Sie sind auf Seite 1von 14

International Journal of Antimicrobial Agents 32 (2008) 207–220

Review

A bioinformatic approach to understanding antibiotic resistance in


intracellular bacteria through whole genome analysis
Silpak Biswas, Didier Raoult, Jean-Marc Rolain ∗
URMITE UMR 6236, CNRS IRD, Faculté de Médecine et de Pharmacie, Université de la Méditerranée, 27 Bd Jean Moulin,
13385 Marseille Cedex 05, France
Received 19 March 2008; accepted 19 March 2008

Abstract
Intracellular bacteria survive within eukaryotic host cells and are difficult to kill with certain antibiotics. As a result, antibiotic resistance
in intracellular bacteria is becoming commonplace in healthcare institutions. Owing to the lack of methods available for transforming these
bacteria, we evaluated the mechanisms of resistance using molecular methods and in silico genome analysis. The objective of this review
was to understand the molecular mechanisms of antibiotic resistance through in silico comparisons of the genomes of obligate and facultative
intracellular bacteria. The available data on in vitro mutants reported for intracellular bacteria were also reviewed. These genomic data were
analysed to find natural mutations in known target genes involved in antibiotic resistance and to look for the presence or absence of differ-
ent resistance determinants. Our analysis revealed the presence of tetracycline resistance protein (Tet) in Bartonella quintana, Francisella
tularensis and Brucella ovis; moreover, most of the Francisella strains possessed the blaA gene, AmpG protein and metallo-␤-lactamase family
protein. The presence or absence of folP (dihydropteroate synthase) and folA (dihydrofolate reductase) genes in the genome could explain
natural resistance to co-trimoxazole. Finally, multiple genes encoding different efflux pumps were studied. This in silico approach was an
effective method for understanding the mechanisms of antibiotic resistance in intracellular bacteria. The whole genome sequence analysis will
help to predict several important phenotypic characteristics, in particular resistance to different antibiotics. In the future, stable mutants should
be obtained through transformation methods in order to demonstrate experimentally the determinants of resistance in intracellular bacteria.
© 2008 Elsevier B.V. and the International Society of Chemotherapy. All rights reserved.

Keywords: Antibiotic resistance; Genomics; Sequence analysis; Intracellular bacteria; In silico; In vitro mutant

1. Introduction Treating bacterial infections is increasingly complicated


by the ability of bacteria to develop resistance to different
Intracellular bacteria are defined by their capacity to antibiotics. Resistance to antibiotics can be caused by a
survive and live inside eukaryotic host cells. As a result, variety of mechanisms: (i) the presence of an enzyme that
they have developed diverse strategies to survive within this inactivates the antimicrobial agent; (ii) a mutation in the target
compartment. These bacteria are responsible for enormous of the antimicrobial agent that reduces its binding capacity;
morbidity and mortality worldwide and are difficult to kill (iii) post-transcriptional and post-translational modification
with certain antibiotics. The barrier to antibiotic treatment of of the target of the antimicrobial agent, which reduces its
obligate intracellular bacteria is the difference between the binding capacity; (iv) reduced uptake of the antimicrobial
localisation of antibiotics within the cellular compartments agent; and (v) active efflux of the antimicrobial agent [2].
of infected cells and the localisation of the bacteria [1]. Bacteria can develop resistance to antibiotics through
two genetic processes: (i) mutation and selection (vertical
gene transfer); and (ii) exchange of genes between strains
∗ Corresponding author. Present address: Unité des Rickettsies, CNRS
and species (horizontal gene transfer). Among intracellular
UMR 6020, IFR 48, Faculté de Médecine et de Pharmacie, Université de la
Méditerranée, 27 Bd Jean Moulin, 13385 Marseille Cedex 05, France.
bacteria, antibiotic resistance is primarily due to spontaneous
Tel.: +33 4 91 38 55 17; fax: +33 4 91 83 03 90. mutations or multiple mutations in the bacterial genome (i.e.
E-mail address: jm.rolain@medecine.univ-mrs.fr (J.-M. Rolain). vertical gene transfer). To date, there is only one example of

0924-8579/$ – see front matter © 2008 Elsevier B.V. and the International Society of Chemotherapy. All rights reserved.
doi:10.1016/j.ijantimicag.2008.03.017
208 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

horizontal gene transfer conferring resistance to tetracycline


among intracellular bacteria, namely Chlamydia suis [3].
In Mycobacterium tuberculosis, the causative agent of
tuberculosis, all of the drug resistance determinants are
chromosomally encoded, arising exclusively through the
acquisition and maintenance of spontaneous chromosomal
mutations in target or complementary genes [4]. Resistance-
associated point mutations, deletions or insertions in M.
tuberculosis have been previously described for all first-line
drugs (e.g. isoniazid, rifampicin, pyrazinamide, ethambutol
and streptomycin) in addition to several second-line and Fig. 1. Lifestyles of intracellular bacteria: (1) bacterial escape into the
cytosol after exit from the endosomal compartment (e.g. Rickettsia, Shigella,
newer drugs (e.g. ethionamide, fluoroquinolones, macrolides Listeria); (2) survival in non-fused phagosomes (e.g. Bartonella, Brucella,
and nitroimidazopyrans) [5]. Legionella); (3) segregation from the endolytic route and formation of a
The objective of this review was to present an overview unique inclusion vacuole (e.g. Chlamydia) and (4) survival by fusion with
of the molecular evidence for bacterial resistance to the lysosome (e.g. Coxiella, Tropheryma, Francisella).
antimicrobial agents among intracellular bacteria using a
bioinformatic analysis of whole genome sequences and in after exit from an endosomal compartment with or without
silico analysis of target genes. fusion of the phagosomal vacuole with lysosomes (e.g.
Rickettsia, Shigella and Listeria); (b) survival in non-fused
phagosomes (e.g. Bartonella, Brucella and Legionella); (c)
survival in fused phagosomes (e.g. Chlamydia); and (d) sur-
2. Intracellular behaviour of bacteria and antibiotic vival in fused phagolysosomes (e.g. Coxiella, Tropheryma
activity and Francisella) (Fig. 1).
Antibiotic activity against intracellular bacteria depends
The intracellular localisation of some bacteria remains on several factors, including pharmacodynamic and phar-
a critical point explaining the failure of some antibiotic macokinetic properties of antibiotics. First, in order to
treatments in infected hosts. Parasites that multiply only be active, antibiotics must reach the infected cells in
within eukaryotic cells are obligate intracellular pathogens, their tissue compartments via the systemic route. Second,
whereas facultative intracellular pathogens can also multiply antibiotics need to reach and concentrate within intracellular
in cell-free models [6]. compartments. The intracellular to extracellular ratio (C/E)
Four categories of mechanisms exist to explain the sur- is a very important factor and can be determined by several
vival of intracellular bacteria: (a) survival in the cytoplasm methods, including radiometric, fluorometric and chemical

Table 1
Antibiotic susceptibility results for various intracellular bacteria
Intracellular bacteria Antibiotic

ERY AMG TET QUI CHL SXT RIF BL


Bartonella spp. S S S S S S S S
Tropheryma whipplei S S S R S S S S
Francisella tularensis S/R S S S S R S R
Rickettsia spp.
Typhus group
R. prowazekii S R S S S R S R
R. typhi S R S S S R S R
Spotted fever group
R. conorii subgroup
R. conorii, R. rickettsii R R S S S R S R
R. massiliae subgroup
R. massiliae, R. montanensis R R S S S R R R
Ehrlichia spp.
E. canis R R S R R R S R
E. chaffeensis R R S R R R S R
Wolbachia spp. R R S S/R R R S R
Coxiella burnetii S/R R S S R S S R
Brucella spp. S/R S S S R R S R
ERY, erythromycin; AMG, aminoglycosides; TET, tetracycline; QUI, quinolones; CHL, chloramphenicol; SXT, sulfamethoxazole/trimethoprim (co-
trimoxazole); RIF, rifampicin; BL, ␤-lactams; S, susceptible; R, resistant.
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 209

techniques. For example, the fluoroquinolones accumulate susceptibility should be determined only in cell models.
within phagocytes with a C/E ratio of 6:7 for granulo- The choice of cell system depends on each pathogen, but
cytes, 3:4 for macrophages and ca. 2:1 for epithelial cells in general cell lines that are easy to obtain and grow are
[7–9]. Third, antibiotics should remain active within the used (e.g. Vero, L929 or MRC5 cells) [6]. The methods for
targeted intracellular compartment, without inactivation by evaluating antibiotic susceptibility vary with the nature of
cellular metabolism and/or deleterious effect of pH [6,10]. the intracellular pathogen [11]. One important technique
Some antibiotics are more effective at neutral or basic pH is enumeration of viable intracellular microorganisms
values (e.g. fluoroquinolone compounds) but others (e.g. (colony-forming unit or plaque assay) after various times of
rifampicin) are more effective at acidic pH values [6]. antibiotic exposure compared with drug-free controls (e.g.
Rickettsia, Coxiella) [6]. Other methods of evaluation have
also been used, including determination of the percentage
3. Overview of natural antibiotic susceptibility of infected cells, flow cytometry, immunofluorescence tech-
among intracellular bacteria niques, luciferase techniques and quantitative polymerase
chain reaction (PCR) [7,12]. There are very few reports of
The antibiotic susceptibility of facultative intracellular antibiotic susceptibility testing among fastidious bacteria, as
bacteria can be assessed in cell-free systems using minimum the methods are time consuming and labour intensive.
inhibitory concentration (MIC) determination methods; Table 1 summarises the natural susceptibility to antibi-
for obligate intracellular bacteria, however, antibiotic otics among intracellular bacteria. Bacteria of the genus

Fig. 2. Phylogenetic tree based on 16S sequences of intracellular bacteria used in this study.
210 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

Bartonella are susceptible to all antibiotics in vitro, including Table 2


␤-lactams, aminoglycosides, chloramphenicol, tetracyclines, Candidate genes involved in antibiotic resistance
macrolides, rifampicin, fluoroquinolones and co-trimoxazole Antibiotic Candidate genes
[13,14]. Tropheryma whipplei displays a homogeneous pat- Macrolides erm, mef, msr, 23S rRNA, rplD (L4
tern of antibiotic susceptibility in axenic medium [15], with r-protein), rplV (L22 r-protein)
almost all antibiotics showing at least some activity, except Tetracyclines 16S rRNA, tet gene, rpsL (S12 r-protein),
fluoroquinolones [16]. Francisella tularensis strains are rpsG (S7r-protein)
susceptible to streptomycin, gentamicin, doxycycline, chlo- Aminoglycosides 16S rRNA, rpsL (S12 r-protein),
ramphenicol and quinolones but show heterogeneous sus- aminoglycoside-modifying enzymes
ceptibility to erythromycin [17,18]. Rickettsia are naturally Quinolones gyrA, gyrB, parC, parE
resistant to ␤-lactams, aminoglycosides and co-trimoxazole. Rifampicin rpoB
Chloramphenicol cat, 23S rRNA
The typhus group is susceptible to erythromycin, whereas the ␤-Lactams mecA
spotted fever group is not [19,20] (Table 1). Trimethoprim folA
In vitro antibiotic susceptibility studies have shown that Sulfamethoxazole folP
various species of Ehrlichia and Wolbachia are susceptible erm, erythromycin ribosome methylation; mef, macrolide efflux; msr,
to doxycycline and rifampicin, although these bacteria methionine sulfoxide reductase; r-protein, ribosomal protein; tet, tetracy-
show heterogeneous susceptibility to quinolone compounds cline resistance protein; gyrA, DNA gyrase subunit A; gyrB, DNA gyrase
[21–23] (Table 1). Antibiotic susceptibility testing of subunit B; parC, DNA topoisomerase IV subunit A; parE, DNA topoi-
somerase IV subunit B; rpoB, DNA-directed RNA polymerase subunit ␤;
Coxiella burnetii showed that amikacin and amoxicillin cat, chloramphenicol acetyltransferase; mecA, penicillin-binding protein 2’;
were not effective, whereas co-trimoxazole, rifampicin, folA, dihydrofolate reductase; folP, dihydropteroate synthase.
doxycycline, clarithromycin and quinolones were all bacte-
riostatic [24,25]. There is heterogeneity in susceptibility to action for different antibiotics, including inhibition of protein
erythromycin among the strains tested [25,26]. Strains of synthesis, interference with cell wall synthesis, interference
Brucella spp. also show heterogeneous susceptibility to ery- with nucleic acid synthesis and inhibition of metabolic
thromycin but are susceptible to almost all antibiotics except pathways. Candidate genes involved in the mechanisms of
chloramphenicol, co-trimoxazole and ␤-lactams [27,28]. antibiotic resistance are described in Table 2 and Fig. 3.

5.1. Inhibition of protein synthesis is the main


4. Methods for whole genome sequence analysis mechanism of action for macrolides, lincosamides,
chloramphenicol, aminoglycosides and tetracyclines
Total numbers of bacterial genomes used in this study
are described in Fig. 2, which represents a phylogenetic 5.1.1. Macrolide–lincosamide–streptogramin (MLS)
tree based on 16S sequence comparison. The target genes antibiotics
involved in antibiotic resistance were retrieved from avail- The MLS antibiotics are an important group of translation
able genomes at the Kyoto Encyclopedia of Genes and inhibitors that act on the 50S ribosomes [29]. The MLS group
Genomes (KEGG) (http://www.genome.jp/kegg/) database. was defined on the basis of cross-resistance patterns, which
The nucleotide sequences of target genes and/or amino acid showed that these drugs acted on the peptidyl transferase
sequences were compared and aligned using the ClustalW centre of the 50S subunit. Binding of these drugs was found
program (http://www.ebi.ac.uk/clustalw/) to examine pos- to involve domains II and V of the 23S rRNA [4,5].
sible mutations known to be associated with antibiotic Intrinsic resistance to MLSB (macrolide–lincosamide–
resistance. The keywords used for the in silico genome streptogramin B) antibiotics in bacteria is generally due
study were: efflux; multidrug; ABC transporters; MFS; to low permeability of the outer membrane to these
RND; MATE; cat gene; tet gene; aminoglycoside-modifying hydrophobic compounds [2]. Three different mechanisms
enzymes; chloramphenicol; folate (folA, folP); 23S; L4; of acquired MLS resistance have been found in bacteria
L22; 16S; erm gene; gyrA and gyrB gene; macrolide; [30–32]. The first described mechanism was a result of
aminoglycoside; tetracycline; beta-lactam; fluoroquinolone; post-transcriptional modifications of the 23S rRNA by
rifampin; trimethoprim; and sulfamethoxazole. adenine-N6 -methyltransferase. Ribosomal target modifi-
cation confers cross-resistance to MLSB antibiotics and
remains the most frequent mechanism of resistance. The
5. Mode of action and mechanisms of antibiotic genes encoding these methylases have been termed erm
resistance: in silico genomic analysis and study of (erythromycin ribosome methylation) [2,33,34]. In our study,
natural and in vitro mutants erm genes were not found in the genome of intracellular
bacteria.
Most antimicrobial agents used for the treatment of Another mechanism of resistance is active drug efflux
bacterial infections can be categorised according to their mediated by the membrane-bound efflux protein encoded
main mechanism of action. There are different modes of by mef(A), which confers resistance only to 14- and
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 211

Fig. 3. (a) Action of macrolides on the peptidyl-tRNA molecule during elongation, resulting inhibition of protein synthesis. (b) Candidate genes for macrolide,
chloramphenicol, tetracycline and aminoglycoside resistance. (c) Secondary structure of Escherichia coli 23S rRNA showing the hairpin 35 in domain II and
part of domain V. The nucleotides in domain II and domain V whose mutations cause resistance to macrolides are shown by green dots. (d) Candidate genes
(gyrA, gyrB) for fluoroquinolones resistance. (e) Structure of RNA polymerase showing the ␤ subunit, the binding site of the antibiotic rifampicin. Rifampicin
inhibits DNA-dependent RNA polymerase by binding to the ␤ subunit encoded by the rpoB gene.

15-membered macrolides [2,35,36]. By in silico genome Another mechanism of resistance is mutation of the
analysis, macrolide-specific efflux proteins (e.g. MacA bacterial 23S rRNA, or mutations, insertions or deletions
and MacB) were identified in the Bartonella genome. in ribosomal protein genes [34,39] (Fig. 3b). Mutations in
In the Escherichia coli genome, ybjYZ were suspected the 23S rRNA at position A2058 and/or A2059 remain the
to be genes for ABC drug efflux transporters and were most common and confer the highest levels of macrolide
renamed macAB (macrolide-specific ABC-type efflux resistance. A lower level of drug resistance is provided by
carrier) [37,38]. Plasmids carrying both the macA and macB mutations at positions 2057, 2452 and 2611 of the 23S
genes conferred resistance against macrolides composed of rRNA [34] (Fig. 3c). Mutations in the single copies of genes
14- and 15-membered compounds, but conferred only weak encoding the L4 and L22 ribosomal proteins have also been
resistance against 16-membered compounds. implicated in macrolide resistance [39,40]. Amino acids
212 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

59–90 and 85–87 in the L4 and L22 ribosomal proteins, of nine repeated amino acids between amino acids R71 and
respectively, have been reported to be important mutational A72 in the highly conserved region of the protein [50].
regions for macrolide resistance [39,41–43]. Recently, we reported various changes in the 23S rRNA
Heterogeneity in susceptibility to erythromycin has gene and the L4 ribosomal protein for the B. henselae strain
been shown among F. tularensis subsp. holarctica, with Marseille as well as other B. henselae isolates [51]. Most of
biovar I being erythromycin sensitive and biovar II being the mutations in the 23S rRNA gene (e.g. A2058G, A2058C
erythromycin resistant [19]. The molecular mechanisms of and C2611T) were previously reported to confer ery-
resistance to erythromycin are not known, but alignment of thromycin resistance in other bacteria as well [34,39,44,52].
domain V of the 23S rRNA using in silico methods showed We found amino acid mutations at two different posi-
an A → C transition at position 2059 (E. coli numbering) tions (G71R and H75Y) in ribosomal protein L4 among
in one F. tularensis genome (F. tularensis subsp. holarctica erythromycin-resistant strains of B. henselae [51] (Table 4).
LVS), which could explain the heterogeneity in susceptibility The A2058G mutation in the erythromycin-resistant strain of
to erythromycin for F. tularensis strains (Fig. 4). Recently, Bartonella bacilliformis was also reported by our team [53].
we reported one natural mutation (A2059G) in the 23S rRNA
gene (Table 3) in a Bartonella henselae isolate from a lymph 5.1.2. Chloramphenicol
node from a patient with cat-scratch disease, suggesting Chloramphenicol is a bacteriostatic antimicrobial agent
that naturally occurring erythromycin-resistant strains may that is effective against a wide variety of microorganisms.
infect humans. Chloramphenicol interferes with microbial protein synthesis
A study by Branger et al. [44] confirmed that macrolides by binding to the 50S ribosomal subunit and inhibiting the
and telithromycin lacked antimicrobial activity against peptidyltransferase step in protein synthesis [2].
Ehrlichia [21–23,45]. They reported numerous specific There are three known mechanisms of resistance to
mutations in nucleotides known to confer resistance to chloramphenicol: reduced membrane permeability; muta-
macrolides in the Ehrlichia chaffeensis 23S rRNA gene tion of the 23S ribosomal subunit; and elaboration of
(e.g. T754G, G2057A, A2059G and C2611T) [44]. Wol- chloramphenicol acetyltransferase. Mutations in 23S rRNA
bachia pipientis, a closely related organism, also possesses have been previously reported in chloramphenicol-resistant
mutations T754A, A2058G, A2059C and C2611G, which strains of E. coli and Ehrlichia [44].
could explain the intrinsic resistance of this bacterium to High-level resistance to chloramphenicol is conferred
macrolides [46] (Table 3). by the cat gene, which encodes an enzyme called chloram-
Recently, we found three amino acid differences in phenicol acetyltransferase that inactivates chloramphenicol
the highly conserved region at the C terminus of the L22 [54,55]. This enzyme is usually encoded on a plasmid and
ribosomal protein between typhus group (TG) rickettsiae can be transferred along with genes conferring resistance to
and spotted fever group (SFG) rickettsiae [47], which may a number of other antibiotics [55]. Whole genome analysis
explain the heterogeneity in susceptibility to erythromycin data showed the presence of a cat gene in the genome of
between these two subgroups. Rickettsia typhi and Rickettsia Bartonella.
prowazekii showed two amino acid changes at positions
83 and 84 (Streptococcus pneumoniae numbering) and a 5.1.3. Aminoglycosides
single amino acid change at position 89 compared with the Aminoglycosides kill bacteria by inhibiting protein
seven SFG rickettsial strains [47]. The three amino acid synthesis via binding to the 16S rRNA and disrupting
differences found between the two subgroups of rickettsiae the integrity of the bacterial cell membrane [56,57]. The
were located in a highly conserved region of the L22 most frequently encountered mechanism of resistance
protein. In E. coli, deletion of three amino acids in this to aminoglycosides is their structural modification by
conserved region (Met82 -Lys83 -Arg84 ) conferred resistance specific enzymes produced by resistant organisms. The three
to erythromycin [48]. Similarly, amino acid substitutions classes of such aminoglycoside-modifying enzymes are:
as well as insertions or deletions within the region between (1) aminoglycoside nucleotidyltransferases, which transfer
amino acid positions 80 and 94 have been reported in nucleotide triphosphates; (2) aminoglycoside acetyltrans-
in vitro mutants of Haemophilus influenzae resistant to ferases, which transfer the acetyl group from acetyl-CoA;
macrolide compounds [49]. Finally, we found a macrolide and (3) aminoglycoside phosphotransferases, which transfer
2 -phosphotransferase-like protein in the genome of T. the phosphoryl group from ATP [58–61].
whipplei Twist. It has previously been reported that the genome of
Rickettsia conorii contains one gene encoding a protein
5.1.1.1. In vitro mutants. In Bartonella spp., differ- similar to an aminoglycoside 3 -phosphotransferase, and
ent mechanisms of erythromycin resistance have been there is a streptomycin resistance protein homologue in the
reported using in vitro studies. We demonstrated that the genome of Rickettsia felis; on the other hand, the genomes of
fully erythromycin-resistant strain of Bartonella quintana R. typhi and R. prowazekii do not contain these genes [62]. A
obtained after 16 passages in vitro harboured a 27-base repeat predicted aminoglycoside phosphotransferase is also present
insertion in ribosomal protein L4, resulting in an insertion in the genome of Wolbachia Bma. In silico data showed the
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 213

Fig. 4. Alignment of 23S rRNA (domain V) sequences of different intracellular bacteria showing changes at position 2059 for Francisella tularensis subsp.
holarctica LVS (A2059C) and for Ehrlichia, Neorickettsia and Wolbachia spp. (A2059G).

presence of a probable aminoglycoside efflux pump (AcrD, mechanism of pinocytosis by the eukaryotic cell, explaining
acriflavine resistance protein D) in the genome of Ehrlichia the slow intracellular accumulation of these drugs [45,70].
ruminantium str. Welgevonden (France) and E. ruminantium
str. Gardel. Aminoglycoside N(6 )-acetyltransferase (aacA4) 5.1.4. Tetracyclines
was found in the genome of C. burnetii. Tetracyclines are broad-spectrum antimicrobial agents
Other mechanisms of resistance include alteration of the with activity against a broad range of pathogenic bacteria,
30S ribosomal subunit target by mutation (mutation in 16S including intracellular bacteria [71,72]. Tetracycline is
rRNA gene) (Fig. 3b), methylation of the aminoglycoside- thought to inhibit the growth of bacteria by entering the
binding site, and reduction of the intracellular concentration bacterial cell, binding to ribosomes and inhibiting protein
of aminoglycosides by changes in outer membrane per- synthesis [73]. Several studies have found a single, high-
meability, decreased inner membrane transport and active affinity binding site for tetracyclines in the ribosomal 30S
efflux [63–69]. subunit [74,75] (Fig. 3b).
Previous studies indicated that aminoglycosides could In most species, resistance to tetracycline is conferred
not diffuse passively through the eukaryotic cell membrane by genes with two main modes of action. The first group of
because of their large size and negative charge. Cellular genes encodes efflux systems that transport the drug from
uptake of this class of antibiotics corresponded to an active the inside to the outside of the bacterial cell; the second
214 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

Table 3
Genome analysis data and natural mutations in candidate genes for antibiotic resistance in intracellular bacteria
Intracellular bacteria Antibiotic class and mechanism of resistance

MAC FQa AMG BL RIF (rpoB)a CHL


23S rRNA L22 ribosomal protein gyrA parC
Bartonella henselae A2059G Ser-83 → Ala
Bartonella quintana Ser-83 → Ala
Bartonella bacilliformis Ser-83 → Ala MBL
Rickettsia spp. Triple AA differences APH BLA Phe-973 → Leu
Francisella tularensis A2059C BLA
Ehrlichia chaffeensis T754G Ser-83 → Ala MBL G2057A
G2057A
A2059G
C2611T
Wolbachia pipientis T754A MBL
A2058G
A2059C
C2611G
Tropheryma whipplei Ser-83 → Ala Ser-96 → Ala
Coxiella burnetii AAC MBL
Brucella suis MBL
MAC, macrolides; FQ, fluoroquinolones; AMG, aminoglycosides; BL, ␤-lactams; RIF, rifampicin; CHL, chloramphenicol; AA, amino acid; APH, aminogly-
coside phosphotransferase; AAC, aminoglycoside acetyltransferase; MBL, metallo-␤-lactamase; BLA, ␤-lactamase.
a Amino acid changes resulting from the gene mutation are given.

group encodes ribosomal protection proteins, which remove generally interact with the ribosome and, as a result, protein
tetracycline from the ribosome [72,76]. All of the tet efflux synthesis is unaffected by the presence of the antibiotics [2].
genes encode membrane-associated proteins that export
tetracycline from the cell. These tetracycline resistance 5.2. The cell wall can be affected by drugs that prevent
determinants are often associated with transmissible genetic the production of new cell walls, leading to cell lysis and
elements including plasmids, transposons and integrons [72]. death; β-lactam drugs such as penicillins,
Genome data analysis revealed the presence of a tetracy- cephalosporins and carbapenems all interfere with cell
cline resistance protein (Tet) in the genome of B. quintana, wall production
most F. tularensis strains and Brucella ovis. Among several
tetracycline resistance determinants (TetA, TetB and TetC), 5.2.1. β-Lactam antibiotics
the genome of Brucella melitensis biovar Abortus possessed ␤-Lactam antibiotics are among the most commonly
the tetracycline resistance protein TetB. These proteins used antimicrobial agents. They interfere with the final stage

Table 4
Molecular mechanisms of resistance in intracellular bacteria selected in vitro for different classes of antibiotics
Intracellular bacteria Antibiotic class and mechanism of resistance

MAC FQ (gyrA)a RIF (rpoB)a SMX (dhps)a

23S rRNA L4 ribosomal protein


Bartonella henselae A2058G G71R
A2058C H75Y
C2611T
Bartonella quintana Insertion GRARHSSAR
Bartonella bacilliformis A2058G Asp-87 → Asn Ser-531 → Phe
Rickettsia spp. Leu-151 → Phe
Phe-201 → Leu
Val-271 → Ile
Arg-546 → Lys
Francisella tularensis Thr-83 → Ile
Coxiella burnetii Glu-87 → Gly
Tropheryma whipplei Val-57 → Ile
Thr-102 → Pro
Leu-162 → Ile
MAC, macrolides; FQ, fluoroquinolones; RIF, rifampicin; SMX, sulfamethoxazole.
a Amino acid changes resulting from the gene mutation are given.
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 215

of cell wall synthesis by inhibiting the bacterial enzymes enzyme that can affect supercoiling of DNA, and inhibition
transpeptidases and carboxypeptidases that catalyse the reac- of this activity by fluoroquinolones is associated with rapid
tions of peptidoglycan synthesis [77,78]. These enzymes, killing of the bacterial cell [2,91–93].
commonly called penicillin-binding proteins (PBPs), Alterations in target enzymes appear to be the most dom-
cross-link the peptidoglycan polymers. Peptidoglycan is an inant factors in expression of resistance to quinolones [92].
essential component of the bacterial cell wall, and inhibition A small region from codon 67 to 106 of gyrA in E. coli was
of PBPs causes bacteriolysis by creating a wall unable to designated the ‘quinolone resistance-determining region’
withstand osmotic forces [77,79]. (QRDR) [2] and variations in this region were found in
The greatest single cause of resistance to ␤-lactam antibi- species with natural resistance to fluoroquinolones [94,95].
otics is antibiotic-inactivating enzymes, the ␤-lactamases, Analysis of the T. whipplei genome allowed the identifi-
which efficiently catalyse irreversible hydrolysis of the cation of the gyrA and parC gene encoding the ␣ subunit of
amide bond of the ␤-lactam ring resulting in biologically the natural fluoroquinolone targets DNA gyrase and topoiso-
inactive products [80]. Over 250 ␤-lactamases have been merase IV, respectively [16]. Heterogeneity in susceptibility
described, varying in their substrate profiles, inhibition to fluoroquinolones in T. whipplei [16] was found to be
profiles, molecular mass, isoelectric point, amino acid associated with mutations in the DNA gyrase gene. In the T.
sequence and molecular structure [77,81,82]. Genes encod- whipplei GyrA and ParC sequences, alanine residues were
ing ␤-lactamases can be localised either on plasmids or found at positions 81 and 96, respectively, corresponding
on the bacterial chromosome and are found both among to a serine at position 83 in E. coli GyrA [96] and a serine
Gram-negative and Gram-positive organisms [2]. at position 80 in E. coli ParC [97], respectively (Table 3).
Whole genome analysis showed that most Francisella GyrA-mediated natural resistance to fluoroquinolones has
strains possessed a blaA (␤-lactamase class A) gene and also been described in Mycobacterium spp., which are
AmpG protein. ␤-Lactamases of Amber’s class A are the closely phylogenetically related to T. whipplei. In silico
most commonly found in bacteria resistant to ␤-lactam genome analysis revealed a natural mutation at position 83
antibiotics [83]. The AmpG protein is an integral membrane of the QRDR region (Ser-83 → Ala) of the GyrA protein
protein that functions as a peptidoglycan-specific permease for three Bartonella species (Table 3). Many examples exist
and can be used to transport new drugs mimicking the demonstrating that species naturally bearing a serine residue
murein recycled compounds into the cytoplasm [84]. at position 83 of GyrA protein are usually susceptible to
Metallo-␤-lactamase family proteins were found in the fluoroquinolones, whereas the presence of an alanine at this
genomes of most of the intracellular bacteria used in this critical position corresponds to natural resistance to these
study (Table 3). antibiotics [98–101]. Similarly, Maurin et al. [102] observed
Five specific open-reading frames (ORFs) related to that a serine residue at position 83 of the GyrA protein in
antibiotic resistance have been previously identified in susceptible species of Ehrlichia is replaced by an alanine
the genome of R. felis, including a class C ␤-lactamase, a residue in fluoroquinolone-resistant species.
penicillin acylase homologue and an ABC-type multidrug
transporter system [85]. Interestingly, a previous study 5.3.1.1. In vitro mutants. Resistance to fluoroquinolones
showed the presence of two genes encoding ␤-lactamases in has been described in some strains of C. burnetii and it was
the genome of R. conorii and none in the genome of R. typhi shown that the mechanism involved two distinct nucleotide
and R. prowazekii, which possessed PBPs and ampG genes mutations in the GyrA protein (Glu-87 → Gly and Glu-
instead [63]. 87 → Lys) [103,104] (Table 4). An amino acid change
(Asp-87 → Asn) in its GyrA has also been reported recently
5.3. Nucleic acid synthesis can be interrupted by several in a ciprofloxacin-resistant strain of B. bacilliformis (Table 4).
mechanisms
5.3.2. Rifampicin
5.3.1. Fluoroquinolones The molecular mechanism of rifampicin activity involves
Quinolones or fluoroquinolones are among the most inhibition of DNA-dependent RNA polymerase. This enzyme
important antibacterial drugs and are used extensively for is a complex oligomer composed of four different subunits (␣,
the treatment of bacterial infections both in human and ␤, ␤ and ␴ encoded by rpoA, rpoB, rpoC and rpoD, respec-
veterinary medicine [86]. tively) (Fig. 3e). Rifampicin binds to the ␤ subunit of RNA
Fluoroquinolones exert their antibacterial effects by polymerase and results in transcription inhibition [105,106].
inhibition of certain bacterial topoisomerase enzymes, Resistance to rifampicin is primarily caused by mutations
namely DNA gyrase and topoisomerase IV. DNA gyrase and in the rpoB gene. In the majority of rifampicin-resistant
topoisomerase IV are heterotetrameric proteins composed isolates, mutations occurred within an 81-bp hotspot region
of two subunits, designated A and B [87–90]. The genes (the rifampicin resistance-determining region (RRDR),
encoding the A and B subunits are referred to as gyrA and codons 507–533 according to E. coli numbering) in the
gyrB (DNA gyrase) or parC and parE (DNA topoisomerase rpoB gene [4,107]. Previously, Rolain et al. [19] found that
IV), respectively (Table 2; Fig. 3d). DNA gyrase is the only susceptibility to rifampicin varied, with R. prowazekii, R.
216 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

typhi, Rickettsia canada, Rickettsia bellii and most SFG The target for sulphonamide action is dihydropteroate
rickettsiae being susceptible to rifampicin, whilst the Rick- synthase (DHPS), which catalyses the condensation of
ettsia massiliae subgroup (R. massiliae, Rickettsia montana, para-aminobenzoic acid with 7,8-dihydro-6-hydroxy-
Rickettsia rhipicephalus and Rickettsia aeschlimannii) methylopterine pyrophosphate to form 7,8-dihydropteroate
were more resistant to rifampicin. Drancourt and Raoult [115–118]. Sulphonamide resistance is commonly mediated
[108] investigated the genetic basis for natural rifampicin by the presence of alternative drug-resistant forms of DHPS.
resistance in representatives of the TG and the two SFG Chromosomal mutations in the dhps gene that confer
subgroups of rickettsiae by sequence analysis of the rpoB resistance to sulphonamides have also been identified in a
gene. They found a single point mutation resulting in a number of bacteria [116,119].
Phe → Leu change at position 973 of the R. conorii rpoB Mutation in folA and folP (structural genes for DHFR and
sequence (Table 3). This single point mutation, which DHPS, respectively) could confer resistance to trimethoprim
appeared to be specific for the naturally rifampicin-resistant and sulfamethoxazole, respectively. Interestingly, Rickettsia
subgroup, was not previously implicated in rifampicin spp. are resistant to co-trimoxazole and it has been found
resistance in other bacteria. Resistance to rifampicin can that the folP and folA genes are absent in most Rickettsia
also be due to expression of an efflux system [109,110]. spp. Coxiella burnetii, which is naturally susceptible to
co-trimoxazole compounds, showed both folP and folA genes
in its genome. Interestingly, T. whipplei is susceptible to
5.3.2.1. In vitro mutants. Amino acid substitutions in the
co-trimoxazole, although only the folP gene is present in the
RNA polymerase and rpoB point mutations have been
Tropheryma genome. A recent study [120] demonstrated that
demonstrated following in vitro selection of rifampicin-
the MICs against the two strains of T. whipplei ranged from
resistant R. prowazekii [111] and R. typhi [112]. Troyer et al.
0.5 mg/L to 1 mg/L for sulfadiazine compared with 0.5 mg/L
[112] reported the detection of a rifampicin-resistant strain of
for sulfamethoxazole, leading the authors to suggest that
R. typhi (Ethiopian). The basis of this resistance was inves-
sulfadiazine was as effective as sulfamethoxazole in vitro.
tigated by sequencing and mapping point mutations in the
rpoB gene of the mutant and then comparing the sequences
5.4.1.1. In vitro study. We developed a new method to study
of wild-type and rifampicin-resistant R. typhi rpoB genes. A
antibiotic susceptibility in fastidious bacteria such as T. whip-
total of eight nucleotide substitutions occurred, three of which
plei using E. coli gene complementation. In the genome of T.
resulted in amino acid substitutions in the mutant strain:
whipplei, a typical DHPS-encoding gene, the target gene for
leucine for phenylalanine at residue 151, phenylalanine for
sulfamethoxazole, is not found as an individual ORF. DNA
leucine at residue 201 and valine for isoleucine at residue
sequencing of two samples (before and after failure) from
271 [112] (Table 4). In another study by Rachek et al., com-
a patient with clinically acquired resistance to trimetho-
parison of the rpoB sequences from the rifampicin-sensitive
prim/sulfamethoxazole, using specific oligonucleotide
R. prowazekii Madrid E strain and a rifampicin-resistant
primers for the candidate gene folP, showed three amino
mutant identified a single point mutation that resulted in an
acid changes [121]. Gene complementation in E. coli showed
arginine-to-lysine change at position 546 of the rpoB gene
that the mutated sequence was associated with resistance.
[111].
A recently reported rifampicin-resistant strain of B.
bacilliformis from our group showed a mutation at serine
6. Efflux pumps
531 (Ser → Phe) in the RRDR of the rpoB gene [53]
(Table 4). This 531 site is one of the most frequent sites
Efflux pumps are present in all living cells. They partici-
of mutation, also conferring rifampicin resistance in other
pate in the detoxification process, expelling various harmful
bacterial species [106].

5.4. Folate synthesis, which is necessary for DNA


replication, is blocked by sulphonamides and
trimethoprim

5.4.1. Trimethoprim and sulphonamides


Trimethoprim is an analogue of dihydrofolic acid, an
essential component in the synthesis of amino acids and
nucleotides, which competitively inhibits the enzyme
dihydrofolate reductase (DHFR). Resistance can be caused
by a number of mechanisms, including overproduction of
host DHFR, mutations in the structural gene for DHFR
and acquisition of a foreign gene (dfr) encoding a resistant Fig. 5. Correlation between genome size of intracellular bacteria and number
DHFR enzyme [2,113,114]. of ATP-binding cassette (ABC) transporters present in the genome.
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 217

compounds and xenobiotics [122,123]. Bacterial drug efflux rial gene products that may be involved in antibiotic transport
pumps are currently classified into five families [124–130]: and efflux from bacterial cells. Among intracellular bacteria,
(i) the ATP-binding cassette (ABC) superfamily; (ii) MFS the spread of antibiotic resistance is mainly due to vertical
(major facilitator superfamily) transporters; (iii) the RND gene transfer, which is also observed in M. tuberculosis. This
(resistance–nodulation–cell division) superfamily; (iv) the group of organisms could be an interesting paradigm to iden-
SMR (small multidrug resistance) family; and (v) the MATE tify different resistance determinants using whole genome
(multidrug and toxic compound extrusion) family. analysis. Microarray studies could also be used to determine
antibiotic resistance genes in intracellular bacteria. The
6.1. In silico genome analysis study of natural mutants and in vitro resistant mutants will
additionally help in the understanding of the efficacy of
Whole genome analysis using an in silico method showed different classes of antibiotics among intracellular bacteria.
that the number of genes encoding ABC transporters in We developed a new method to study antibiotic resistance
intracellular bacteria varied from four (R. prowazekii, E. in fastidious bacteria (e.g. T. whipplei) using E. coli gene
ruminantium Welgevonden France and E. ruminantium complementation. Transformation of intracellular bacteria
Gardel) to 277 (Brucella suis). We found a correlation is one of the difficulties we face in working with these
between the numbers of encoded ABC transporters and microorganisms. This complementation approach could be
genome size, which is depicted in Fig. 5. Genome size used with other intracellular bacteria for further study and
and number of ABC transporters in different intracellular could open a door to the identification and demonstration
bacteria are given in Supplementary Data 1. of resistance determinants among these bacteria. Genomic
Several of the efflux systems found among intracellular studies will further clarify how resistance to novel classes
bacteria are given in Supplementary Data 2. Efflux systems of antibiotics arises, in addition to the fitness costs to the
including pH adaptation potassium efflux system proteins organism that result from resistance.
(PhaAB, PhaC, PhaD, PhaE, PhaF and PhaG) have been Funding: No funding sources.
found in multiple different Bartonella spp. The genome of T. Competing interests: None declared.
whipplei Twist contains the multidrug efflux protein QacA Ethical approval: Not required.
and a cation efflux protein. The plasmid-encoded multidrug
resistance gene, qacA, from Staphylococcus aureus mediates
resistance to a number of classes of antimicrobial organic Appendix A. Supplementary data
cations, including intercalating dyes, quaternary ammonium
compounds, diamidines and biguanidines [70,76]. Interest- Supplementary data associated with this arti-
ingly, T. whipplei has been shown to be resistant to several cle can be found, in the online version, at
antiseptics, including glutaraldehyde and peracetic acid doi:10.1016/j.ijantimicag.2008.03.017.
[131].
In the genome of F. tularensis we found outer membrane
efflux proteins along with other efflux systems. The OEP References
(outer membrane efflux protein) family forms trimeric
channels allowing export of a variety of substrates in [1] McOrist S. Obligate intracellular bacteria and antibiotic resistance.
Gram-negative bacteria [132]. Trends Microbiol 2000;8:483–6.
[2] Fluit AC, Visser MR, Schmitz FJ. Molecular detection of antimicro-
The glutathione-regulated potassium efflux system bial resistance. Clin Microbiol Rev 2001;14:836–71.
protein, KefB, is present in the genome of all Rickettsia spp. [3] Dugan J, Rockey DD, Jones L, Andersen AA. Tetracycline resis-
and is also known as K+ /H+ antiporter. It is localised to the tance in Chlamydia suis mediated by genomic islands inserted
inner membrane of the cell and facilitates potassium efflux into the chlamydial inv-like gene. Antimicrob Agents Chemother
2004;48:3989–95.
[133].
[4] Ramaswamy S, Musser JM. Molecular genetic basis of antimicrobial
agent resistance in Mycobacterium tuberculosis: 1998 update. Tuber
Lung Dis 1998;79:3–29.
7. Concluding remarks and perspectives [5] Somoskovi A, Parsons LM, Salfinger M. The molecular basis of resis-
tance to isoniazid, rifampin, and pyrazinamide in Mycobacterium
The emergence and spread of antimicrobial resistance tuberculosis. Respir Res 2001;2:164–8.
[6] Rolain JM, Raoult D. Treatment of intracellular infections. In: Hooper
determinants continues to challenge our ability to treat DC, Rubinstein E, editors. Quinolone antimicrobial agents. 3rd ed.
serious infections. The past two decades have produced Washington, DC: ASM Press; 2002.
substantial research into the mechanisms by which bacteria [7] García I, Pascual A, Ballesta S, Perea EJ. Uptake and intracellular
develop and disperse resistance determinants, although there activity of ofloxacin isomers in human phagocytic and non-phagocytic
is still much to be learned. In silico genome analysis will cells. Int J Antimicrob Agents 2000;15:201–5.
[8] Pascual A, García I, Ballesta S, Perea EJ. Uptake and intracellular
help to predict possible molecular mechanisms of resistance activity of moxifloxacin in human neutrophils and tissue-
among intracellular bacteria. Studying microbial genomes cultured epithelial cells. Antimicrob Agents Chemother 1999;43:
will also aid in the discovery process by identifying all bacte- 12–5.
218 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

[9] Perea EJ, García I, Pascual A. Comparative penetration of lome- [29] Tsui WH, Yim G, Wang HH, McClure JE, Surette MG, Davies J. Dual
floxacin and other quinolones into human phagocytes. Am J Med effects of MLS antibiotics: transcriptional modulations and interac-
1992;92(4A):48S–51S. tions on the ribosome. Chem Biol 2004;11:1307–16.
[10] Maurin M, Benoliel AM, Bongrand P, Raoult D. Phagolysosomal [30] Leclercq R, Giannattasio RB, Jin HJ, Weisblum B. Bacterial resistance
alkalinization and the bactericidal effect of antibiotics: the Coxiella to macrolide, lincosamide and streptogramin antibiotics by target
burnetii paradigm. J Infect Dis 1992;166:1097–102. modification. Antimicrob Agents Chemother 1991;35:1267–72.
[11] Maurin M, Raoult D. Optimum treatment of intracellular infection. [31] Weisblum B. Erythromycin resistance by ribosome modification.
Drugs 1996;52:45–59. Antimicrob Agents Chemother 1995;39:577–85.
[12] Gant VA, Warnes G, Phillips I, Savidge GF. The application of flow [32] Weisblum B. Macrolide resistance. Drug Resist Updat 1998;1:29–41.
cytometry to the study of bacterial responses to antibiotics. J Med [33] Robert MC, Sutcliffe J, Courvalin P, Jensen LB, Rood J, Sep-
Microbiol 1993;39:147–54. pala H. Nomenclature for macrolide and macrolide–lincosamide–
[13] Ives TJ, Manzewitsch P, Regnery RL, Butts JD, Kebede M. In streptogramin B resistance determinants. Antimicrob Agents
vitro susceptibilities of Bartonella henselae, B. quintana, B. eliza- Chemother 1999;43:2823–30.
bethae, Rickettsia rickettsii, R. conorii, R. akari, and R. prowazekii to [34] Vester B, Douthwaite S. Macrolide resistance conferred by base sub-
macrolide antibiotics as determined by immunofluorescent-antibody stitutions in 23S rRNA. Antimicrob Agents Chemother 2001;45:1–12.
analysis of infected Vero cell monolayers. Antimicrob Agents [35] Ardanuy C, Tubau F, Liñares J, Domínguez MA, Pallarés R, Martín R;
Chemother 1997;41:578–82. Spanish Pneumococcal Infection Study Network (G03/103). Distri-
[14] Maurin M, Gasquet S, Ducco C, Raoult D. MICs of 28 antibi- bution of subclasses mefA and mefE of the mefA gene among clinical
otic compounds for 14 Bartonella (formerly Rochalimaea) isolates. isolates of macrolide-resistant (M-phenotype) Streptococcus pneu-
Antimicrob Agents Chemother 1995;39:2387–91. moniae, viridans group streptococci, and Streptococcus pyogenes.
[15] Boulos A, Rolain JM, Mallet MN, Raoult D. Molecular evaluation of Antimicrob Agents Chemother 2005;49:827–9.
antibiotic susceptibility of Tropheryma whipplei in axenic medium. J [36] Luna VA, Coates P, Eady EA, Cove JH, Nguyen TT, Roberts MC. A
Antimicrob Chemother 2005;55:178–81. variety of gram-positive bacteria carry mobile mef genes. J Antimicrob
[16] Masselot F, Boulos A, Maurin M, Rolain JM, Raoult D. Molecu- Chemother 1999;44:19–25.
lar evaluation of antibiotic susceptibility: the Tropheryma whipplei [37] Paulsen IT, Sliwinski MK, Saier Jr MH. Microbial genome analyses:
paradigm. Antimicrob Agents Chemother 2003;47:1658–64. global comparisons of transport capabilities based on phylogenies,
[17] Ikäheimo I, Syrjälä H, Karhukorpi J, Schildt R, Koskela M. In bioenergetics and substrate specificities. J Mol Biol 1998;277:573–92.
vitro antibiotic susceptibility of Francisella tularensis isolated from [38] Kobayashi N, Nishino K, Yamaguchi A. Novel macrolide-specific
humans and animals. J Antimicrob Chemother 2000;46:287–90. ABC-type efflux transporter in Escherichia coli. J Bacteriol
[18] Kudelina RI, Olsufiev NG. Sensitivity to macrolide antibiotics and 2001;183:5639–44.
lincomycin in Francisella tularensis holarctica. J Hyg Epidemiol [39] Tait-Kamradt A, Davies T, Cronan M, Jacobs MR, Appelbaum PC,
Microbiol Immunol 1980;24:84–91. Sutcliffe J. Mutations in 23S rRNA and ribosomal protein L4 account
[19] Rolain JM, Maurin M, Vestris G, Raoult D. In vitro susceptibilities of for resistance in pneumococcal strains selected in vitro by macrolide
27 Rickettsiae to 13 antimicrobials. Antimicrob Agents Chemother passage. Antimicrob Agents Chemother 2000;44:2118–25.
1998;42:1537–41. [40] Gregory ST, Dahlberg AE. Erythromycin resistance mutations in ribo-
[20] Raoult D, Bres P, Drancourt M, Vestris G. In vitro susceptibili- somal proteins L22 and L4 perturb the higher order structure of 23 S
ties of Coxiella burnetii, Rickettsia rickettsii, and Rickettsia conorii ribosomal RNA. J Mol Biol 1999;289:827–34.
to the fluoroquinolone sparfloxacin. Antimicrob Agents Chemother [41] Davydova N, Streltzov V, Wilce M, Liljas A, Garber M. L22 ribo-
1991;35:88–91. somal protein and effect of its mutation on ribosome resistance to
[21] Rolain JM, Maurin M, Bryskier A, Raoult D. In vitro activities of erythromycin. J Mol Biol 2002;322:635–44.
telithromycin (HMR 3647) against Rickettsia rickettsii, Rickettsia [42] Doktor SZ, Shortridge VD, Beyer JM, Flamm RK. Epidemiology
conorii, Rickettsia africae, Rickettsia typhi, Rickettsia prowasekii, of macrolide and/or lincosamide resistant Streptococcus pneumoniae
Coxiella burnetii, Bartonella henselae, Bartonella quintana, Bar- clinical isolates with ribosomal mutations. Diagn Microbiol Infect
tonella bacilliformis, and Ehrlichia chaffeensis. Antimicrob Agents Dis 2004;49:47–52.
Chemother 2000;44:1391–3. [43] Tait-Kamradt A, Davies T, Appelbaum PC, Depardieu F, Courvalin
[22] Brouqui P, Raoult D. In vitro susceptibility of Ehrlichia sennetsu to P, Petitpas J, et al. Two new mechanisms of macrolide resistance in
antibiotics. Antimicrob Agents Chemother 1990;34:1593–6. clinical strains of Streptococcus pneumoniae from Eastern Europe and
[23] Brouqui P, Raoult D. In vitro antibiotic susceptibility of the newly North America. Antimicrob Agents Chemother 2000;44:3395–401.
recognized agent of ehrlichiosis in humans, Ehrlichia chaffeensis. [44] Branger S, Rolain JM, Raoult D. Evaluation of antibiotic suscep-
Antimicrob Agents Chemother 1992;36:2799–803. tibilities of Ehrlichia canis, Ehrlichia chaffeensis, and Anaplasma
[24] Jabarit-Aldighieri N, Torres H, Raoult D. Susceptibility of Rick- phagocytophilum by real-time PCR. Antimicrob Agents Chemother
ettsia conorii, R. rickettsii, and Coxiella burnetii to PD 127,391, PD 2004;48:4822–8.
131,628, pefloxacin, ofloxacin, and ciprofloxacin. Antimicrob Agents [45] Maurin M, Raoult D. Use of aminoglycosides in treatment of infec-
Chemother 1992;36:2529–32. tions due to intracellular bacteria. Antimicrob Agents Chemother
[25] Maurin M, Raoult D. In vitro susceptibilities of spotted fever 2001;45:2977–86.
group rickettsiae and Coxiella burnetii to clarithromycin. Antimicrob [46] Fenollar F, Maurin M, Raoult D. Wolbachia pipientis growth kinetics
Agents Chemother 1993;37:2633–7. and susceptibilities to 13 antibiotics determined by immunofluores-
[26] Raoult D, Torres H, Drancourt M. Shell-vial assay: evaluation of cence staining and real-time PCR. Antimicrob Agents Chemother
a new technique for determining antibiotic susceptibility, tested 2003;47:1665–71.
in 13 isolates of Coxiella burnetii. Antimicrob Agents Chemother [47] Rolain JM, Raoult D. Prediction of resistance to erythromycin in the
1991;35:2070–7. genus Rickettsia by mutations in L22 ribosomal protein. J Antimicrob
[27] Rolain JM, Maurin M, Raoult D. Bactericidal effect of antibiotics Chemother 2005;56:396–8.
on Bartonella and Brucella spp.: clinical implications. J Antimicrob [48] Canu A, Malbruny B, Coquemont M, Davies TA, Appelbaum PC,
Chemother 2000;46:811–4. Leclerq R. Diversity of ribosomal mutations conferring resistance to
[28] Baykam N, Esener H, Ergönül O, Eren S, Celikbas AK, Dokuzoguz B. macrolides, clindamycin, streptogramin, and telithromycin in Strep-
In vitro antimicrobial susceptibility of Brucella species. Int J Antimi- tococcus pneumoniae. Antimicrob Agents Chemother 2002;46:125–
crob Agents 2004;23:405–7. 31.
S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220 219

[49] Kosowska K, Credito K, Pankuch GA, Hoellman D, Lin G, Clark [72] Chopra I, Roberts M. Tetracycline antibiotics: mode of action, appli-
C, et al. Activities of two novel macrolides, GW 773546 and cations, molecular biology, and epidemiology of bacterial resistance.
GW 708408, compared with those of telithromycin, erythromycin, Microbiol Mol Biol Rev 2001;65:232–60.
azithromycin, and clarithromycin against Haemophilus influenzae. [73] Speer BS, Shoemaker MB, Salyers AA. Bacterial resistance to tetracy-
Antimicrob Agents Chemother 2004;48:4113–9. cline: mechanisms, transfer, and clinical significance. Clin Microbiol
[50] Meghari S, Rolain JM, Grau GE, Platt E, Barrassi L, Mege JL, Raoult Rev 1992;5:387–99.
D. Anti-angiogenic effect of erythromycin in Bartonella quintana: in [74] Oehler R, Polacek N, Steiner G, Barta A. Interaction of tetracycline
vitro model of infection. J Infect Dis 2006;193:380–6. with RNA: photoincorporation into ribosomal RNA of Escherichia
[51] Biswas S, Raoult D, Rolain JM. Molecular characterization of resis- coli. Nucleic Acids Res 1997;25:1219–24.
tance to macrolides in Bartonella henselae. Antimicrob Agents [75] Schnappinger D, Hillen W. Tetracyclines: antibiotic action, uptake,
Chemother 2006;50:3192–3. and resistance mechanisms. Arch Microbiol 1996;165:359–69.
[52] Ng LK, Martin I, Liu G, Bryden L. Mutation in 23S rRNA associ- [76] Connell SR, Tracz DM, Nierhaus KH, Taylor DE. Ribosomal pro-
ated with macrolide resistance in Neisseria gonorrhoeae. Antimicrob tection proteins and their mechanism of tetracycline resistance.
Agents Chemother 2002;46:3020–5. Antimicrob Agents Chemother 2003;47:3675–81.
[53] Biswas S, Raoult D, Rolain JM. Molecular mechanisms of resistance [77] Essack SY. The development of beta-lactam antibiotics in response
to antibiotics in Bartonella bacilliformis. J Antimicrob Chemother to the evolution of beta-lactamases. Pharm Res 2001;18:1391–9.
2007;59:1065–70. [78] Danziger LH, Pendland SL. Bacterial resistance to beta-lactam antibi-
[54] Shaw WV. Chloramphenicol acetyltransferase: enzymology and otics. Am J Health Syst Pharm 1995;52(6 Suppl. 2):S3–8.
molecular biology. CRC Crit Rev Biochem 1983;14:1–46. [79] Jacoby GA. Extended-spectrum beta-lactamases and other enzymes
[55] Calia KE, Calia FM. Chloramphenicol. In: Yu VL, Merigan Jr TC, providing resistance to oxyimino-beta-lactams. Infect Dis Clin North
Barriere SL, editors. Antimicrobial therapy and vaccines. Baltimore, Am 1997;11:875–87.
MD: Williams & Wilkins; 1998. p. 765–74. [80] Matagne A, Lamotte-Brasseur J, Frère JM. Catalytic properties
[56] Mingeot-Leclercq MP. Aminoglycosides: activity and resistance. of class A beta-lactamases: efficiency and diversity. Biochem J
Antimicrob Agents Chemother 1999;43:727–37. 1998;330:581–98 [Erratum. Biochem J 1998;331:975.].
[57] Davies J, Wright G. Bacterial resistance to aminoglycoside antibi- [81] Frère JM, Dubus A, Galleni M, Matagne A, Amicosante G. Mechanis-
otics. Trends Microbiol 1997;5:234–40. tic diversity of beta-lactamases. Biochem Soc Trans 1999;27:58–63.
[58] Azucena E, Mobashery S. Aminoglycoside-modifying enzymes: [82] Jacoby GA. Beta-lactamase nomenclature. Antimicrob Agents
mechanisms of catalytic processes and inhibition. Drug Resist Updat Chemother 2006;50:1123–9.
2001;4:106–17. [83] Yang Y, Rasmussen BA, Shlaes DM. Class A beta-lactamases—
[59] Rather PN. Origins of the aminoglycoside modifying enzymes. Drug enzyme–inhibitor interactions and resistance. Pharmacol Ther
Resist Updat 1998;1:285–91. 1999;83:141–51.
[60] Wright GD. Aminoglycoside-modifying enzymes. Curr Opin Micro- [84] Chahboune A, Decaffmeyer M, Brasseur R, Joris B. Membrane topol-
biol 1999;2:499–503. ogy of the Escherichia coli AmpG permease required for recycling of
[61] Smith CA. Aminoglycoside antibiotic resistance by enzymatic deac- cell wall anhydromuropeptides and AmpC beta-lactamase induction.
tivation. Curr Drug Targets Infect Disord 2002;2:143–60. Antimicrob Agents Chemother 2005;49:1145–9.
[62] Rolain JM, Raoult D. Genome comparison analysis of molecular [85] Ogata H, Renesto P, Audic S, Robert C, Blanc G, Fournier PE, et al.
mechanisms of resistance to antibiotics in the Rickettsia genus. Ann The genome sequence of Rickettsia felis identifies the first putative
N Y Acad Sci 2005;1063:222–30. conjugative plasmid in an obligate intracellular parasite. PLoS Biol
[63] Magnet S, Courvalin P, Lambert T. Resistance–nodulation–cell 2005;3:e248.
division-type efflux pump involved in aminoglycoside resistance [86] Ball P. Quinolone generations: natural history or natural selection? J
in Acinetobacter baumannii strain BM4454. Antimicrob Agents Antimicrob Chemother 2000;46:17–24.
Chemother 2001;45:3375–80. [87] Drlica K, Zhao XL. DNA gyrase, topoisomerase IV, and the 4-
[64] Magnet S, Smith TA, Zheng R, Nordmann P, Blanchard JS. Amino- quinolones. Microbiol Rev 1997;61:377–92.
glycoside resistance resulting from tight drug binding to an altered [88] Everett MJ, Piddock LJV. Mechanisms of resistance to fluoroquino-
aminoglycoside acetyltransferase. Antimicrob Agents Chemother lones. In: Kuhlmann J, Dahlhoff A, Zeiler HJ, editors. Quinolone anti-
2003;47:1577–83. bacterials. Berlin, Germany: Springer-Verlag KG; 1998. p. 259–97.
[65] Keggs PA, Thompson J, Cundliffe E. Methylation of 16S ribosomal [89] Hooper DC. Bacterial topoisomerases, anti-topoisomerases and anti-
RNA and resistance to aminoglycoside antibiotics in clones of Strep- topoisomerase resistance. Clin Infect Dis 1998;27:54–63.
tomyces lividans carrying DNA from Streptomyces tenjimariensis. [90] Drlica K, Malik M. Fluoroquinolones: action and resistance. Curr Top
Mol Gen Genet 1985;200:415–21. Med Chem 2003;3:249–82.
[66] Recht MI, Douthwaite S, Dahlquist KD, Puglisi JD. Effect [91] Drlica K. Mechanism of fluoroquinolone action. Curr Opin Microbiol
of mutations in the A site of 16 S rRNA on aminoglyco- 1999;2:504–8.
side antibiotic–ribosome interaction. J Mol Biol 1999;286:33– [92] Hooper DC. Mechanisms of fluoroquinolone resistance. Drug Resist
43. Updat 1999;2:38–55.
[67] Shakil S, Khan R, Zarrilli R, Khan AU. Aminoglycosides versus [93] Blondeau JM. Fluoroquinolones: mechanism of action, classifica-
bacteria—a description of the action, resistance mechanism, and noso- tion, and development of resistance. Surv Ophthalmol 2004;49(Suppl.
comial battleground. J Biomed Sci 2008;15:5–14. 2):S73–8.
[68] Magnet S, Blanchard JS. Molecular insights into aminoglycoside [94] Chen JY, Siu LK, Chen YH, Lu PL, Ho M, Peng CF. Molecular
action and resistance. Chem Rev 2005;105:477–97. epidemiology and mutations at gyrA and parC genes of ciprofloxacin-
[69] Taber HW, Mueller JP, Miller PF, Arrow AS. Bacterial uptake of resistant Escherichia coli isolates from a Taiwan medical center.
aminoglycoside antibiotics. Microbiol Rev 1987;51:439–57. Microb Drug Resist 2001;7:47–53.
[70] Tulkens PM. Intracellular pharmacokinetics and localization of antibi- [95] Waters B, Davies J. Amino acid variation in the GyrA subunit of bac-
otics as predictors of their efficacy against intraphagocytic infections. teria potentially associated with natural resistance to fluoroquinolone
Scand J Infect Dis Suppl 1990;74:209–17. antibiotics. Antimicrob Agents Chemother 1997;41:2766–9.
[71] Roberts MC. Tetracycline resistance determinants: mechanisms of [96] Herrera G, Aleixandra V, Urios A, Blanco M. Quinolone action
action, regulation of expression, genetic mobility, and distribution. in Escherichia coli cells carrying gyrA and gyrB mutations. FEMS
FEMS Microbiol Rev 1996;19:1–24. Microbiol Lett 1993;106:187–91.
220 S. Biswas et al. / International Journal of Antimicrobial Agents 32 (2008) 207–220

[97] Vila J, Ruiz J, Goni P, De Anta MT. Detection of mutations in parC in Streptococcus pneumoniae. Antimicrob Agents Chemother 2001;45:
quinolone-resistant clinical isolates of Escherichia coli. Antimicrob 1104–8.
Agents Chemother 1996;40:491–3. [115] Padayachee T, Klugman KP. Novel expansions of the gene encod-
[98] Yoshida H, Kojima T, Yamagishi JI, Nakamura S. Quinolone-resistant ing dihydropteroate synthase in trimethoprim–sulfamethoxazole-
mutations of the gyrA gene of Escherichia coli. Mol Gen Genet resistant Streptococcus pneumoniae. Antimicrob Agents Chemother
1988;211:1–7. 1999;43:2225–30.
[99] Oram M, Fisher L. 4-Quinolone resistance mutations in the DNA [116] Shiota T, Baugh M, Jackson R, Dillard R. The enzymatic synthesis
gyrase of Escherichia coli clinical isolates identified by using of hydroxymethyldihydropteridine pyrophosphate and dihydrofolate.
the polymerase chain reaction. Antimicrob Agents Chemother Biochemistry 1969;8:5022–8.
1991;35:387–9. [117] Sköld O. Sulfonamide resistance: mechanisms and trends. Drug Resist
[100] Houssaye S, Gutmann L, Varon E. Topoisomerase mutations Updat 2000;3:155–60.
associated with in vitro selection of resistance to moxifloxacin [118] Skold O. Resistance to trimethoprim and sulfonamides. Vet Res
in Streptococcus pneumoniae. Antimicrob Agents Chemother 2001;32:261–73.
2002;46:2712–5. [119] Swedberg G, Ringertz S, Skold O. Sulfonamide resistance in Strep-
[101] Cullen M, Wyke A, Kuroda R, Fisher L. Cloning and characteri- tococcus pyogenes is associated with differences in the amino acid
zation of a DNA gyrase A gene from Escherichia coli that confers sequence of its chromosomal dihydropteroate synthase. Antimicrob
clinical resistance to 4-quinolones. Antimicrob Agents Chemother Agents Chemother 1998;42:1062–7.
1989;33:886–94. [120] Bakkali N, Fenollar F, Rolain JM, Raoult D. Comment on: therapy
[102] Maurin M, Abergel C, Raoult D. DNA gyrase-mediated natural for Whipple’s disease. J Antimicrob Chemother 2008;61:968–9.
resistance to fluoroquinolones in Ehrlichia spp. Antimicrob Agents [121] Bakkali N, Fenollar F, Biswas S, Rolain JM, Raoult D. Acquired
Chemother 2001;45:2098–105. resistance to trimethoprim–sulfamethoxazole during Whipple’s dis-
[103] Musso D, Drancourt M, Osscini S, Raoult D. Sequence of quinolone ease and expression of the causative target gene. J Infect Dis 2008 (in
resistance-determining region of gyrA gene for clinical isolates and press).
for an in vitro-selected quinolone-resistant strain of Coxiella burnetii. [122] Pagès JM, Masi M, Barbe J. Inhibitors of efflux pumps in Gram-
Antimicrob Agents Chemother 1996;40:870–3. negative bacteria. Trends Mol Med 2005;11:382–9.
[104] Spyridaki I, Psaroulaki A, Aransay A, Scoulica E, Tselentis Y. Diag- [123] Poole K. Efflux-mediated multiresistance in Gram-negative bacteria.
nosis of quinolone-resistant Coxiella burnetii strains by PCR-RFLP. Clin Microbiol Infect 2004;10:12–26.
J Clin Lab Anal 2000;14:59–63. [124] Kumar A, Schweizer HP. Bacterial resistance to antibiotics: active
[105] Lisitsyn NA, Sverdlov ED, Moiseyeva EP, Danilevskaya ON, Niki- efflux and reduced uptake. Adv Drug Deliv Rev 2005;57:1486–513.
forov VG. Mutation to rifampicin resistance at the beginning of the [125] Fath MJ, Kolter R. ABC transporters: bacterial exporters. Microbiol
RNA polymerase beta subunit gene in Escherichia coli. Mol Gen Rev 1993;57:995–1017.
Genet 1984;196:173–4. [126] Borges-Walmsley MI, Walmsley AR. The structure and function of
[106] Telenti A, Imboden P, Marchesi F, Lowrie D, Cole S, Colston MJ, drug pumps. Trends Microbiol 2001;9:71–9.
et al. Detection of rifampicin-resistance mutations in Mycobacterium [127] Yoshida H, Bogaki M, Nakamura S, Ubukata K, Konno M.
tuberculosis. Lancet 1993;341:647–50. Nucleotide sequence and characterization of the Staphylococcus
[107] Yam WC, Tam CM, Leung CC, Tong HL, Chan KH, Leung ET, et aureus norA gene, which confers resistance to quinolones. J Bacteriol
al. Direct detection of rifampin-resistant Mycobacterium tuberculosis 1990;172:6942–9.
in respiratory specimens by PCR-DNA sequencing. J Clin Microbiol [128] Hansen LH, Johannesen E, Burmolle M, Sorensen AH, Sorensen
2004;42:4438–43. SJ. Plasmid-encoded multidrug efflux pump conferring resistance
[108] Drancourt M, Raoult D. Characterization of mutations in the rpoB to olaquindox in Escherichia coli. Antimicrob Agents Chemother
gene in naturally rifampin-resistant Rickettsia species. Antimicrob 2004;48:3332–7.
Agents Chemother 1999;43:2400–3. [129] Schuldiner S, Lebendiker M, Yerushalmi H. EmrE, the small-
[109] Chandrasekaran S, Lalithakumari D. Plasmid-mediated rifampicin est ion-coupled transporter, provides a unique paradigm for
resistance in Pseudomonas fluorescens. J Med Microbiol structure–function studies. J Exp Biol 1997;200:335–41.
1998;47:197–200. [130] Omote H, Hiasa M, Matsumoto T, Otsuka M, Moriyama Y. The MATE
[110] Bambeke VF, Balzi E, Tulkens PM. Antibiotic efflux pumps. Biochem proteins as fundamental transporters of metabolic and xenobiotic
Pharmacol 2000;60:457–70. organic cations. Trends Pharmacol Sci 2006;27:587–93.
[111] Rachek LI, Tucker AM, Winkler HH, Wood DO. Transforma- [131] La Scola B, Rolain JM, Maurin M, Raoult D. Can Whipple’s dis-
tion of Rickettsia prowazekii to rifampin resistance. J Bacteriol ease be transmitted by gastroscopes? Infect Control Hosp Epidemiol
1998;180:2118–24. 2003;24:191–4.
[112] Troyer JM, Radulovic S, Andersson SG, Azad AF. Detection of point [132] Zhao Q, Li X, Srikumar R, Poole K. Contribution of outer mem-
mutations in rpoB gene of rifampin-resistant Rickettsia typhi. Antimi- brane efflux protein OprM to antibiotic resistance in Pseudomonas
crob Agents Chemother 1998;42:1845–6. aeruginosa independent of MexAB. Antimicrob Agents Chemother
[113] Huovinen P. Resistance to trimethoprim–sulfamethoxazole. Clin 1998;42:1682–8.
Infect Dis 2001;32:1608–14. [133] Bakker EP, Booth IR, Dinnbier U, Epstein W, Gajewska A. Evi-
[114] Maskell JP, Sefton AM, Hall LM. Multiple mutations modulate dence for multiple K+ export systems in Escherichia coli. J Bacteriol
the function of dihydrofolate reductase in trimethoprim-resistant 1987;169:3743–9.

Das könnte Ihnen auch gefallen