Open access

Pseudomonas aeruginosa: Multi-Drug-Resistance Development and Treatment Options

Written By

Georgios Meletis and Maria Bagkeri

Submitted: 25 August 2012 Published: 29 May 2013

DOI: 10.5772/55616

From the Edited Volume

Infection Control

Edited by Silpi Basak

Chapter metrics overview

5,754 Chapter Downloads

View Full Metrics

1. Introduction

Antibiotic resistance is a worldwide problem of major importance. Isolations in some countries of multi-drug-resistant (resistant to three or more classes of antimicrobials), extensively-drug-resistant (resistant to all but one or two classes) or even pan-drug-resistant (resistant to all available classes) Gram-negative pathogens are causing therapeutic problems and- in the same time- are posing infection control issues in many hospitals. In fact, numerous studies highlight the link between multi-drug-resistance and increased morbidity and mortality, increased length of hospital stay and higher hospital costs [1-4].

Pseudomonas aeruginosa is a Gram-negative opportunistic nosocomial pathogen responsible for a wide range of infections that may present high rates of antimicrobial resistance. The genome of this microorganism is among the largest in the bacterial world allowing for great genetic capacity and high adaptability to environmental changes. In fact, P. aeruginosa has 5567 genes encoded in 6.26 Mbp of DNA while Escherichia coli K12 for example has 4279 genes encoded in 4.46 Mbp and Haemophilus influenzae Rd has 1.83 Mbp encoding 1714 genes [5]. This large genetic armamentarium- that can be further enriched with the addition of genes acquired by transferable genetic elements via horizontal gene transfer- is a major contributing factor to its formidable ability to develop resistance against all known antibiotics.

Generally, antibiotic resistance mechanisms of P. aeruginosa can be divided in intrinsic and acquired. Intrinsic refers to resistance that is a consequence of a large selection of genetically-encoded mechanisms and acquired refers to resistance that is achieved via the acquisition of additional mechanisms or is a consequence of mutational events under selective pressure.

Advertisement

2. Intrinsic resistance of Pseudomonas aeruginosa

P. aeruginosa shows inherent resistance to antimicrobial agents through a variety of mechanisms: (1) decreased permeability of the outer membrane, (2) efflux systems which actively pump antibiotics out of the cell, and (3) production of antibiotic-inactivating enzymes [6].

2.1. Outer membrane permeability

The outer membrane of Gram-negative bacteria is a barrier which prevents large hydrophilic molecules to pass through it. Aminoglycosides and colistin interact with lipopolysaccharides changing the permeability of the membrane in order to pass whereas beta-lactams and quinolones need to diffuse through certain porin channels.

Bacteria produce two major classes of porins: general; which allow almost any hydrophilic molecule to pass [7] and specific; which have binding sites for certain molecules, allowing them to be oriented and pass in the most energy-efficient way [8].

Most bacteria posses lots of general porins and relatively few specific ones. However, the exact opposite occurs for P. aeruginosa that expresses mainly specific porins [7].

2.2. Efflux systems

P. aeruginosa expresses several efflux pumps that expel drugs together with other substances out of the bacterial cell. These pumps consist of three proteins: (1) a protein transporter of the cytoplasmatic membrane that uses energy in the form of proton motive force, (2) a periplasmic connective protein, and (3) an outer membrane porin [5].

Efflux system Efflux pump family Substrates References
MexAB-OprM Resistance Nodulation Division (RND) Fluoroquinolones Aminoglycosides
β-Lactams (preferably Meropenem, Ticarcillin)
Tetracycline
Tigecycline
Chloramphenicol
[17]
MexCD-OprJ Resistance Nodulation Division (RND) Fluoroquinolones
β-Lactams (preferably Meropenem, Ticarcillin) Tetracycline
Tigecycline
Chloramphenicol
Erythromycin
Roxythromycin
[17]
MexEF-OprN Resistance Nodulation Division (RND) Fluoroquinolones
β-Lactams (preferably Meropenem, Ticarcillin) Tetracycline
Tigecycline
Chloramphenicol
[17]
[18]
MexXY-OprM Resistance Nodulation Division (RND) Fluoroquinolones Aminoglycosides
β-Lactams (preferably Meropenem, Ticarcillin, Cefepime)
Tetracycline
Tigecycline
Chloramphenicol
[17]
AmrAB-OprA Resistance Nodulation Division (RND) Aminoglycosides [19]
PmpM Multidrug And Toxic compound Extrusion (MATE) Fluoroquinolones [17]
Mef(A) Major Facilitator Superfamily (MFS) Macrolides [20]
ErmEPAF Small Multidrug Resistance (SMR) Aminoglycosides [21]

Table 1.

Efflux systems of P. aeruginosa.

Most antibiotics- except polymyxins- are pumped out [9,10] by these efflux systems (Table 1) therefore their first two components are named multidrug efflux (Mex) along with a letter (e.g. MexA and MexB). The outer membrane porin is called Opr along with a letter (e.g. OprM) [11].

2.3. Antibiotic-inactivating enzymes

P. aeruginosa belongs to the SPICE group of bacteria (Serratia spp., P. aeruginosa, Indole positive Proteus, Citrobacter spp., Enterobacter spp.). These microorganisms share a common characteristic: the ability to produce chromosomal-encoded and inducible AmpC beta-lactamases. These are cephalosporinases that hydrolyze most beta-lactams and are not inhibited by the beta lactamase inhibitors.

Another endogenous beta-lactamase produced by P. aeruginosa is the class D oxacillinase PoxB [12,13]. This enzyme however has only been found in laboratory mutants and is not clinically significant.

Advertisement

3. Antipseudomonal treatment

Despite the intrinsic resistance of P. aeruginosa to many antimicrobials, some antibiotics are active against this microorganism [14]. Those used more frequently belong to three antibiotic classes: (1) Beta-lactams, (2) Quinolones and (3) Aminoglycosides (Table 2).

3.1. Beta-lactams

Beta-lactams bind to and inactivate penicillin-binding proteins (PBPs) that are transpeptidases involved in bacterial cell wall synthesis [15]. The group of beta-lactam antibiotics includes penicillins, cepholosporins, monobactams and carbapenems. The beta-lactams that are most active against P. aeruginosa are: Piperacillin and ticarcillin (penicillins), ceftazidime (3rd generation cephalosporin), cefepime (4th generation cephalosporin), aztreonam (monobactam), imipenem, meropenem and doripenem (carbapenems).

3.2. Quinolones

Quinolones are synthetic antimicrobials that block DNA replication by inhibiting the activity of DNA gyrase and topoisomerase IV [16]. The fluorquinolones with anti-pseudomonal activity are ciprofloxacin, levofloxacin and ofloxacin.

Antibiotic Class Mechanism of action Drug
Penicillins Bacterial cell wall synthesis inhibition Ticarcillin
Penicillin / Beta-lactamase inhibitor Bacterial cell wall synthesis inhibition Ticarcillin/Clavulanic acid
Piperacillin/Tazobactam
Cefalosporins Bacterial cell wall synthesis inhibition Ceftazidime
Cefepime
Monobactams Bacterial cell wall synthesis inhibition Aztreonam
Carbapenems Bacterial cell wall synthesis inhibition Imipenem
Meropenem
Doripenem
Fluoroquinolones Block of DNA synthesis Ciprofloxacin
Levofloxacin
Ofloxacin
Aminoglycosides Protein synthesis inhibition Gentamycin
Tobramycin
Amikacin

Table 2.

Commonly used anti-pseudomonal drugs.

3.3. Aminoglycosides

Aminoglycosides inhibit protein synthesis by binding to the 30S or 50S ribosomal subunit [22]. Drugs of this antibiotic class that can be used against P. aeruginosa are tobramycin, amikacin and gentamicin. Aminoglycosides are associated with ototoxicity and nefrotoxicity [23]. Because of these adverse effects and because of their narrow therapeutic range, aminoglycosides are used in combination with agents belonging to other antibiotic classes. The only treatment in which aminoglycosides are recommended as monotherapy is that of urinary tract infections due to P. aeruginosa [14].

Advertisement

4. Acquired resistance of Pseudomonas aeruginosa

Apart from being resistant to a variety of antimicrobial agents, P. aeruginosa develops resistance to anti-pseudomonal drugs as well. This acquired resistance is a consequence of mutational changes or the acquisition of resistance mechanisms via horizontal gene transfer and can occur during chemotherapy [24]. Mutational events may lead to over-expression of endogenous beta-lactamases or efflux pumps, diminished expression of specific porins and target site modifications while acquisition of resistance genes mainly refers to transferable beta-lactamases and aminoglycoside-modifying enzymes (Table 3).

Resistance to Resistance mechanism
Beta-lactams Endogenous beta-lactamases
Acquired beta-lactamases
Efflux
Diminished permeability
Fluoroquinolones Target site mutations
Efflux
Aminoglycosides Aminoglycoside-modifying enzymes
Efflux
16S rRNA methylases
Polymyxins LPS modification

Table 3.

Resistance mechanisms of P. aeruginosa to anti-pseudomonal drugs.

4.1. Resistance to beta-lactams

Resistance to beta-lactam antibiotics is multi-factorial but is mediated mainly by inactivating enzymes called beta-lactamases. These enzymes cleave the amide bond of the beta-lactam ring causing antibiotic inactivation and are classified according to a structural [25] and a functional [26] classification.

Among the beta-lactams, carbapenems are the most efficient against P. aeruginosa. These agents are stable to the hydrolytic effect of the majority of the beta-lactamases including the Extended Spectrum Beta-Lactamases (ESBLs) [27]. For this reason, the enzymes that possess carbapenemase activity, namely the carbapenemases [28], will be discussed separately in this section.

4.1.1. Expression of endogenous beta-lactamases

Resistance to beta-lactams in clinical isolates is commonly due to the presence of AmpC beta-lactamases [29-36]. Furthermore, the production of AmpC beta-lactamases in P. aeruginosa can be induced by a number of beta-lactam antibiotics such as benzyl penicillines, narrow spectrum cephalosporins and imipenem [37]. In fact, this mutational derepression is one of the most common mechanisms of resistance to beta-lactams in P. aeruginosa [29,32,33,36].

AmpC enzymes are not carbapenemases, they posses however a low potential of carbapenem hydrolysis and their overproduction combined with efflux pumps over-expression and/or diminished outer membrane permeability has been proven to lead also to carbapenem resistance in P. aeruginosa [38].

4.1.2. Acquired beta-lactamases

Acquired beta-lactamases are typically encoded by genes which are located in transferable genetic elements such as plasmids or transposons [39] often on integrons [40-49]. Integrons are genetic elements that capture and mobilize genes [50]. Other genetic elements associated with transferable resistance in P. aeruginosa are the mobile insertion sequences called ISCR elements [49,51-53].

Different types of transferable beta-lactamases have been found in clinical P. aeruginosa isolates around the world (Table 4).

Ambler molecular class Bush-Jacoby-Madeiros group Enzymes References
A 2b TEM-1, -2, -90,
-110, SHV-1
[57,58]
2be PER-1, -2
VEB-1, -2, -3
TEM-4, -21, -24,
-42, -116
SHV-2a, -5, -12
GES/IBC-1, -2, -5,
-8, -9
BEL
LBT 802
CTX-M-1, -2, -43
[10]
[53]
[59-62]
2c PSE-1 (CARB-2), PSE-4 (CARB-1), CARB-3, CARB-4, CARB-like, AER-1 [10]
[63]
2f KPC-2, -5 [64,65]
B 3 IMP-1, -4, -6, -7, -9, -10, -12, -13, -15,
-16, -18, -22
VIM-1, -2, -3, -4, -5, -7, -8, -11, -13, -15, -16,-17, -18
SPM-1
GIM-1
AIM-1
NDM-1
[10]
[47]
[66-76]
C 1 AmpC [77]
D 2d OXA
LCR-1
NPS-1
[10]
[12]
[54]
[57]
[78-80]

Table 4.

Beta-lactamases found in P. aeruginosa isolates.

Among them, carbapenemases are of major clinical importance because they inactivate carbapenems together with other beta-lactams. Ambler class A ESBLs hydrolyze penicillins, narrow- and broad-spectrum cephalosporins and aztreonam [54]. Some TEM and SHV enzymes do not possess broad-spectrum cephalosporinase activity and are called restricted-spectrum beta-lactamases. Class D OXA beta-lactamases are a heterogenous group of enzymes and not all share the same properties. Generally, most of them show a preference for cloxacillin over benzylpenicillin. They confer resistance to amino- and carboxypenicillins and narrow –spectrum cephalosporins even though some of them are ESBLs and a few members of the class present carbapenemase activity [24].

4.1.3. Carbapenemases

P. aeruginosa is the species in which all types of transferable carbapenemases, except SIM-1 [55], have been detected. The class B carbapenemases that bear Zn2+ in their active center [56] are the most frequent around the world in P. aeruginosa isolates and are called metallo-beta-lactamases (MBLs). They hydrolyse in vitro all beta-lactams except aztreonam and are the major cause of high-level carbapenem resistance. Genes that encode MBLs are commonly found as gene cassettes in integrons and are transferable [42]. Interestingly, more resistance genes for other antibiotic classes can be present in the same integrons contributing thus in the development of a multi-drug resistant phenotype.

IMP and VIM type MBLs were first identified in Japan [81] and Italy [82] respectively and have spread though all continents since then. Other metallo-enzymes are more geographically restricted. SPM-1, after causing outbreaks in Brazil [28], has been found in Basel [83] in a single isolate recovered from a patient previously hospitalized in Brazil. GIM-1 and AIM-1 were reported from Germany [41] and Australia [84] and did not spread elsewhere. Finally, the only report for NDM-1 in P. aeruginosa was made from Serbia [76].

Ambler class A carbapenemase KPC was first reported in P. aeruginosa isolates in Colombia [64] but KPC-producing P. aeruginosa isolates have not been reported from other continents except Latin America. KPCs present high rates of carbapenem hydrolysis and inactivate all other beta-lactams including aztreonam.

Enzymes GES/IBC belong to the same enzymatic class but their carbapenemase activity is not as high as that of the KPCs. It may become important however if combined with diminished outer membrane permeability or efflux over-expression. For P. aeruginosa, GES-2 has been reported in South Africa [85] and IBC-2 in Greece [86].

Class D carbapenemases like OXA-198 have been found in P. aeruinosa isolates although such findings are rather rare for this species [87]. The most clinically important carbapenemases are summarized in Table 5.

Ambler molecular class Bush-Jacoby-Madeiros group Carbapenemases
A 2f KPC
B 3 IMP enzymes
VIM enzymes
SPM-1
GIM-1
AIM-1
NDM-1

Table 5.

Clinically important carbapenemases found in P. aeruginosa isolates.

4.1.4. Efflux systems over-expression

Among the various efflux systems of P. aeruginosa, MexAB-OprM, MexXY-OprM and MexCD-OprJ play an important role in developing beta-lactam resistance [88]. Between these three, MexAB-OprM accommodates the broadest range of beta-lactams [24], is by far the better exporter of meropenem [24] and is most frequently related to beta-lactam resistance in clinical P. aeruginosa isolates [33,89]. The efflux pumps may be over-expressed in some isolates [90] contributing thus, together with other mechanisms in the development of multi-drug resistance [24].

4.1.5. Diminished permeability

OprD is a specific porin of the outer membrane of P. aeruginosa through which carbapenems (mainly imipenem) enter into the periplasmic space [91]. Diminished expression [92] or mutational loss [93] of this porin is the most common mechanism of resistance to carbapenems [24,94] and is frequently associated with efflux pumps and/or AmpC over-expression [36,38]. Diminished expression or loss of the OprD porin is a frequent phenomenon during imipenem treatment [95].

4.2. Resistance to fluoroquinolones

High-level resistance to fluoroquinolones is mediated by target site modifications. Efflux plays a contributing role as well [96,97] and the two mechanisms often coexist [32,98-100].

4.2.1. DNA gyrase and topoisomerase IV mutations

Gyrase and topoisomerase are comprised by two subunits each. DNA gyrase (GyrA and GyrB) is the main target of fluoroquinolones in P. aeruginosa. Consequently, mutations are most common for this enzyme rather than for topoisomerase IV (ParC and ParE) [98-102]. Highly resistant isolates have multiple mutations in gyrA and/or parC [98,101-103] while mutations regarding the other subunits are less frequently encountered [100-102,104].

4.2.2. Efflux pumps contribution

Four efflux pumps contribute to fluoroquinolone resistance: MexAB-OprM, MexCD-OprJ, MexEF-OprN and MexXY-OprM [105] as a consequence of mutational events in their repressor genes [24]. Among these, MexAB-OprM, MexCD-OprJ, and MexEF-OprN have been associated to fluoroquinolone resistance in clinical isolates [31,105-107] whereas MexXY-OprM has only been linked rarely to such type of resistance [106].

4.3. Resistance to aminoglycosides

Acquired resistance to aminoglycosides is mediated by transferable aminoglycoside-modifying enzymes (AMEs), rRNA methylases and derepression of endogenous efflux systems [24,108,109].

4.3.1. Aminoglycoside-modifying enzymes

Modification and subsequent inactivation of aminoglycosides is achieved by three deferent mechanisms: (1) acetylation, by aminoglycoside acetyltransferases (AACs), (2) adenylation, by aminoglycoside nucleotidyltransferases (ANTs), and (3) phosphorylation, by aminoglycoside posphoryltransferases (APHs) [108].

Genes encoding AMEs are typically found on integrons together with other genes responsible for transferable resistance for other antibiotic classes. This way AMEs become important determinants for the development of multi-drug resistance in P. aeruginosa and other species [24,108,109].

Enzymatic families that acetylate the 3 and 6’ position of the antibiotic are the most common. Five subfamilies of AAC(3) and two of AAC(6’) have been described for P. aeruginosa, each one presenting different preferences for aminoglycoside substrates (Table 6).

Among the nucleotidyltransferases, ANT(2’)-I is the most frequently encountered in P. aerugiosa. This enzyme is present in isolates showing resistance to gentamicin and tobramycin but not to amikacin [109].

Almost all phosphoryltransferases of P. aeruginosa act in the 3’ position of the aminoglycoside molecule [24]. However, they have less clinical importance because of the fact that they inactivate aminoglycosides that are not routinely used for the treatment of P. aeruginosa infections such as kanamycin and neomycin [109]. The enzymes of this family that inactivate anti-pseudomonal aminoglycosides are APH(3’)-VI [110-112], APH(3’)-IIb-like [113] and APH(2’’) [110]. Despite being reported in some cases, these enzymes remain rare for clinical P. aeruginosa isolates [24].

4.3.2. Efflux systems

Resistance to aminoglycosides in P. aeruginosa can occur independently of aminoglycoside-modifying enzymes in cystic fibrosis patients. This type of resistance has been reported in several studies [99,118-120] and is attributable to over-expression of the MexXY-OprM efflux pump.

Category Enzymatic family Subfamily Substrates References
Acetyltransferases
(AAC)
AAC(3) I Gentamicin [11]
[48]
[108,109]
II Gentamicin
Tobramycin
III Gentamicin
Tobramycin
IV Gentamicin
VI Gentamicin
Tobramycin
AAC(6΄) I Tobramycin
Amikacin
[108,109]
II Tobramycin
Gentamicin
Nucleotidyltransferases
(ANT)
ANT(2΄) Ι Gentamicin
Tobramycin
[109]
ΑΝΤ(4΄) IIa Tobramycin
Amikacin
[114,115]
IIb Tobramycin
Amikacin
ΑΝΤ(3΄) Streptomycin [108]
Phosphoryltransferases
(APH)
APH(3΄) ΙΙ Kanamycin
Neomycin
[109]
[116]
IIb Kanamycin [117]
IIb-like Amikacin
(weakly)
[113]
VI Amikacin
Isepamicin
[110-112]
APH(2΄΄) Gentamicin
Tobramycin
[110]

Table 6.

Aminoglycoside-modifying enzymes found in P. aeruginosa isolates.

4.3.3. 16S rRNA methylases

Methylation of the 16S rRNA of the A site of the 30S ribosomal subunit interferes with aminoglycoside binding and consequently promotes high-level resistance to all aminoglycosides [24]. Different 16S rRNA methylases have been described for P. aeruginosa: RmtA [112,121], RmtB [122], ArmA [122,123] and RmtD which is commonly found together with the MBL SPM-1 in Brazil [124,125].

Advertisement

5. Treatment options for MDR Pseudomonas aeruginosa

Different combinations of the aforementioned mechanisms may be present in a single P. aeruginosa isolate leading to simultaneous resistance to various anti-pseudomonal compounds. The most potent combination is obviously that of a carbapenemase producing isolate usually enriched by resistance to quinolones and aminoglycosides leaving very limited options for antimicrobial treatment.

As far as newer carbapenem compounds are concerned, data suggest that doripenem does not offer advantages over other carbapenems against carbapenemase producing strains [126].

Tigecycline is an option for Gram-negative MDR pathogens but it cannot be used against P. aeruginosa, Morganella morganii, Proteus spp. and Providencia spp. because it is intrinsically vulnerable to their chromosomal-encoded efflux pumps [127].

Furthermore, time-kill studies on 12 MBL-producing P. aeruginosa isolates performed with aztreonam alone and in combination with ceftazidime and amikacin, showed bactericidal activity against one and eight isolates respectively. In the same study, colistin was bactericidal against all 12 isolates [128].

In fact, polymyxins and colistin in particular, are quite effective in the treatment of MDR P. aeruginosa infections [129,130]. The target of colistin is the bacterial cell membrane. More precisely, colistin interacts with the lipid A of lipopolysaccharides, allowing penetration through the outer membrane by displacing Ca2+ and Mg2+. The insertion between the phospholipids leads to loss of membrane integrity and consequent bacterial cell death [131]. There are reports of resistance to polymyxin B [132-134] and colistin [135-137] in clinical isolates but they remain to date relatively rare for P. aeruginosa [24]. While in many cases the mechanism of clinical polymyxin resistance is unknown, substitution of the lipopolysaccharide lipid A with aminoarabinose has been shown to contribute to polymyxin resistance in vitro [138] and cystic fibrosis isolates [139]. Colistin is frequently associated with nephro- and neurotoxicity but both these adverse effects seem to be dose-dependent and reversible [140].

Another interesting option for the treatment of MDR P. aeruginosa is fosfomycin, an old antibacteial that has regained attention because of its in vitro activity against such isolates [140]. Fosfomycin inactivates the enzyme pyruvil-transferase, which is required for the synthesis of the cell wall peptidoglycan. In a review of the existing fosfomycin studies, 81.1% of 1529 patients were successfully treated for infections caused by P. aeruginosa, Staphylococcus aureus, Staphylococcus epidermidis, Enterobacter spp. and Klebsiella spp. Fosfomycin was administered together with aminoglycosides, cephalosporins and penicillines [141]. More studies are needed however to determine the future role of fosfomycin against MDR P. aeruginosa isolates.

Advertisement

6. Combination therapy

The application of combination therapy instead of monotherapy in cases of non-MDR P. aeruginosa remains to date a controversial issue [14]. Combination treatment against MDR strains instead seems to be some times necessary (for example in cases of pan-resistance or resistance to all except a single agent). In such cases better results are expected by the additive or subadditive activity of a combination or by the enhancement of a single active agent by an otherwise inactive drug [142].

Several old and newer studies have showed the increased activity in vitro of various antibiotic combinations against MDR P. aeruginosa (Table 7) even though, the mechanisms of positive interaction between the various agents are rarely known [142].

Antibiotic combination References
Ticarcillin, Tobramycin, Rifampin [143]
Cephalosporins, Quinolones [144]
Ceftazidime, Colistin [145]
Macrolides, Tobramycin, Trimethoprim, Rifampin [146]
Polymyxin B, Rifampin [147]
Polymyxin B, Imipenem [148]
Colistin, Meropenem [149]

Table 7.

Enhanced activity of antibiotic combinations against MDR P. aeruginosa.

Advertisement

7. Conclusion

P. aeruginosa is a nosocomial pathogen of particular clinical concern not only because of its extraordinary resistance mechanisms armamentarium but also for its formidable ability to adapt very well to the hospital environment. There are important challenges in the treatment of MDR P. aeruginosa strains and their isolation in healthcare settings poses serious infection control issues. For these reasons, the prudent use of antibiotics, mainly those used as last resort treatment like carbapenems is of outmost importance in order to prevent evolutionary pressure that may lead to the emergence of highly resistant clones.

References

  1. 1. Slama TG. Gram-negative antibiotic resistance: there is a price to pay. Crit Care 2008;12. (Sup4) S4.
  2. 2. Kerr KG, Snelling AM. Pseudomonas aeruginosa: a formidable and ever-present adversary. J Hosp Infect 2009;73. 338–344.
  3. 3. Mauldin PD, Salgado CD, Hansen IS, et al. Attributable hospital cost and length of stay associated with health care-associated infections caused by antibiotic-resistant Gram- negative bacteria. Antimicrob Agents Chemother 2010;54. 109–115.
  4. 4. Tumbarello M, Repetto E, Trecarichi EM, et al. Multidrug- resistant Pseudomonas aeruginosa bloodstream infections: risk factors and mortality. Epidemiol Infect 2011;139. 1740-1749.
  5. 5. Lambert PA. Mechanisms of antibiotic resistance in Pseudomonas aeruginosa. J R Soc Med 2002;95. 22-26.
  6. 6. Moore NM, Flaws ML. Antimicrobial resistance mechanisms in Pseudomonas aeruginosa. Clin Lab Sci 2011; 24. 47-51.
  7. 7. Hancock REW, & Brinkman F. Function of Pseudomonas porins in uptake and efïlux. Annu Rev Microbiol 2002;56. 17-38.
  8. 8. Tamber S, Ochs MM, Hancock REW. Role of the novel OprD family of porins in nutrient uptake in Pseudomonas aeruginosa. J Bacteriol 2006;188. 45-54.
  9. 9. Lister PD, Wolter DJ, Hanson ND. Antibacterial-resistant Pseudomonas aeruginosa: Clinical impact and complex regulation of chromosomally encoded resistance mechanisms. Clin Micro Rev 2009;22. 582-610.
  10. 10. Strateva T, Yordanov D. Pseudomonas aeruginosa – a phenomenon of bacterial resistance. J Med Microbiol 2009;58. 1133–1148.
  11. 11. Schweizer HP. Efflux as a mechanism of resistance to antimicrobials in Pseudomonas aeruginosa and related bacteria: unanswered questions. Genet Mol Res 2003;2. 48-62.
  12. 12. Girlich D, Naas T, Nordmann P. Biochemical characterization of the naturally occurring oxacillinase OXA-50 of Pseudomonas aeruginosa. Antimicrob Agents Chemother 2004;48.2043–2048.
  13. 13. Kong KF, Jayawardena SR, Del Puerto A, et al. Characterization of poxB, a chromo-somal-encoded Pseudomonas aeru-ginosa oxacillinase. Gene 2005;358. 82–92.
  14. 14. Moore NM, Flaws ML. Treatment strategies and recommendations for Pseudomonas aeruginosa infections. Clin Lab Sci 2011;24. 52-56.
  15. 15. Tipper DJ. Mode of action of beta-lactam antibiotics. Pharmacol Ther 1985;27.1-35.
  16. 16. Hooper DC. Quinolone mode of action--new aspects. Drugs 1993;45. 8-14.
  17. 17. Poole K. Efflux-mediated antimicrobial resistance. J Antimicrob Chemother 2005;56. 20–51.
  18. 18. Kohler T, Van Delden C, Curty LK, et al. Overexpression of the MexEF-OprN multidrug efflux system affects cell-to-cell signaling in Pseudomonas aeruginosa. J Bacteriol 2001;183. 5213–5222.
  19. 19. Westbrock-Wadman S, Sherman DR, Hickey MJ, et al. Characterization of a Pseudomonas aeruginosa efflux pump contributing to aminoglycoside impermeability. Antimicrob Agents Chemother 2004;43. 2975–2983.
  20. 20. Pozzi G, Iannelli F, Oggioni MR, et al. Genetic elements carrying macrolide-efflux genes in streptococci. Curr. Drug Targets Infect Disord 2004; 4. 203–206.
  21. 21. Li XZ, Poole K, Nikaido H. Contributions of MexAB-OprM and an ErmE homolog to intrinsic resistance of Pseudomonas aeruginosa to aminoglycosides and dyes. Antimicrob Agents Chemother, 2003;47. 27–33.
  22. 22. Dozzo P, Moser, HE. New aminoglycoside antibiotics . Expert Opin Ther Pat 2010;20.1321-1341.
  23. 23. Pagkalis S, Mantadakis E, Mavros MN, et al. Pharmacological considerations for the proper clinical use of aminoglycosides. Drugs 2011;71. 2277-2294.
  24. 24. Poole K. Pseudomonas aeruginosa: resistance to the max. Front Microbiol 2011;2.65.
  25. 25. Ambler RP. The structure of beta-lactamases. Philos Trans R Soc Lond B Biol Sci 1980;289. 321-331.
  26. 26. Bush K, Jacoby GA, Medeiros AA. A functional classification scheme for beta-lactamases and its correlation with molecular structure. Antimicrob Agents Chemother 1995;39. 1211-1233.
  27. 27. Falagas ME, Karageorgopoulos DEJ. Extended-spectrum beta-lactamase-producing organisms. J Hosp Infect 2009;73.345-354.
  28. 28. Queenan AM, Bush K. Carbapenemases: the versatile beta-lactamases. Clin Microbiol 2007;20. 440-458.
  29. 29. Arora S, Bal, M. AmpC β-lactamase producing bacterial iso-lates from Kolkata hospital. Indian J Med Res 2005;122. 224–233.
  30. 30. Bratu S, Landman D, Gupta, J, et al. Role of AmpD, OprF and penicillin-binding proteins in β-lactam resistance in clinical isolates of Pseudomonas aeruginosa. J Med Microbiol 2007;56. 809–814.
  31. 31. Reinhardt A, Kohler T, Wood P, et al. Development and persistence of antimicrobial resistance in Pseudomonas aeruginosa: a longitudinal observation in mechanically ventilated patients. Antimicrob Agents Chemother 2007;51. 1341–1350.
  32. 32. Tam VH, Schilling AN, LaRocco MT, et al. Prevalence of AmpC over-expression in blood-stream isolates of Pseudomonas aeruginosa. Clin Microbiol Infect 2007;13, 413–418.
  33. 33. Drissi M, Ahmed ZB, Dehecq B, et al. Antibiotic susceptibility and mechanisms of β-lactam resistance among clinical strains of Pseudomonas aeruginosa: first report in Algeria. Med Mal Infect 2008;38. 187-191.
  34. 34. Vettoretti L, Floret N, Hocquet D, et al. Emergence of extensive-drug-resistant Pseudomonas aeruginosa in a French university hospital. Eur J Clin Microbiol Infect Dis 2009;28. 1217–1222.
  35. 35. Upadhyay S, Sen MR, Bhattacharjee A. Presence of different β-lactamase classes among clinical isolates of Pseudomonas aeruginosa expressing AmpC β-lactamase enzyme. J Infect Dev Ctries 2010;4. 239–242.
  36. 36. Xavier DE, Picao RC, Girardello R, et al. Efflux pumps expression and its association with porin down- regulation and β-lactamase production among Pseudomonas aeruginosa causing bloodstream infections in Brazil. BMC Microbiol 2010;10. 217.
  37. 37. Dunne WM Jr, Hardin DJ.J. Use of several inducer and substrate antibiotic combinations in a disk approximation assay format to screen for AmpC induction in patient isolates of Pseudomonas aeruginosa, Enterobacter spp., Citrobacter spp., and Serratia spp. Clin Microbiol 2005;43. 5945-5949.
  38. 38. Quale J, Bratu S, Gupta J, et al. Interplay of efflux system, ampC, and oprD expression in carbapenem resistance of Pseudomonas aeruginosa clinical isolates. Antimicrob Agents Chemother 2006;50. 1633-1641.
  39. 39. Giedraitienė A, Vitkauskienė A, Naginienė R, et al. Antibiotic resistance mechanisms of clinically important bacteria. Medicina (Kaunas) 2011;47.137-146.
  40. 40. Poirel L, Nordmann P. Acquired carbapenem-hydrolyzing β-lactamases and their genetic support. Curr Pharm Biotechnol 2002; 3. 117–127.
  41. 41. Castanheira M, Toleman MA, Jones RN, et al. Molecular characterization of a β-lactamase gene, blaGIM-1, encoding a new subclass of metallo-β-lactamase. Antimicrob Agents Chemother 2004;48. 4654–4661.
  42. 42. Walsh TR, Toleman MA, Poirel L, et al. Metallo-β-lactamases: the quiet before the storm? Clin Microbiol Rev 2005;18. 306–325.
  43. 43. Naas T, Aubert D, Lambert T, et al. Complex genetic structures with repeated elements, a sul-type class 1 integron, and the blaVEB extended-spectrum β-lactamase gene. Antimicrob Agents Chemother 2006;50. 1745–1752.
  44. 44. Bogaerts P, Bauraing C, Deplano A, et al. Emergence and dissemination of BEL- 1-producing Pseudomonas aeruginosa isolates in Belgium. Antimicrob Agents Chemother 2007;51. 1584–1585.
  45. 45. Gupta V. Metallo β-lactamases in Pseudomonas aeruginosa and Acinetobacter species. Expert Opin Investig Drugs 2008;17. 131–143.
  46. 46. Li H, Toleman MA, Bennett PM, et al. Complete Sequence of p07-406, a 24,179-base-pair plasmid harboring the blaVIM-7 metallo-β-lactamase gene in a Pseudomonas aeruginosa isolate from the United States. Antimicrob Agents Chemother 2008;52. 3099–3105.
  47. 47. Castanheira M, Bell JM, Turnidge JD. Carbapenem resistance among Pseudomonas aeruginosa strains from India: evidence for nationwide endemicity of multiple metallo-β-lactamase clones (VIM-2, -5, -6, and -11 and the newly characterized VIM-18). Antimicrob Agents Chemother 2009;53. 1225–1227.
  48. 48. Zhao WH, Chen G, Ito R, et al. Relevance of resistance levels to carbapenems and integron-borne blaIMP-1, blaIMP-7, blaIMP-10 and blaVIM-2 in clinical isolates of Pseudomonas aeruginosa. J Med Microbiol 2009;58. 1080–1085.
  49. 49. Kotsakis SD, Papagiannitsis CC, Tzelepi E, et al. GES-13, a β-lactamase variant pos-sessing Lys-104 and Asn-170 in Pseudomonas aeruginosa. Antimicrob Agents Chemother 2010;54. 1331–1333.
  50. 50. Cambray G, Guerout AM, Mazel D. Integrons. Annu Rev Genet 2010;44. 141–166.
  51. 51. Poirel L, Magalhaes M, Lopes M, et al. Molecular analysis of metallo-β-lactamase gene blaSPM-1-surrounding sequences from disseminated Pseudomonas aeruginosa isolates in Recife, Brazil. Antimicrob Agents Chemother 2004;48. 1406–1409.
  52. 52. Picao RC, Poirel L, Gales AC, et al. Diversity of β-lactamases produced by ceftazidime-resistant Pseudomonas aeruginosa isolates causing bloodstream infections in Brazil. Antimicrob Agents Chemother 2009;53. 3908–3913.
  53. 53. Picao RC, Poirel L, Gales AC, et al. Further identification of CTX-M-2 extended-spectrum β-lactamase in Pseudomonas aeruginosa. Antimicrob Agents Chemother 2009;53. 2225–2226.
  54. 54. Paterson DL, Bonomo RA. Extendedspectrum β-lactamases: a clinical update. Clin Microbiol Rev 2005;18. 657–686.
  55. 55. Lee K, Yum JH, Yong D, et al. Novel acquired metallo-β-lactamase gene, blaSIM-1, in a class 1 integron from Acinetobacter baumannii clinical isolates from Korea. Antimicrob Agents Chemother 2005;49. 4485–4491.
  56. 56. Sacha P, Wieczorek P, Hauschild T, et al. Metallo-beta-lactamases of Pseudomonas aeruginosa--a novel mechanism of resistance to beta-lactam antibiotics. Folia Histochem Cytobiol 2008;46. 137-142.
  57. 57. Pai H, Jacoby GA. Sequences of the NPS-1 and TLE-1 β-lactamase genes. Antimicrob Agents Chemother 2001;45. 2947–2948.
  58. 58. Kalai BS, Achour W, Bejaoui M, et al. Detection of SHV-1 β-lactamase in Pseudomonas aeruginosa strains by genetic methods. Pathol Biol (Paris) 2009;57. e73–75.
  59. 59. Celenza G, Pellegrini C, Caccamo M, et al. Spread of bla(CTX-M-type) and bla(PER-2) β-lactamase genes in clinical isolates from Bolivian hospitals. J Antimicrob Chemother 2006;57. 975–978.
  60. 60. Jiang X, Zhang Z., Li, M, et al. Detection of extended-spectrum β-lactamases in clinical isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother 2006;50. 2990–2995.
  61. 61. Rejiba S, Limam F, Belhadj C. Biochemical characterization of a novel extendedspectrum β-lactamase from Pseudomonas aeruginosa. Microb Drug Resist 2002;8. 9–13.
  62. 62. al Naiemi N, Duim B, Bart, A. A CTX-M extended-spectrum β-lactamase in Pseudomonas aeruginosa and Stenotrophomonas maltophilia. J Med Microbiol 2006;55. 1607–1608.
  63. 63. Sanschagrin F, Bejaoui N, Levesque RC. Structure of CARB-4 and AER-1 carbenicillin-hydrolyzing β-lactamases. Antimicrob Agents Chemother 42. 1998;1966–1972.
  64. 64. Villegas MV, Lolans K, Correa A, et al. First identification of Pseudomonas aeruginosa isolates producing a KPC-type carbapenem-hydrolyzing β-lactamase. Antimicrob Agents Chemother 2007;51. 1553–1555.
  65. 65. Wolter DJ, Khalaf N, Robledo IE, et al. Surveillance of carbapenem-resistant Pseudomonas aeruginosa isolates from Puerto Rican Medical Center Hospitals: dissemination of KPC and IMP-18 β-lactamases. Antimicrob Agents Chemother 2009;53. 1660–1664.
  66. 66. Bert F, Vanjak D, Leflon-Guibout V, et al. IMP-4-producing Pseudomonas aeruginosa in a French patient repatriated from Malaysia: impact of early detection and control measures. Clin Infect Dis 2007;44. 764–765.
  67. 67. Ryoo NH, Lee K, Lim JB, et al. Outbreak by meropenemresistant Pseudomonas aeruginosa producing IMP-6 metallo-β-lactamase in a Korean hospital. Diagn Microbiol Infect Dis 2009;63. 115–117.
  68. 68. Iyobe S, Kusadokoro H, Takahashi A, et al. Detection of a variant metallo-β-lactamase, IMP-10, from two unrelated strains of Pseudomonas aeruginosa and an Alcaligenes xylosoxidans strain. Antimicrob Agents Chemother 2002;46. 2014–2016.
  69. 69. Docquier JD, Riccio ML, Mugnaioli C, et al. IMP-12, a new plasmid-encoded metallo-β-lactamase from a Pseudomonas putida clinical isolate. Antimicrob Agents Chemother 2003;47. 1522–1528.
  70. 70. Garza-Ramos U, Morfin-Otero R, Sader HS, et al. Metallo-β-lactamase gene bla(IMP-15) in a class 1 integron, In95, from Pseudomonas aeruginosa clinical isolates from a hospital in Mexico. Antimicrob Agents Chemother 2008;52. 2943–2946.
  71. 71. Mendes RE, Toleman MA, Ribeiro J, et al. Integron carrying a novel metallo-β-lactamase gene, blaIMP-16, and a fused form of aminoglycoside-resistant gene aac(6’)-30/ aac(6’)-Ib’: report from the SENTRY Antimicrobial Surveillance Program. Antimicrob Agents Chemother 2004;48. 4693–4702.
  72. 72. Duljasz W, Gniadkowski M, Sitter S, et al. First organisms with acquired metallo-β-lactamases (IMP-13, IMP-22, and VIM-2) reported in Austria. Antimicrob Agents Chemother 2009;53. 2221–2222.
  73. 73. Koh TH, Wang GC, Sng LH. IMP-1 and a novel metallo-β-lactamase, VIM-6, in fluorescent Pseudomonads isolated in Singapore. Antimicrob Agents Chemother 2004;48. 2334–2336.
  74. 74. Siarkou VI, Vitti D, Protonotariou E, et al. Molecular epidemiology of outbreak-related Pseudomonas aeruginosa strains carrying the novel variant blaVIM-17 metallo-β-lactamase gene. Antimicrob Agents Chemother 2009;53. 1325–1330.
  75. 75. Toleman MA, Simm AM, Murphy TA, et al. Molecular characterization of SPM-1, a novel metallo-β-lactamase isolated in Latin America: report from the SENTRY antimicrobial surveillance programme. J Antimicrob Chemother 2002;50. 673–679.
  76. 76. Jovcic B, Lepsanovic Z, Suljagic V, et al. Emergence of NDM-1 Metallo-{beta}-Lactamase in Pseudomonas aeruginosa Clinical Isolates from Serbia. Antimicrob Agents Chemother 2011;55. 3929-3931.
  77. 77. Rodriguez-Martinez JM, Poirel L, Nordmann P. Extended-spectrum cephalosporinases in Pseudomonas aeruginosa. Antimicrob Agents Chemother 2009;53. 1766–1771.
  78. 78. Giuliani F, Docquier JD, Riccio ML, et al. OXA-46, a new class D β-lactamase of narrow substrate specificity encoded by a blaVIM-1-containing integron from a Pseudomonas aeruginosa clinical isolate. Antimicrob Agents Chemother 2005;49. 1973–1980.
  79. 79. Juan C, Mulet X, Zamorano L, et al. Detection of the novel extended spectrum β-lactamase (ESBL) OXA-161 from a plasmid-located integron in Pseudomonas aeruginosa clinical isolates in Spain. Antimicrob Agents Chemother 2009;53. 5288–5290.
  80. 80. Sevillano E, Gallego L, Garcia-Lobo JM. First detection of the OXA-40 carbapenemase in P. aeruginosa isolates, located on a plasmid also found in A. baumannii. Pathol Biol (Paris) 2009;57. 493–495.
  81. 81. Watanabe M, Iyobe S, Inoue M, et al. Transferable imipenem resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother 1991;35.147–151.
  82. 82. Lauretti L, Riccio ML, Mazzariol A, et al. Cloning and characterization of blaVIM, a new integron-borne metallo-β-lactamase gene from a Pseudomonas aeruginosa clinical isolate. Antimicrob Agents Chemother 1999;43. 1584–1590.
  83. 83. Salabi AE, Toleman MA, Weeks J, et al. First report of the metallo-beta-lactamase SPM-1 in Europe. Antimicrob Agents Chemother 2010;54. 582.
  84. 84. Yong D, Bell JM, Ritchie B, et al. A novel sub-group metallo-b-lactamase (MBL), AIM-1, emerges in Pseudomonas aeruginosa (PSA) from Australia. 47th Interscience Conference on Antimicrobial Agents and Chemotherapy. Chicago, IL, USA, 2007; Abstract C1–593.
  85. 85. Poirel L, Weldhagen GF, Naas T, et al. GES-2, a class A beta-lactamase from Pseudomonas aeruginosa with increased hydrolysis of imipenem. Antimicrob Agents Chemother 2001;45. 2598-2603.
  86. 86. Mavroidi A, Tzelepi E, Tsakris A, et al. An integron–associated β-lactamase (IBC-2) from Pseudomonas aeruginosa is a variant of the extended-spectrum β-lactamase IBC-1. J Antimicrob Chemother 2001;48. 627-630.
  87. 87. El Garch F, Bogaerts P, Bebrone C, et al. OXA-198, an acquired carbapenem-hydrolyzing class D beta-lactamase from Pseudomonas aeruginosa. Antimicrob Agents Chemother 2011;55. 4828-4833.
  88. 88. Poole K. Resistance to β-lactam antibiotics. Cell Mol Life Sci 2004;61. 2200–2223.
  89. 89. Tomas M, Doumith M, Warner M, et al. Efflux pumps, OprD porin, AmpC β-lactamase, and multiresistance in Pseudomonas aeruginosa isolates from cystic fibrosis patients. Antimicrob Agents Chemother 2010;54. 2219–2224.
  90. 90. Yoneda K, Chikumi H, Murata T, et al. Measurement of Pseudomonas aeruginosa multidrug efflux pumps by quantitative real-time polymerase chain reaction. FEMS Microbiol Lett 2005;243. 125-131.
  91. 91. Farra A, Islam S, Strålfors A, et al. Role of outer membrane protein OprD and penicillin-binding proteins in resistance of Pseudomonas aeruginosa to imipenem and meropenem. Int J Antimicrob Agents 2008;31. 427-433.
  92. 92. Gutiérrez O, Juan C, Cercenado E, et al. Molecular epidemiology and mechanisms of carbapenem resistance in Pseudomonas aeruginosa isolates from Spanish hospitals. Antimicrob Agents Chemother 2007;51. 4329-4335.
  93. 93. Horii T, Muramatsu H, Morita M, et al. Characterization of Pseudomonas aeruginosa isolates from patients with urinary tract infections during antibiotic therapy. Microb Drug Resist 2003;9. 223-229.
  94. 94. Wang J, Zhou JY, Qu TT, et al. Molecular epidemiology and mechanisms of carbapenem resistance in Pseudomonas aeruginosa isolates from Chinese hospitals. Int J Antimicrob Agents 2010;3. 486–491.
  95. 95. Carmeli Y, Troillet N, Eliopoulos GM, et al. Emergence of antibiotic-resistant Pseudomonas aeruginosa: comparison of risks associated with different antipseudomonal agents. Antimicrob Agents Chemother 1999;43. 1379-1382.
  96. 96. Jacoby GA. Mechanisms of resistance to quinolones. Clin Infect Dis 2005;41. 120–126.
  97. 97. Drlica K, Hiasa H, Kerns R, et al. Quinolones: action and resistance updated. Curr Top Med Chem 2009;9. 981–998.
  98. 98. Higgins PG, Fluit AC, Milatovic D, et al. Mutations in GyrA, ParC, MexR and NfxB in clinical isolates of Pseudomonas aeruginosa. Int J Antimicrob Agents 2003;21. 409–413.
  99. 99. Henrichfreise B, Wiegand I, Pfister W, et al. Resistance mechanisms of multiresistant Pseudomonas aeruginosa strains from Germany and correlation with hypermutation. Antimicrob Agents Chemother 2007;51. 4062–4070.
  100. 100. Rejiba S, Aubry A, Petitfrere S, et al. Contribution of parE mutation and efflux to ciprofloxacin resistance in Pseudomonas aeruginosa clinical isolates. J Chemother 2008;20. 749–752.
  101. 101. Lee JK, Lee YS, Park YK, et al. Alterations in the GyrA and GyrB subunits of topoisomerase II and the ParC and ParE subunits of topoi-somerase IV in ciprofloxacin-resistant clinical isolates of Pseudomonas aeruginosa. Int J Antimicrob Agents 2005;25. 290–295.
  102. 102. Muramatsu H, Horii T, Takeshita A, et al. Characterization of fluoroquinolone and carbapenem susceptibilities in clinical isolates of levofloxacin-resistant Pseudomonas aeruginosa. Chemotherapy 2005;51. 70–75.
  103. 103. Nakano M, Deguchi T, Kawamura T, et al. Mutations in the gyrA and parC genes in fluoroquinolone-resistant clinical isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother 1997;41. 2289–2291.
  104. 104. Schwartz T, Volkmann H, Kirchen S, et al. Real-time PCR detection of Pseudomonas aeruginosa in clinical and municipal wastewater and genotyping of the ciprofloxacin- resistant isolates. FEMS Microbiol Ecol 2006;57. 158–167.
  105. 105. Poole K. Efflux-mediated resistance to fluoroquinolones in Gram-negative bacteria. Antimicrob Agents Chemother 2000;44. 2233–2241.
  106. 106. Wolter DJ, Smith-Moland E, Goering RV, et al. Multidrug resistance associated with mexXY expression in clinical isolates of Pseudomonas aeruginosa from a Texas hospital. Diagn Microbiol Infect Dis 2004;50. 43–50.
  107. 107. Zhanel GG, Hoban DJ, Schurek K, et al. Role of efflux mechanisms on fluoroquinolone resistance in Streptococcus pneumoniae and Pseudomonas aeruginosa. Int J Antimicrob Agents 2004;24. 529–535.
  108. 108. Ramirez MS, Tolmasky ME. Aminoglycoside modifying enzymes. Drug Resist 2010;13. 151–171.
  109. 109. Poole K. Aminoglycoside resistance in Pseudomonas aeruginosa. Antimicrob Agents Chemother 2005;49. 479–487.
  110. 110. Kettner M, Milosovic P, Hletkova M, et al. Incidence and mechanisms of aminoglycoside resistance in Pseudomonas aeruginosa serotype O11 isolates. Infection 1995;23. 380–383.
  111. 111. Kim JY, Park YJ, Kwon HJ, et al. Occurrence and mechanisms of ami-kacin resistance and its association with β-lactamases in Pseudomonas aeruginosa: a Korean nationwide study. J Antimicrob Chemother 2008;62. 479–483.
  112. 112. Jin JS, Kwon KT, Moon DC, et al. Emergence of 16S rRNA methylase rmtA in colistin-only- sensitive Pseudomonas aeruginosa in South Korea. Int J Antimicrob Agents 2009;33. 490–491.
  113. 113. Riccio ML, Pallecchi L, Fontana R, et al. In70 of plasmid pAX22, a blaVIM-1-containing integron carrying a new aminogly-coside phosphotransferase gene cas-sette. Antimicrob Agents Chemother 2001;45. 1249–1253.
  114. 114. Shaw KJ, Munayyer H, Rather PN, et al. Nucleotide sequence analysis and DNA hybridization studies of the ant(4′)-IIa gene from Pseudomonas aeruginosa. Antimicrob Agents Chemother 1993;37. 708–714.
  115. 115. Sabtcheva S, Galimand M, Gerbaud G, et al. Aminoglycoside resistance gene ant(4′)-IIb of Pseudomonas aeruginosa BM4492, a clinical isolate from Bulgaria. Antimicrob Agents Chemother 2003;47. 1584–1588.
  116. 116. Miller GH, Sabatelli FJ, Naples L, et al. Resistance to aminoglycosides in Pseudomonas. Aminoglycoside Resistance Study Groups. Trends Microbiol 1994;2. 347–353.
  117. 117. Hachler H, Santanam P, Kayser FH. Sequence and characterization of a novel chromosomal aminogly-coside phosphotransferase gene, aph (3′)-IIb, in Pseudomonas aeruginosa. Antimicrob Agents Chemother 1996;40. 1254–1256.
  118. 118. Sobel ML, McKay GA, Poole K. Contribution of the MexXY multidrug transporter to aminogly-coside resistance in Pseudomonas aeruginosa clinical isolates. Antimicrob Agents Chemother 2003;47. 3202–3207.
  119. 119. Hocquet D, Nordmann P, El Garch, et al. Involvement of the MexXY- OprM efflux system in emergence of cefepime resistance in clinical strains of Pseudomonas aeruginosa. Antimicrob Agents Chemother 2006;50. 1347–1351.
  120. 120. Islam S, Oh H, Jalal S, et al. Chromosomal mechanisms of aminoglycoside resistance in Pseudomonas aeruginosa isolates from cystic fibrosis patients. Clin Microbiol Infect 2009;15. 60–66.
  121. 121. Yamane K, Doi Y, Yokoyama K, et al. Genetic environments of the rmtA gene in Pseudomonas aeruginosa clinical isolates. Antimicrob Agents Chemother 2004;48. 2069–2074.
  122. 122. Zhou Y, Yu H, Guo Q, et al. Distribution of 16S rRNA methylases among different species of Gram- negative bacilli with high-level resist-ance to aminoglycosides. Eur J Clin Microbiol Infect Dis 2010;29. 1349–1353.
  123. 123. Gurung M, Moon DC, Tamang MD, et al. Emergence of 16S rRNA methylase gene armA and cocarriage of blaIMP-1 in Pseudomonas aeruginosa isolates from South Korea. Diagn Microbiol, Infect Dis 2010;68. 468–470.
  124. 124. Doi Y, Ghilardi AC, Adams J, et al. High prevalence of metallo-β-lactamase and 16S rRNA methylase coproduction among imipenem- resistant Pseudomonas aeruginosa isolates in Brazil. Antimicrob Agents Chemother 2007;51. 3388–3390.
  125. 125. Lincopan N, Neves P, Mamizuka EM, et al. Balanoposthitis caused by Pseudomonas aeruginosa co-producing metallo-β-lactamase and 16S rRNA methylase in children with hematological malignancies. Int J Infect Dis 2010;14. 344–347.
  126. 126. Mushtaq S, Ge Y, Livermore DL. Comparative activities of doripenem versus isolates, mutants, and transconjugants of Enterobacteriaceae and Acinetobacter spp. with characterized b-lactamases. Antimicrob Agents Chemother 2004;48. 1113–1119.
  127. 127. Dean CR, Visalli MA, Projan SJ, et al. Efflux-mediated resistance to tigecycline (GAR-936) in Pseudomonas aeruginosa PAO1. Antimicrob Agents Chemother 2003;47. 972-978.
  128. 128. Oie S, Fukui Y, Yamamoto M, et al. In vitro antimicrobial effects of aztreonam, colistin, and the 3-drug combination of aztreonam, ceftazidime and amikacin on metallo b-lactamase-producing Pseudomonas aeruginosa. BMC Infect Dis 2009;9. 123.
  129. 129. Montero M, Horcajada JP, Sorli L, et al. Effectiveness and safety of colistin for the treatment of multidrug-resistant Pseudomonas aeruginosa infections. Infection 2009;37. 461–465.
  130. 130. Falagas ME, Rafailidis PI, Ioannidou E, et al. Colistin therapy for micro-biologically documented multidrug- resistant Gram-negative bacterial infections: a retrospective cohort study of 258 patients. Int J Antimicrob Agents 2010;35. 194–199.
  131. 131. Giamarellou H. Treatment options for multidrug-resistant bacteria. Expert Rev Anti Infect Ther 2006;4. 601-618.
  132. 132. Landman D, Bratu S, Alam M, et al. Citywide emergence of Pseudomonas aeruginosa strains with reduced susceptibility to poly-myxin B. J Antimicrob Chemother 2005;55. 954–957.
  133. 133. Abraham N, Kwon DH. A single amino acid substitution in PmrB is associated with polymyxin B resistance in clinical isolate of Pseudomonas aeruginosa. FEMS Microbiol Lett 2009;298. 249–254.
  134. 134. Barrow K, Kwon DH. Alterations in two-component regulatory systems of phoPQ and pmrAB are associated with polymyxin B resistance in clinical isolates of Pseudomonas aeruginosa. Antimicrob Agents Chemother 2009;53. 5150–5154.
  135. 135. Johansen HK, Moskowitz SM, Ciofu O, et al. Spread of colistin resistant non- mucoid Pseudomonas aeruginosa among chronically infected Danish cystic fibrosis patients. J Cyst Fibros 2008;7. 391–397.
  136. 136. Matthaiou DK, Michalopoulos A, Rafailidis PI, et al. Risk factors associated with the isolation of colistin-resistant Gram-negative bacteria: a matched case-control study. Crit Care Med 2008;36. 807–811.
  137. 137. Samonis G, Matthaiou DK, Kofteridis D, et al. In vitro susceptibility to various antibiotics of colistin-resistant Gram-negative bacterial isolates in a general tertiary hospital in Crete, Greece. Clin Infect Dis 2010;5. 1689–1691.
  138. 138. Moskowitz SM, Ernst RK, Miller SI. PmrAB, a two-component regulatory system of Pseudomonas aeruginosa that modulates resistance to cationic antimicrobial peptides and addition of aminoarabinose to lipid A. J Bacteriol 2004;186. 575–579.
  139. 139. Ernst RK, Yi EC, Guo L, et al. Specific lipopolysaccha-ride found in cystic fibrosis airway Pseudomonas aeruginosa. Science 1999;286. 1561–1565.
  140. 140. Giamarellou H, Poulakou G. Multidrug-resistant Gram-negative infections: what are the treatment options? Drugs 2009;69. 1879–1901.
  141. 141. Falagas ME, Giannopoulou KP, Kokolakis GN, et al. Fosfomycin: use beyond urinary tract and gastrointestinal infections. Clin Infect Dis 2008;46. 1069-1077
  142. 142. Rahal JJ. Novel antibiotic combinations against infections with almost completely resistant Pseudomonas aeruginosa and Acinetobacter species. Clin Infect Dis 2006; 43. Suppl 2:S95-99.
  143. 143. Zuravleff JJ, Yu VL, Yee RB. Ticarcillin-tobramycin-rifampin: in vitro synergy of the triple combination against Pseudomonas aeruginosa. J Lab Clin Med 1983;101. 896–902.
  144. 144. Fish DN, Choi MK, Jung R. Synergic activity of cephalosporins plus fluoroquinolones against Pseudomonas aeruginosa with resistance to one or both drugs. J Antimicrob Chemother 2002;50. 1045–1049.
  145. 145. Gunderson BW, Ibrahim KH, Hovde LB, et al. Synergistic activity of colistin and ceftazidime against multiantibiotic-resistant Pseudomonas aeruginosa in an in vitro pharmacodynamic model. Antimicrob Agents Chemother 2003;47. 905–909.
  146. 146. Saiman L, Chen Y, San Gabriel P, et al. Synergistic activities of macrolides antibiotics against Pseudomonas aeruginosa, Burkholderia cepacia, Stenotrophomonas maltophilia, and Alcaligines xylosoxidans isolated from patients with cystic fibrosis. Antimicrob Agents Chemother 2002;46. 1105–1107.
  147. 147. Perez Urena MT, Barasoain I, Espinosa M, et al. Evaluation of different antibiotic actions combined with rifampicin. Chemotherapy 1975;27. 82–89.
  148. 148. Chini NX, Scully B, DellaLatta P. Synergy of polymyxin B with imipenem and other antimicrobial agents against Acinetobacter, Klebsiella, and Pseudomonas species. Program and abstracts of the 38th Interscience Conference on Antimicrobial Agents and Chemotherapy (San Diego). Washington, DC: American Society for Microbiology 1998; Abstract E-56.
  149. 149. Pankuch GA, Lin G, Seifect H, et al. Activity of meropenem with and without ciprofloxacin and colistin against Pseudomonas aeruginosa and Acinetobacter baumannii. Antimicrob Agents Chemother 2008;52. 333-336.

Written By

Georgios Meletis and Maria Bagkeri

Submitted: 25 August 2012 Published: 29 May 2013