteria, such as
why mutations in these residues in DNA gyrase and topo IV confer S. aureus, in vitro reactions with norfloxacin and gyrase fail to block
QL resistance in diverse bacterial species. Mg2þ coordination also replication fork movement [54], explaining why norfloxacin in-
explains why the carbonyl at C4 of a QL is essential as well as why hibits DNA replication largely through topo IV in S. aureus [55].
only limited substitution is permitted at positions 3 and 4. The The stability of QL-induced cleaved complexes contributes to
bulky substituent at position 7 of moxifloxacin occupies a large and their ability to block replication fork movement. It was shown that
open pocket between the DNA and the ParE subunit, consistent S. aureus gyrase wrapped DNA to catalyze DNA supercoiling and
with the C7 position being the most versatile position for substi- arrest replication fork progression in vitro [54]. E. coli mutant gyrase
tution in QLs. The C7 substituent is proximal to Arg418 from ParE. In with lack of the entire GyrA CTD (GyrA59) cannot wrap DNA,
E. coli, mutation of the equivalent residue Lys447 to glutamate showing that gyrase-mediated DNA wrapping is required for
enhances the activity of amphoteric QLs, such as moxifloxacin, replication fork arrest [56]. Although topo IV does not wrap DNA,

Fig. 2. Mg2þ ion mediates QL interactions with topo IV. (A) Moxifloxacin (yellow) in complex with ParE28eParC58 (gray carbons) and DNA (green carbons) at 3.27 Å. Mg2þ (green
sphere) is coordinated by four water molecules (red spheres) and two oxygen atoms from the QL. (B) Cartoon/stick representation of the drug-binding pockets of topo IV-DNA in
complex with levofloxacin (magenta). The figure has been adapted with modifications from Refs. [45,47]. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
4 G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8

QL-induced cleaved complexes formed with topo IV are more stable


than those formed with GyrA59 gyrase [56]. Helix-4, part of the
DNA-binding domain of GyrA or ParC, is also important on complex
stability. Replacement of helix-4 of E. coli ParC with that of E. coli
GyrA [57] and replacement of an extended region around helix-4 of
E. coli GyrA with that of S. aureus GyrA [58] reveal that subtle dif-
ferences of helix-4 affect the QL sensitivity of a topoisomerase.
Thus, QL sensitivity of a topoisomerase correlates with the stability
of the cleaved complex.
Formation of cleaved complexes, inhibition of DNA synthesis,
and inhibition of growth are correlated [34,59]. However, they are
all reversible phenomena, while lethal action is not [12,60].
Nevertheless, blocked replication forks could stimulate secondary
events that kill cells slowly. When drug treatment is long (over-
night with E. coli), cell death can occur at QL concentrations that are
only twice MIC (rapid killing generally requires concentrations 5e
10 times the MIC). Slow killing, which is commonly expressed by
the parameter called the minimal bactericidal concentration (MBC),
is poorly understood mechanistically [5].
A few double-strand DNA breaks (DSBs) are generated by
collision of replication forks with cleaved complexes, since in a
plasmid model system where cleaved complex formation blocked
replication, double-strand breaks were observed [61]. It has been
suggested that DSBs arising after replication fork stalling are
generated by a recombination nuclease. One candidate is the
RuvABC complex, as RuvAB may reverse and dissociate the cleaved
complex at a stalled replication fork before RuvC cleaves DNA to
generate a DSB [61]. DNA cleavage could provide RecBC with access Fig. 3. Schematic representation of QL action. (a) Binding of gyrase to DNA.
to circular DNA, as required for the induction of SOS response [62]. (b) Reversible formation of QLegyraseeDNA complexes that rapidly block DNA repli-
cation. (c) Inhibition of replication leads to induction of the SOS response and cell
It is important to distinguish the few DNA breaks associated with filamentation. (d) Lethal chromosome fragmentation that requires ongoing protein
replication fork arrest from the extensive chromosome fragmen- synthesis in aerobic conditions, as seen with nalidixic acid treatment of E. coli.
tation associated with cell death (discussed below). (e) Lethal chromosome fragmentation that requires ongoing protein synthesis but not
aerobic conditions, as seen with norfloxacin treatment of E. coli. (f) Lethal chromosome
fragmentation that requires neither ongoing protein synthesis nor aerobic conditions,
4. Lethal actions of QLs
as seen for PD161144 with E. coli. (g) DNA breakage detected after treatment of cell
lysates with an ionic detergent such as SDS. (h) Chromosome fragmentation stimulates
4.1. Pathways of chromosome fragmentation a surge of ROS with hydroxyl radicals as the terminal product. (i) Formation of highly
toxic hydroxyl radicals, the end product of an ROS cascade, is blocked by 2,20 -bipyridyl,
Cell death arises from the release of DNA breaks resulting in an iron chelator, and thiourea, a hydroxyl radical scavenger, which prevent cell death.
Question marks indicate uncertainty about slow death and the nature of the DNA ends.
chromosome fragmentation from protein-mediated constraint The figure has been drawn based on Refs. [5,74].
existing in the cleaved complexes [23]. The QLs kill E. coli in diverse
pathways, distinguished by the effects of protein synthesis inhibitors
and anaerobiosis on QL lethality [7]. Nalidixic and oxolinic acids of with chloramphenicol [23]. Thus, it appears that an identified
the first-generation QLs are not lethal in the presence of chloram- protein factor is involved in releasing DNA ends from the QLe
phenicol or during anaerobic growth [20,23] (Fig. 3, pathway d). gyraseeDNA complexes in three suggested mechanisms: 1) prote-
Norfloxacin which is the second-generation QL fails to kill E. coli in ase digestion of gyrase, 2) nuclease-mediated cleavage on either
the presence of chloramphenicol, but at high concentrations it kills side of the cleaved complex, and 3) protein denaturation [44]. On
cells under anaerobic condition (Fig. 3, pathway e); while cipro- the other hand, lethal effect of chloramphenicol-insensitive mode
floxacin kills under both conditions but requires higher concentra- also arises from the release of DNA ends from QLegyraseeDNA
tions during anaerobiosis [20,23,63] (Fig. 3, pathway f). PD161144, a complexes [23]. An explanation for the protein synthesis-
third generation C-8-methoxy derivative, is affected little by chlor- independent mode emerged from the observation that the QLs
amphenicol or anaerobic growth for the lethal activity [63] (Fig. 3, stimulated a form of illegitimate recombination and deletion for-
pathway f). Since even the fourth-generation compounds are sensi- mation arising from subunit dissociationereassociation [67e71],
tive to chloramphenicol if QL concentrations are low enough, the suggesting that QL molecules might be able to force gyraseeDNA
choice of pathway depends on QL concentration. The different lethal complexes apart (Fig. 3, step f). An E. coli mutant with A67S in
pathways are also observed in mycobacteria [64,65]. GryA allowed nalidixic acid to kill E. coli and fragment chromo-
Much less was said about cell death arising from QL-topo somes in the presence of chloramphenicol [64]. Thus, the mutation
IV-DNA lesions. From the observation that killing of an E. coli gyrA appears to shift lethal activity from step d to step f in Fig. 3, perhaps
(Nalr) mutant was drastically reduced by rifampin [20] or chlor- by the alanine-to-serine change weakening hydrophobic in-
amphenicol [23], it is proved a protein synthesis-dependent mode. teractions between the GyrA subunits. Lethal effects arising from
In the case of C-8-methoxy fluoroquinolones (PD161148), the this chloramphenicol-insensitive mode of QL action become more
methoxy group increased lethality against wild-type S. aureus cells prominent as the QL concentration increases [23], perhaps by a
when protein synthesis was inhibited [66]. cooperative effect of drug stacking on gyrase subunit dissociation.
As for the mechanism of protein synthesis-dependent pathway, Furthermore, the effect of QL structure may explain the connection
it was shown that oxolinic acid failed to eliminate the topological between cleaved complex destabilization and chloramphenicol-
constraint needed for the maintenance of supercoils in cells treated insensitive killing [5].

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8 5

4.2. Oxidative damage cellular death pathway involved in a variety of activities, including DNA repair, recombi-
nation and mutagenesis [77,78]. RecBCD complex exhibit a single-
Recent findings indicate that production of ROS is a common strand (ss) DNA exonuclease activity associated with the genera-
mechanism of cell death induced by bactericidal antibiotics, tion of free DNA ends followed by endonucleolytic activities of the
including QLs [6,72]. Hydroxyl radical concentrations were elevated RuvABC or SbcCD complex of E. coli [79]. Free ssDNA ends are bound
in E. coli following treatment with norfloxacin. Thiourea or 2,20 - by RecA molecules to form nucleoprotein filaments, activating the
bipyridyl, agents that reduce the level of hydroxyl radical, inhibited auto-self-degradation activity of the LexA repressor. As a conse-
norfloxacin lethality [72] (Fig. 3, step i). When both sodA and sodB quence, the complete SOS response is induced [80].
were deficient, norfloxacin lethality was reduced, in consistence QLs appear to produce at least two types of lethal damages that
with that superoxide dismutase normally promoted QL lethality by can be distinguished by protective elements of the SOS response.
stimulating formation of peroxide [73]. A deficiency in catalase/ For some QLs, such as nalidixic acid, survival depends largely on the
peroxidase (katG) elevated the lethal activity of norfloxacin, recombination function of the RecA and RecBCD proteins [81,82].
because a buildup of peroxide lead to accumulation of highly toxic Moreover, nalidixic acid lethality is unaffected by the lexA mutation
hydroxyl radicals [73]. in a gyrAþ background [64]. However, for the potent fluo-
How do protein synthesis-dependent and -independent path- roquinolones (e.g. norfloxacin and ciprofloxacin), rapid lethality is
ways fit with the observation that hydroxyl radical accumulation is increased by a lexA allele [83]. Since norfloxacin and ciprofloxacin
associated with QL lethality? It was recently demonstrated that are able to attack topo IV, and recA and lexA mutations can increase
lethal activity of oxolinic acid was completely blocked by thiourea the ciprofloxacin susceptibility of gyrA mutants [23], topo IV-
plus 2,20 -bipyridyl and by chloramphenicol, but only chloram- mediated killing appears to be mitigated by both recA-dependent
phenicol blocked both chromosome fragmentation and hydroxyl recombination and SOS induction.
radical accumulation [74]. Thus, the chromosome fragmentation Induction of the SOS response not only causes QL-induced
step occurs before the ROS step (Fig. 3, step h). In contrast, lethal mutagenesis, contributing to the acquisition of QL resistant muta-
action of PD161144 by the subunit dissociation pathway (Fig. 3, step tions, but also enhances bacterial survival in the presence of fluo-
f) was affected little by treatment with chloramphenicol or thiourea roquinolones. sfiA encodes an inhibitor of cell division that function
plus 2,20 -bipyridyl or all three agents together [74]. With moxi- to allow time for the repair of DNA damage, and recA encodes the
floxacin, both chloramphenicol and thiourea plus 2,20 -bipyridyl key factor in homologous recombination (HR) reactions [84].
separately exhibited the same partial inhibition of QL lethality, and Additionally, genes umuCD are activated to express low-fidelity
no additivity in protection from moxifloxacin lethality was DNA polymerase V, which is involved in SOS-mediated mutagen-
observed when these three agents were combined [74]. These esis by inducing the accumulation of replication errors in surviving
studies indicated that hydroxyl radical action contributes to QL- cells. dinB encoding DNA polymerase IV is activated to restart DNA
mediated cell death occurring via the chloramphenicol-sensitive replication behind DSBs which is called translesion synthesis (TLS)
but not via the chloramphenicol-insensitive lethal pathway. and repair independently from other DNA polymerases [85]. Cells
may benefit from this enhanced mutational activity by ensuring
4.3. Toxineantitoxin and programmed cell death (PCD) long-term survival as well as evolutionary fitness [78]. It is shown
that the development of fluoroquinolone resistance is impaired in
Toxineantitoxin modules, which appear to be part of the bac- the absence of SOS response [86]. Thus, the error-prone activity of
terial stress responses, also contribute to QL lethality. These mod- the SOS response is considered to be an intrinsic system of bacteria
ules act when elimination of a short-lived antitoxin allows the for adaptive response to DNA damaging stress in those survival
cognate toxin to interfere with the bacterial transcriptione cells.
translation machinery, leading to programmed cell death (PCD), Besides SOS response, which preferentially repairs DSBs via
an irreversible series of events that causes the loss of colony- error-free HR or tolerates them via error-prone TLS, alternative
forming ability even after removal of the stressor [75]. When cells pathways resulting in large chromosomal rearrangements also
were treated with nalidixic acid at 500 to 1000 times the MIC for exist. Illegitimate recombination is accomplished by subunit ex-
10 min, a 90% decrease in the number of CFU followed unless the change between two type II topoisomerase holoenzymes bound in
chpAIK (mazEF) toxineantitoxin module was absent, suggesting two separate cleavage complexes, which results in the ligation of
that triggering the E. coli chromosomal toxineantitoxin system DNA ends from the two separate DSBs [70]. Non-homologous
mazEF was an additional determinant in the lethal action of QLs [8]. end joining (NHEJ) has also been reported in mycobacteria as an
Further study is needed to examine whether the action of alternative repair pathway for DSBs by religating modified DNA
mazEF-mediated cell death in E. coli involve the production of ROS, ends [87].
which is a common mechanism of cell death induced by bacteri-
cidal antibiotics [72]. It was found that antibiotics as DNA damaging 5.2. Other QL-induced stress responses
reagents (i.e. nalidixic acid) caused mazEF-mediated cell death in an
ROS-independent pathway [76]. Furthermore, a signaling molecule, Bacteria can survive antibiotic treatment without acquiring
Extracellular Death Factor (EDF), participated in mazEF induction. heritable antibiotic resistance, but to induce stress responses that
This mode of action implies that using synthetic EDF together with protect the cell from lethal factors such as DNA damaging agents.
QLs can increase the efficiency of bacterial killing. Besides, bacte- Several TA (toxin/antitoxin) genes were induced in E. coli in
riostatic antibiotics can be turned into bactericidal antibiotics by response to ciprofloxacin, whose promoters contain a Lex box:
using EDF to turn on the mazEF system. symER, hokE, yafN/yafO, and tisAB/istR, which is necessary for
persister formation (dormant cells formation) [88]. The mechanism
5. Bacterial responses to QLs of this ciprofloxacin-induced persister formation was recently
investigated. It was shown that a knockout of tisAB/istR had a
5.1. SOS response of bacteria sharply decreased level of persisters tolerant to ciprofloxacin, and
step-wise administration of ciprofloxacin-induced persister for-
One consequence of QL treatment is the induction of the SOS mation in a tisAB-dependent manner. Moreover, cells producing
regulon, a set of more than 30 lexA repressor-controlling genes TisB toxin were tolerant to multiple antibiotics [89]. TisB is a

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
6 G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8

membrane peptide that functions to decrease proton motive force induce irreversible oxidative DNA damage by ROS that occurs in a
and ATP levels, consistent with its role in forming dormant cells. protein synthesis-dependent manner. Besides, the toxin MazF is
Lon protease degrades abnormal proteins and proteins pro- required under certain conditions for efficient killing by QLs with
duced in excess, influencing a variety of physiological phenomena EDF participating in mazEF induction. In contrast to the killing ef-
[90]. The role of Lon protease in chromosome maintenance was fect of QLs, bacteria can survive and acquire resistance by SOS
firstly noticed when examining paradoxical survival of bacteria at response and other stress responses.
very high concentrations of QLs [91]. A deficiency of Lon protease Nevertheless, our knowledges about QL actions are far from
eliminated paradoxical survival, and plasmid-borne protease ac- complete and needed to be explored. For example, (1) the already
tivity of Lon restored the paradoxical behavior of QLs [92]. These resolved crystal structures of the drugeenzymeeDNA complexes
observations confirm that Lon is necessary for paradoxical survival. still cannot explained several amino acid substitutions and drug
It was also found that a Lon-deficient mutant, PA1803, exhibited structure variations, additional studies should be required to
more than 4-fold-increased susceptibilities to fluoroquinolones, explain what is likely to be a multistep binding [99]; (2) Nor-
and complementation of the lon mutant restored wild-type anti- floxacin appears to represent an intermediate situation (Fig. 3, step
biotic susceptibility and cell morphology [93]. The data suggest that e), how norfloxacin-mediated replication fork breakage fits into
the induction of Lon is involved in adaptive resistance of bacteria. these categories is not currently understood; (3) Proteins involved
The precise roles of Lon protease are described in the review in steps d and e (Fig. 3) need to be identified to show how DNA
written by Drlica et al. [44]. breaks are released from the protein-mediated constraint, and
YrbB, a component of an ABC transport system that maintains cell-free tests for step f need to be extended to plasmid systems
lipid asymmetry in the Gram-negative outer membrane by pre- that can be readily studied; (4) It is clear that death ultimately
venting surface exposure of phospholipids, was found to be one of arises from the accumulation of hydroxyl radicals with the older
the proteins involved in protecting E. coli from QL-mediated death QLs, but how DNA breaks promote a cascade of ROS is still un-
by genetic searching [94]. A yrbB mutation rendered E. coli cells known; (5) Blocked replication forks could stimulate secondary
more susceptible to the lethal action of QLs under conditions in events that kill cells slowly with QL concentrations that are only
which bacteriostatic susceptibility was unaffected. YrbB worked in twice MIC, and the mechanism of slow killing is poorly under-
both protein synthesis-dependent and -independent lethal path- stood; (6) The paradoxical loss of lethality at very high QL con-
ways of QL action. In the absence of YrbB, nalidixic acid could kill centrations need to be explained in molecular terms; (7) Whether
cells independently of hydroxyl radical action, in which the E. coli Lon-mediated repair involves direct removal of the complexes or
toxineantitoxin system ChpB was found to be involved. In addition, an indirect effect due to rapidly removing an unidentified lethal
proteases ClpP and Lon were also involved in the action of factor involved in fragmentation of DNA is not known; (8) Those
YrbB [95]. genes and proteins involved in bacterial response contributing to
For some time the fluoroquinolones have been suspected to survival and resistance should be identified and studied in mo-
have a secondary mode of action involving membrane damage lecular terms, which would serve as potential targets for designing
[96,97]. By using Bacillus subtilis biosensors, both inhibitions of DNA enhancers for QLs.
synthesis and membrane damage by ciprofloxacin were detected Although QLs are broad-spectrum agents, they have encoun-
[98]. The yorB promoter indicating DNA synthesis and the ypuA tered resistance that makes new topoisomerase inhibitors attrac-
promoter indicating cell wall synthesis/cell envelope stress were tive. Among the newer QLs under development, delafloxacin,
induced by ciprofloxacin, supporting a weak secondary mode of nemonoxacin and JNJ-Q2 are with enhanced activity against Gram-
action involving membrane damage. Probably, ROS is one of the positive bacteria, including MDR (multiple drug resistance)
factors contributing to the membrane damage. Other mechanisms S. pneumoniae and ciprofloxacin-resistant MRSA (Methicillin-
involved in membrane damage, if there is any, still need to be resistant S. aureus). However, the molecular mechanism underlying
examined. the anti-resistant activity of these newer QLs has not been eluci-
The diverse set of genes protecting against the lethal effects dated yet. As reviewed here, the available crystal structures of the
of QLs may serve as potential targets for future screening of QL-type IIA topoisomeraseeDNA complexes have explained some
small-molecular antimicrobial potentiators, which can be used as determinants for resistance, and it is of great interest to study the
enhancers for QLs. anti-resistant activity of these newer QLs on the structural basis.
Moreover, the biological events in bacteria under QL pressures
6. Conclusions and prospective contribute to the cell death and development of resistance of bac-
teria, providing plenty of novel targets for drug design to improve
The QLs are of an important class of antimicrobial agents. It is activity and tackle resistance. As expected, the growing under-
clear that formation of drugeenzymeeDNA complexes is the cen- standing of mechanisms of QL actions will give insights into new
tral event in QL action. QL lethality can be described as a two step clinical treatment and approaches for fighting threat from QL
process in which the first step is reversible formation of cleaved resistant pathogens.
complexes (bacteriostatic). This step blocks bacterial DNA replica-
tion, induces the SOS response, and leads to cell filamentation
(Fig. 3, pathway c). Although these events do not appear to be
Declaration of interest
rapidly lethal, their involvement in slow death has not been ruled
out. The second lethal step requires higher QL concentrations, and
The authors report no declarations of interest.
DNA breaks are released from constraint by two pathways, one
requires protein synthesis (Fig. 3, steps d and e) and the other does
not (Fig. 3, step f). Each pathway to cell death depends on QLs
structures. The protein synthesis-dependent pathway involves Acknowledgments
removal of gyraseedrug complexes from DNA and liberation of
lethal double-strand DNA breaks, while the protein synthesis- This work was supported by National High Technology
independent pathway involves gyrase subunit dissociation and Research and Development Program of China (grant number
release of DNA ends with the gyrase subunits attached. DNA breaks 2011AA10A214).

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8 7

References [29] H. Peng, K.J. Marians, Escherichia coli topoisomerase IV. Purification, charac-
terization, subunit structure, and subunit interactions, J. Biol. Chem. 268
(1993) 24481e24490.
[1] G.M. Keating, L.J. Scott, Moxifloxacin: a review of its use in the management of
[30] F. Blanche, B. Cameron, F.X. Bernard, L. Maton, B. Manse, L. Ferrero, N. Ratet,
bacterial infections, Drugs 64 (2004) 2347e2377.
C. Lecoq, A. Goniot, D. Bisch, J. Crouzet, Differential behaviors of Staphylo-
[2] K. Mdluli, Z. Ma, Mycobacterium tuberculosis DNA gyrase as a target for drug
coccus aureus and Escherichia coli type II DNA topoisomerases, Antimicrob.
discovery, Infect. Disord. Drug Targets 7 (2007) 159e168.
Agents Chemother. 40 (1996) 2714e2720.
[3] T.M. File Jr., G.S. Tillotson, Gemifloxacin: a new, potent fluoroquinolone for the
[31] R. Munoz, A.G. De La Campa, ParC subunit of DNA topoisomerase IV of
therapy of lower respiratory tract infections, Expert Rev. Anti. Infect. Ther. 2
Streptococcus pneumoniae is a primary target of fluoroquinolones and co-
(2004) 831e843.
operates with DNA gyrase A subunit in forming resistance phenotype, Anti-
[4] P.E. Donahue, D.L. Smith, A.E. Yellin, S.J. Mintz, F. Bur, D.R. Luke, Trovafloxacin
microb. Agents Chemother. 40 (1996) 2252e2257.
in the treatment of intra-abdominal infections: results of a double-blind,
[32] X.S. Pan, J. Ambler, S. Mehtar, L.M. Fisher, Involvement of topoisomerase IV
multicenter comparison with imipenem/cilastatin. Trovafloxacin Surgical
and DNA gyrase as ciprofloxacin targets in Streptococcus pneumoniae, Anti-
Group, Am. J. Surg. 176 (1998) 53Se61S.
microb. Agents Chemother. 40 (1996) 2321e2326.
[5] K. Drlica, M. Malik, R.J. Kerns, X. Zhao, Quinolone-mediated bacterial death,
[33] G. Orphanides, A. Maxwell, Evidence for a conformational change in the DNA
Antimicrob. Agents Chemother. 52 (2008) 385e392.
gyraseeDNA complex from hydroxyl radical footprinting, Nucleic Acids Res.
[6] D.J. Dwyer, M.A. Kohanski, B. Hayete, J.J. Collins, Gyrase inhibitors induce an
22 (1994) 1567e1575.
oxidative damage cellular death pathway in Escherichia coli, Mol. Syst. Biol. 3
[34] M. Snyder, K. Drlica, DNA gyrase on the bacterial chromosome: DNA cleavage
(2007) 91.
induced by oxolinic acid, J. Mol. Biol. 131 (1979) 287e302.
[7] M. Malik, S. Hussain, K. Drlica, Effect of anaerobic growth on quinolone lethality
[35] S.H. Manes, G.J. Pruss, K. Drlica, Inhibition of RNA synthesis by oxolinic acid is
with Escherichia coli, Antimicrob. Agents Chemother. 51 (2007) 28e34.
unrelated to average DNA supercoiling, J. Bacteriol. 155 (1983) 420e423.
[8] I. Kolodkin-Gal, H. Engelberg-Kulka, Induction of Escherichia coli chromosomal
[36] A. Morrison, N.R. Cozzarelli, Site-specific cleavage of DNA by E. coli DNA
mazEF by stressful conditions causes an irreversible loss of viability,
gyrase, Cell 17 (1979) 175e184.
J. Bacteriol. 188 (2006) 3420e3423.
[37] L.M. Fisher, K. Mizuuchi, M.H. O’Dea, H. Ohmori, M. Gellert, Site-specific
[9] G.Y. Lesher, E.J. Froelich, M.D. Gruett, J.H. Bailey, R.P. Brundage, 1,8-
interaction of DNA gyrase with DNA, Proc. Natl. Acad. Sci. U. S. A. 78 (1981)
Naphthyridine derivatives. A new class of chemotherapeutic agents, J. Med.
4165e4169.
Pharm. Chem. 91 (1962) 1063e1065.
[38] K. Kirkegaard, J.C. Wang, Mapping the topography of DNA wrapped around
[10] F.J. Turner, S.M. Ringel, J.F. Martin, P.J. Storino, J.M. Daly, B.S. Schwartz, Oxo-
gyrase by nucleolytic and chemical probing of complexes of unique DNA se-
linic acid, a new synthetic antimicrobial agent. I. In vitro and in vivo activity,
quences, Cell 23 (1981) 721e729.
Antimicrob. Agents Chemother. 7 (1967) 475e479.
[39] D. Lockshon, D.R. Morris, Sites of reaction of Escherichia coli DNA gyrase on
[11] W.A. Goss, W.H. Deitz, T.M. Cook, Mechanism of action of nalidixic acid on
pBR322 in vivo as revealed by oxolinic acid-induced plasmid linearization,
Escherichia coli, J. Bacteriol. 88 (1964) 1112e1118.
J. Mol. Biol. 181 (1985) 63e74.
[12] W.A. Goss, W.H. Deitz, T.M. Cook, Mechanism of action of nalidixic acid on
[40] G. Condemine, C.L. Smith, Transcription regulates oxolinic acid-induced DNA
Escherichia coli. Ii. Inhibition of deoxyribonucleic acid synthesis, J. Bacteriol. 89
gyrase cleavage at specific sites on the E. coli chromosome, Nucleic Acids Res.
(1965) 1068e1074.
18 (1990) 7389e7396.
[13] M. Gellert, K. Mizuuchi, M.H. O’Dea, H.A. Nash, DNA gyrase: an enzyme that
[41] R.J. Franco, K. Drlica, DNA gyrase on the bacterial chromosome. Oxolinic
introduces superhelical turns into DNA, Proc. Natl. Acad. Sci. U. S. A. 73 (1976)
acid-induced DNA cleavage in the dnaA-gyrB region, J. Mol. Biol. 201 (1988)
3872e3876.
229e233.
[14] M. Gellert, K. Mizuuchi, M.H. O’Dea, T. Itoh, J.I. Tomizawa, Nalidixic acid
[42] K.C. Dong, J.M. Berger, Structural basis for gate-DNA recognition and bending
resistance: a second genetic character involved in DNA gyrase activity, Proc.
by type IIA topoisomerases, Nature 450 (2007) 1201e1205.
Natl. Acad. Sci. U. S. A. 74 (1977) 4772e4776.
[43] I. Laponogov, M.K. Sohi, D.A. Veselkov, X.S. Pan, R. Sawhney, A.W. Thompson,
[15] A. Sugino, C.L. Peebles, K.N. Kreuzer, N.R. Cozzarelli, Mechanism of action of
K.E. McAuley, L.M. Fisher, M.R. Sanderson, Structural insight into the
nalidixic acid: purification of Escherichia coli nalA gene product and its rela-
quinolone-DNA cleavage complex of type IIA topoisomerases, Nat. Struct. Mol.
tionship to DNA gyrase and a novel nicking-closing enzyme, Proc. Natl. Acad.
Biol. 16 (2009) 667e669.
Sci. U. S. A. 74 (1977) 4767e4771.
[44] K. Drlica, H. Hiasa, R. Kerns, M. Malik, A. Mustaev, X. Zhao, Quinolones: action
[16] N.P. Higgins, C.L. Peebles, A. Sugino, N.R. Cozzarelli, Purification of subunits of
and resistance updated, Curr. Top. Med. Chem. 9 (2009) 981e998.
Escherichia coli DNA gyrase and reconstitution of enzymatic activity, Proc.
[45] A. Wohlkonig, P.F. Chan, A.P. Fosberry, P. Homes, J. Huang, M. Kranz,
Natl. Acad. Sci. U. S. A. 75 (1978) 1773e1777.
V.R. Leydon, T.J. Miles, N.D. Pearson, R.L. Perera, A.J. Shillings, M.N. Gwynn,
[17] Y. Pommier, E. Leo, H. Zhang, C. Marchand, DNA topoisomerases and their
B.D. Bax, Structural basis of quinolone inhibition of type IIA topoisomerases
poisoning by anticancer and antibacterial drugs, Chem. Biol. 17 (2010)
and target-mediated resistance, Nat. Struct. Mol. Biol. 17 (2010) 1152e1153.
421e433.
[46] H. Yoshida, M. Bogaki, M. Nakamura, L.M. Yamanaka, S. Nakamura, Quinolone
[18] J. Kato, Y. Nishimura, R. Imamura, H. Niki, S. Hiraga, H. Suzuki, New topo-
resistance-determining region in the DNA gyrase gyrB gene of Escherichia coli,
isomerase essential for chromosome segregation in E. coli, Cell 63 (1990)
Antimicrob. Agents Chemother. 35 (1991) 1647e1650.
393e404.
[47] I. Laponogov, X.S. Pan, D.A. Veselkov, K.E. McAuley, L.M. Fisher,
[19] J. Kato, H. Suzuki, H. Ikeda, Purification and characterization of DNA topo-
M.R. Sanderson, Structural basis of gate-DNA breakage and resealing by type II
isomerase IV in Escherichia coli, J. Biol. Chem. 267 (1992) 25676e25684.
topoisomerases, PLoS One 5 (2010) e11338.
[20] C.S. Lewin, B.M. Howard, J.T. Smith, Protein- and RNA-synthesis independent
[48] M. Malik, K.R. Marks, A. Mustaev, X. Zhao, K. Chavda, R.J. Kerns, K. Drlica,
bactericidal activity of ciprofloxacin that involves the A subunit of DNA gyr-
Fluoroquinolone and quinazolinedione activities against wild-type and gyrase
ase, J. Med. Microbiol. 34 (1991) 19e22.
mutant strains of Mycobacterium smegmatis, Antimicrob. Agents Chemother.
[21] M. Trucksis, J.S. Wolfson, D.C. Hooper, A novel locus conferring fluo-
55 (2011) 2335e2343.
roquinolone resistance in Staphylococcus aureus, J. Bacteriol. 173 (1991)
[49] E. Cambau, F. Bordon, E. Collatz, L. Gutmann, Novel gyrA point mutation in a
5854e5860.
strain of Escherichia coli resistant to fluoroquinolones but not to nalidixic acid,
[22] C.J. Soussy, J.S. Wolfson, E.Y. Ng, D.C. Hooper, Limitations of plasmid
Antimicrob. Agents Chemother. 37 (1993) 1247e1252.
complementation test for determination of quinolone resistance due to
[50] S.E. Critchlow, A. Maxwell, DNA cleavage is not required for the binding of
changes in the gyrase A protein and identification of conditional quinolone
quinolone drugs to the DNA gyraseeDNA complex, Biochemistry 35 (1996)
resistance locus, Antimicrob. Agents Chemother. 37 (1993) 2588e2592.
7387e7393.
[23] C.R. Chen, M. Malik, M. Snyder, K. Drlica, DNA gyrase and topoisomerase IV on
[51] K.J. Marians, H. Hiasa, Mechanism of quinolone action. A drug-induced
the bacterial chromosome: quinolone-induced DNA cleavage, J. Mol. Biol. 258
structural perturbation of the DNA precedes strand cleavage by topoisomer-
(1996) 627e637.
ase IV, J. Biol. Chem. 272 (1997) 9401e9409.
[24] P. Heisig, Genetic evidence for a role of parC mutations in development of
[52] A.B. Khodursky, N.R. Cozzarelli, The mechanism of inhibition of topoisomerase
high-level fluoroquinolone resistance in Escherichia coli, Antimicrob. Agents
IV by quinolone antibacterials, J. Biol. Chem. 273 (1998) 27668e27677.
Chemother. 40 (1996) 879e885.
[53] K. Drlica, E.C. Engle, S.H. Manes, DNA gyrase on the bacterial chromosome:
[25] A.B. Khodursky, E.L. Zechiedrich, N.R. Cozzarelli, Topoisomerase IV is a target
possibility of two levels of action, Proc. Natl. Acad. Sci. U. S. A. 77 (1980)
of quinolones in Escherichia coli, Proc. Natl. Acad. Sci. U. S. A. 92 (1995)
6879e6883.
11801e11805.
[54] H. Hiasa, M.E. Shea, C.M. Richardson, M.N. Gwynn, Staphylococcus aureus
[26] R.J. Belland, S.G. Morrison, C. Ison, W.M. Huang, Neisseria gonorrhoeae ac-
gyraseequinolone-DNA ternary complexes fail to arrest replication fork pro-
quires mutations in analogous regions of gyrA and parC in fluoroquinolone-
gression in vitro. Effects of salt on the DNA binding mode and the catalytic
resistant isolates, Mol. Microbiol. 14 (1994) 371e380.
activity of S. aureus gyrase, J. Biol. Chem. 278 (2003) 8861e8868.
[27] L. Ferrero, B. Cameron, B. Manse, D. Lagneaux, J. Crouzet, A. Famechon,
[55] B. Fournier, X. Zhao, T. Lu, K. Drlica, D.C. Hooper, Selective targeting of
F. Blanche, Cloning and primary structure of Staphylococcus aureus DNA
topoisomerase IV and DNA gyrase in Staphylococcus aureus: different patterns
topoisomerase IV: a primary target of fluoroquinolones, Mol. Microbiol. 13
of quinolone-induced inhibition of DNA synthesis, Antimicrob. Agents Che-
(1994) 641e653.
mother. 44 (2000) 2160e2165.
[28] K. Hoshino, A. Kitamura, I. Morrissey, K. Sato, J. Kato, H. Ikeda, Comparison of
[56] H. Hiasa, M.E. Shea, DNA gyrase-mediated wrapping of the DNA strand is
inhibition of Escherichia coli topoisomerase IV by quinolones with DNA gyrase
required for the replication fork arrest by the DNA gyraseequinoloneeDNA
inhibition, Antimicrob. Agents Chemother. 38 (1994) 2623e2627.
ternary complex, J. Biol. Chem. 275 (2000) 34780e34786.

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057
8 G. Cheng et al. / European Journal of Medicinal Chemistry xxx (2013) 1e8

[57] E.S. Pfeiffer, H. Hiasa, Replacement of ParC alpha4 helix with that of GyrA [77] G.C. Walker, Mutagenesis and inducible responses to deoxyribonucleic acid
increases the stability and cytotoxicity of topoisomerase IV-quinolone-DNA damage in Escherichia coli, Microbiol. Rev. 48 (1984) 60e93.
ternary complexes, Antimicrob. Agents Chemother. 48 (2004) 608e611. [78] W.L. Kelley, Lex marks the spot: the virulent side of SOS and a closer look at
[58] J. Strahilevitz, A. Robicsek, D.C. Hooper, Role of the extended alpha4 domain of the LexA regulon, Mol. Microbiol. 62 (2006) 1228e1238.
Staphylococcus aureus gyrase A protein in determining low sensitivity to [79] P. Heisig, Type II topoisomerases-inhibitors, repair mechanisms and muta-
quinolones, Antimicrob. Agents Chemother. 50 (2006) 600e606. tions, Mutagenesis 24 (2009) 465e469.
[59] R.T. Chow, T.J. Dougherty, H.S. Fraimow, E.Y. Bellin, M.H. Miller, Association [80] P.P. Khil, R.D. Camerini-Otero, Over 1000 genes are involved in the DNA
between early inhibition of DNA synthesis and the MICs and MBCs of car- damage response of Escherichia coli, Mol. Microbiol. 44 (2002) 89e105.
boxyquinolone antimicrobial agents for wild-type and mutant [gyrA [81] C.S. Lewin, B.M. Howard, N.T. Ratcliffe, J.T. Smith, 4-quinolones and the SOS
nfxB(ompF) acrA] Escherichia coli K-12, Antimicrob. Agents Chemother. 32 response, J. Med. Microbiol. 29 (1989) 139e144.
(1988) 1113e1118. [82] L.S. McDaniel, L.H. Rogers, W.E. Hill, Survival of recombination-deficient
[60] W.H. Deitz, T.M. Cook, W.A. Goss, Mechanism of action of nalidixic acid on mutants of Escherichia coli during incubation with nalidixic acid, J. Bacteriol.
Escherichia coli. 3. Conditions required for lethality, J. Bacteriol. 91 (1966) 134 (1978) 1195e1198.
768e773. [83] B.M. Howard, R.J. Pinney, J.T. Smith, Function of the SOS process in repair of
[61] J.R. Pohlhaus, K.N. Kreuzer, Norfloxacin-induced DNA gyrase cleavage com- DNA damage induced by modern 4-quinolones, J. Pharm. Pharmacol. 45
plexes block Escherichia coli replication forks, causing double-stranded breaks (1993) 658e662.
in vivo, Mol. Microbiol. 56 (2005) 1416e1429. [84] S.L. Lusetti, M.M. Cox, The bacterial RecA protein and the recombinational DNA
[62] K.G. Newmark, E.K. O’Reilly, J.R. Pohlhaus, K.N. Kreuzer, Genetic analysis of the repair of stalled replication forks, Annu. Rev. Biochem. 71 (2002) 71e100.
requirements for SOS induction by nalidixic acid in Escherichia coli, Gene 356 [85] W. Kuban, M. Banach-Orlowska, R.M. Schaaper, P. Jonczyk, I.J. Fijalkowska,
(2005) 69e76. Role of DNA polymerase IV in Escherichia coli SOS mutator activity, J. Bacteriol.
[63] M.A. Cooper, J.M. Andrews, R. Wise, Bactericidal activity of sparfloxacin and 188 (2006) 7977e7980.
ciprofloxacin under anaerobic conditions, J. Antimicrob. Chemother. 28 (1991) [86] R.T. Cirz, F.E. Romesberg, Induction and inhibition of ciprofloxacin resistance-
399e405. conferring mutations in hypermutator bacteria, Antimicrob. Agents Chemo-
[64] M. Malik, X. Zhao, K. Drlica, Lethal fragmentation of bacterial chromosomes ther. 50 (2006) 220e225.
mediated by DNA gyrase and quinolones, Mol. Microbiol. 61 (2006) 810e825. [87] G.R. Weller, B. Kysela, R. Roy, L.M. Tonkin, E. Scanlan, M. Della, S.K. Devine,
[65] S. Hussain, M. Malik, L. Shi, M.L. Gennaro, K. Drlica, In vitro model of myco- J.P. Day, A. Wilkinson, F. d’Adda di Fagagna, K.M. Devine, R.P. Bowater,
bacterial growth arrest using nitric oxide with limited air, Antimicrob. Agents P.A. Jeggo, S.P. Jackson, A.J. Doherty, Identification of a DNA nonhomologous
Chemother. 53 (2009) 157e161. end-joining complex in bacteria, Science 297 (2002) 1686e1689.
[66] X. Zhao, J.Y. Wang, C. Xu, Y. Dong, J. Zhou, J. Domagala, K. Drlica, Killing of [88] T. Dorr, K. Lewis, M. Vulic, SOS response induces persistence to fluo-
Staphylococcus aureus by C-8-methoxy fluoroquinolones, Antimicrob. Agents roquinolones in Escherichia coli, PLoS Genet. 5 (2009) e1000760.
Chemother. 42 (1998) 956e958. [89] T. Dorr, M. Vulic, K. Lewis, Ciprofloxacin causes persister formation by
[67] H. Ikeda, K. Aoki, A. Naito, Illegitimate recombination mediated in vitro by inducing the TisB toxin in Escherichia coli, PLoS Biol. 8 (2010) e1000317.
DNA gyrase of Escherichia coli: structure of recombinant DNA molecules, Proc. [90] V. Tsilibaris, G. Maenhaut-Michel, L. Van Melderen, Biological roles of the Lon
Natl. Acad. Sci. U. S. A. 79 (1982) 3724e3728. ATP-dependent protease, Res. Microbiol. 157 (2006) 701e713.
[68] H. Ikeda, I. Kawasaki, M. Gellert, Mechanism of illegitimate recombination: [91] G.C. Crumplin, J.T. Smith, Nalidixic acid: an antibacterial paradox, Antimicrob.
common sites for recombination and cleavage mediated by E. coli DNA gyrase, Agents Chemother. 8 (1975) 251e261.
Mol. Gen. Genet. 196 (1984) 546e549. [92] M. Malik, J. Capecci, K. Drlica, Lon protease is essential for paradoxical survival
[69] H. Ikeda, Illegitimate recombination: role of type II DNA topoisomerase, Adv. of Escherichia coli exposed to high concentrations of quinolone, Antimicrob.
Biophys. 21 (1986) 149e160. Agents Chemother. 53 (2009) 3103e3105.
[70] H. Ikeda, K. Shiraishi, Y. Ogata, Illegitimate recombination mediated by [93] M.D. Brazas, E.B. Breidenstein, J. Overhage, R.E. Hancock, Role of lon, an ATP-
double-strand break and end-joining in Escherichia coli, Adv. Biophys. 38 dependent protease homolog, in resistance of Pseudomonas aeruginosa to
(2004) 3e20. ciprofloxacin, Antimicrob. Agents Chemother. 51 (2007) 4276e4283.
[71] Y. Shanado, J. Kato, H. Ikeda, Escherichia coli HU protein suppresses DNA- [94] X. Han, A. Dorsey-Oresto, M. Malik, J.Y. Wang, K. Drlica, X. Zhao, T. Lu,
gyrase-mediated illegitimate recombination and SOS induction, Genes Cells Escherichia coli genes that reduce the lethal effects of stress, BMC Microbiol.
3 (1998) 511e520. 10 (2010) 35.
[72] M.A. Kohanski, D.J. Dwyer, B. Hayete, C.A. Lawrence, J.J. Collins, A common [95] X. Han, J. Geng, L. Zhang, T. Lu, The role of Escherichia coli YrbB in the lethal
mechanism of cellular death induced by bactericidal antibiotics, Cell 130 action of quinolones, J. Antimicrob. Chemother. 66 (2011) 323e331.
(2007) 797e810. [96] S. Suerbaum, H. Leying, H.P. Kroll, J. Gmeiner, W. Opferkuch, Influence of beta-
[73] X. Wang, X. Zhao, Contribution of oxidative damage to antimicrobial lethality, lactam antibiotics and ciprofloxacin on cell envelope of Escherichia coli,
Antimicrob. Agents Chemother. 53 (2009) 1395e1402. Antimicrob. Agents Chemother. 31 (1987) 1106e1110.
[74] X. Wang, X. Zhao, M. Malik, K. Drlica, Contribution of reactive oxygen species [97] B. Lindner, A. Wiese, K. Brandenburg, U. Seydel, A. Dalhoff, Lack of interaction
to pathways of quinolone-mediated bacterial cell death, J. Antimicrob. Che- of fluoroquinolones with lipopolysaccharides, Antimicrob. Agents Chemother.
mother. 65 (2010) 520e524. 46 (2002) 1568e1570.
[75] H. Engelberg-Kulka, R. Hazan, S. Amitai, mazEF: a chromosomal toxineanti- [98] K.R. Mariner, N. Ooi, D. Roebuck, A.J. O’Neill, I. Chopra, Further character-
toxin module that triggers programmed cell death in bacteria, J. Cell. Sci. 118 ization of Bacillus subtilis antibiotic biosensors and their use for antibac-
(2005) 4327e4332. terial mode-of-action studies, Antimicrob. Agents Chemother. 55 (2011)
[76] I. Kolodkin-Gal, B. Sat, A. Keshet, H. Engelberg-Kulka, The communication 1784e1786.
factor EDF and the toxineantitoxin module mazEF determine the mode of [99] S.C. Kampranis, A. Maxwell, Conformational changes in DNA gyrase revealed
action of antibiotics, PLoS Biol. 6 (2008) e319. by limited proteolysis, J. Biol. Chem. 273 (1998) 22606e22614.

Please cite this article in press as: G. Cheng, et al., Antibacterial action of quinolones: From target to network, European Journal of Medicinal
Chemistry (2013), http://dx.doi.org/10.1016/j.ejmech.2013.01.057

You might also like