of carbapenem resistance in mechanisms and dynamics e s c

58
ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2014 Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Medicine 998 Mechanisms and Dynamics of Carbapenem Resistance in Escherichia coli MARLEN ADLER ISSN 1651-6206 ISBN 978-91-554-8950-2 urn:nbn:se:uu:diva-221432

Upload: others

Post on 22-Mar-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2014

Digital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Medicine 998

Mechanisms and Dynamicsof Carbapenem Resistance inEscherichia coli

MARLEN ADLER

ISSN 1651-6206ISBN 978-91-554-8950-2urn:nbn:se:uu:diva-221432

Dissertation presented at Uppsala University to be publicly examined in B42, BMC,Husargatan 3, Uppsala, Thursday, 5 June 2014 at 09:00 for the degree of Doctor ofPhilosophy (Faculty of Medicine). The examination will be conducted in English. Facultyexaminer: Ph. D. Josep Casadesús (University of Seville).

AbstractAdler, M. 2014. Mechanisms and Dynamics of Carbapenem Resistance in Escherichia coli.Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Medicine998. 51 pp. Uppsala: Acta Universitatis Upsaliensis. ISBN 978-91-554-8950-2.

The emergence of extended spectrum β-lactamase (ESBL) producing Enterobacteriaceaeworldwide has led to an increased use of carbapenems and may drive the development ofcarbapenem resistance. Existing mechanisms are mainly due to acquired carbapenemases or thecombination of ESBL-production and reduced outer membrane permeability. The focus of thisthesis was to study the development of carbapenem resistance in Escherichia coli in the presenceand absence of acquired β-lactamases. To this end we used the resistance plasmid pUUH239.2that caused the first major outbreak of ESBL-producing Enterobacteriaceae in Scandinavia.

Spontaneous carbapenem resistance was strongly favoured by the presence of the ESBL-encoding plasmid and different mutational spectra and resistance levels arose for differentcarbapenems. Mainly, loss of function mutations in the regulators of porin expression causedreduced influx of antibiotic into the cell and in combination with amplification of β-lactamasegenes on the plasmid this led to high resistance levels. We further used a pharmacokineticmodel, mimicking antibiotic concentrations found in patients during treatment, to test whetherertapenem resistant populations could be selected even at these concentrations. We found thatresistant mutants only arose for the ESBL-producing strain and that an increased dosage ofertapenem could not prevent selection of these resistant subpopulations. In another study wesaw that carbapenem resistance can even develop in the absence of ESBL-production. We foundmutants in export pumps and the antibiotic targets to give high level resistance albeit with highfitness costs in the absence of antibiotics. In the last study, we used selective amplification ofβ-lactamases on the pUUH239.2 plasmid by carbapenems to determine the cost and stability ofgene amplifications. Using mathematical modelling we determined the likelihood of evolutionof new gene functions in this region. The high cost and instability of the amplified state makesde novo evolution very improbable, but constant selection of the amplified state may balancethese factors until rare mutations can establish a new function.

In my studies I observed the influence of β-lactamases on carbapenem resistance and sawthat amplification of these genes would further contribute to resistance. The rapid disappearanceof amplified arrays of resistance genes in the absence of antibiotic selection may lead to theunderestimation of gene amplification as clinical resistance mechanism. Amplification of β-lactamase genes is an important stepping-stone and might lead to the evolution of new resistancegenes.

Keywords: carbapenem, antibiotic resistance, fitness cost, ESBLs, penicillin-binding proteins,gene amplification

Marlen Adler, Department of Medical Biochemistry and Microbiology, Box 582, UppsalaUniversity, SE-75123 Uppsala, Sweden.

© Marlen Adler 2014

ISSN 1651-6206ISBN 978-91-554-8950-2urn:nbn:se:uu:diva-221432 (http://urn.kb.se/resolve?urn=urn:nbn:se:uu:diva-221432)

Not everything that counts can be counted, and not everything that

can be counted counts.

–Albert Einstein

Für Ellen und Roland Adler.

List of Papers

This thesis is based on the following papers, which are referred to in the text by their Roman numerals.

I Adler M, Anjum M, Andersson DI, Sandegren L. (2012) Influence of

acquired β-lactamases on the evolution of spontaneous carbapenem resistance in Escherichia coli. J Antimicrob Chemother 68: 51-59

II Tängdén T, Adler M, Cars O, Sandegren L, Löwdin E. (2013) Fre-

quent emergence of porin-deficient subpopulations with reduced car-bapenem susceptibility in extended-spectrum-β-lactamase-producing Escherichia coli during exposure to ertapenem in an in vitro pharma-cokinetic model. J Antimicrob Chemother 68(6):1319–1326

III Adler M, Anjum M, Andersson DI, Sandegren L. Mutations in PBP2,

PBP3 and AcrB contribute to high-level carbapenem resistance in Escherichia coli. Manuscript

IV Adler M, Anjum M, Berg OG, Andersson DI, Sandegren L. (2014)

High fitness costs and instability of gene duplications reduce rates of evolution of new genes by duplication-divergence mechanisms. Mol Biol Evol doi:10.1093/molbev/msu111

Reprints were made with permission from the respective publishers.

Contents

Introduction ......................................................................................................... 9  Antibiotics .................................................................................................... 10  

Resistance to antibiotics ........................................................................... 11  Cost of resistance ..................................................................................... 13  

β-lactam antibiotics ...................................................................................... 14  Carbapenems ............................................................................................ 18  

β-lactam resistance ....................................................................................... 20  Role of outer membrane proteins in resistance ........................................ 20  Inactivation by β-lactamases .................................................................... 22  Inactivation by Carbapenemases .............................................................. 25  

ESBL plasmid pUUH239.2 .......................................................................... 26  Gene duplication and amplification .............................................................. 28  

Dynamics ................................................................................................. 28  Mechanism of formation and loss of GDA .............................................. 30  Evolution of new genes by duplication-divergence ................................. 30  

Escherichia coli and Klebsiella pneumoniae as pathogens .......................... 32  

Present Investigations ........................................................................................ 33  ESBL-plasmid influences evolution of carbapenem resistance .................... 33  Ertapenem resistance due to pUUH239.2 in a pharmacokinetic model ....... 34  Altered PBPs and drug efflux cause high-level carbapenem resistance ....... 35  Cost and instability of GDA limit evolution of new genes .......................... 36  Concluding Remarks .................................................................................... 38  

Future Perspectives ........................................................................................... 39  

Deutsche Zusammenfassung ............................................................................. 41  

Acknowledgments ............................................................................................. 44  

References ......................................................................................................... 46  

Abbreviations

ARDB Antibiotic resistance gene database bp Base pair DNA Deoxyribonucleic acid E. coli Escherichia coli ESBL Extended spectrum β-lactamase EUCAST European committee on antimicrobial susceptibility testing GDA Gene duplication and amplification HGT Horizontal gene transfer HMM High-molecular mass IAD Innovation-amplification-divergence IS Insertion sequence K. pneumoniae Klebsiella pneumoniae kb Kilo base pair LMM Low-molecular mass Mb Mega base pair MIC Minimal inhibitory concentration P. aeruginosa Pseudomonas aeruginosa PBP Penicillin-binding protein PFGE Pulsed-field gel electrophoresis qRT-PCR Quantitative real-time polymerase chain reaction rhs Recombination hot spot RNA Ribonucleic acid S. typhimurium Salmonella enterica serovar Typhimurium strain LT2 S. aureus Staphylococcus aureus S. cattleya Streptomyces cattleya SIR Sensitive-intermediate-resistant S. pneumoniae Streptococcus pneumoniae

9

Introduction

The discovery of antibiotics is often referred to as the greatest achievement of the 20th century. Together with the implications of germ theory by Louis Pasteur (1864), antibiotics are one of the main contributors to the increased human life expectancy of today. The introduction of antibiotics revolution-ized medicine and we are relying on antibiotic therapy for treatment of inju-ries and minor infections, but also for deep surgery, transplantation, chemo-therapy, neonatal care and prosthetic surgery. However, resistance to antibi-otic exposure was observed already before the clinical introduction of these compounds and quickly developed into a serious health care problem. We underestimated the genetic capacity of microbes to evolve and spread re-sistance genes among themselves and were not aware of the vast amount of resistance genes that are naturally present after millions of years of microbial evolution (more than 23,000 resistance genes, ARDB – antibiotic resistance gene database, http://ardb.cbcb.umd.edu/). Bacteria benefitted from the mis-use of antibiotics, not only as over the counter drugs but also for growth promotion in livestock, and we witnessed evolution at its best during less than 100 years of antibiotic therapy. Fortunately, not all predicted resistance genes found their way into the genomes of potential pathogenic bacteria or have spread sufficiently to reach epidemic magnitudes. It is therefore imper-ative to describe, understand and predict microbial resistance mechanisms to limit their spread and develop strategies to prolong the life span of antibiot-ics.

In this thesis I studied the mechanisms by which Escherichia coli can de-velop resistance to carbapenems, a group of last resort β-lactam antibiotics aimed for treatment of otherwise multi-resistant bacteria, in the laboratory and at concentrations typically found in patient serum during treatment. De-tailed studies of a frequent bacterial adaptation mechanism provided addi-tional understanding of the evolution of new resistance genes. The results presented in this thesis may be important to decide antibiotic treatment re-gimes and also give insights into bacterial evolution in general.

10

Antibiotics Antibiotics (Greek anti, “against”, bios, “life”) are chemicals that are used clinically to treat bacterial infections. Most antibiotics are initially derived from natural substances produced by microorganisms such as bacteria and fungi to inhibit the growth of other microbes. This effect can be by growth inhibition (bacteriostatic) or killing of other microbes (bactericidal). In addi-tion to the great number of naturally occurring antibiotics identified, semi-synthetic drugs have been developed through chemical modification of natu-ral products. Natural and semi-synthetic products make up the largest pro-portion of antibiotics in therapeutic use today, but three classes of truly syn-thetic antibiotics also exist: sulfa drugs, quinolones and oxazolidinones. Depending on their chemical characteristics, antibiotics exert their effect on differing sets of bacterial species. Narrow-spectrum antibiotics affect a lim-ited range of bacterial species, whereas broad-spectrum antibiotics affect a wide range of microbes including Gram-positive and Gram-negative bacte-ria. The development of antimicrobial drugs has greatly enhanced the control of infectious diseases and facilitated the improvement of advanced invasive medicine.

The most important characteristic to predict the outcome of antimicrobial therapy is the specific antibiotic concentration that is necessary to inhibit growth of a pathogen. This is called the minimal inhibitory concentration (MIC) and can be determined by three methods: i) In disc diffusion tests, antibiotics diffusing from a paper disc cause a zones of growth inhibition and the size of this zone can be compared to reference strains or related to treatment outcome. ii) Growth of bacteria in broth with serial dilutions of antibiotics can be used to determine the MIC (Andrews, 2001). iii) Commer-cially available ‘Etests’ offer a gradient of antibiotic concentration on a plas-tic strip. The antibiotic diffuses off the strip causing a zone with growth inhibition. The inhibiting concentration can be read from the printed scale as the lowest concentration that inhibits growth. Static or kinetic time-kill ex-periments cannot be used to measure the MIC, but are useful to study the effects of stable or varying antibiotic concentrations on bacterial survival. The so-called SIR-system (sensitive-intermediate-resistant) uses two empiri-cal determined antibiotic concentrations to predict the outcome of antimi-crobial therapy. These breakpoint concentrations are called clinical break-points. The outcome of antibiotic treatment is unclear if the MIC of the in-fecting bacterium exceeds the first clinical breakpoint and pathogens that can grow in the presence of antibiotic concentrations above the second clini-cal breakpoint are referred to as resistant. The European committee on anti-microbial susceptibility testing (EUCAST) is working to update and unify clinical breakpoints throughout Europe.

To selectively kill bacterial cells while not harming host cells, clinically useful antibiotics must target essential bacterial pathways not present or

11

sufficiently different from those present in eukaryotic cells. Known targets involve: i) DNA replication, ii) RNA synthesis, iii) Protein synthesis, iv) The cell membrane, and v) Cell wall synthesis (Walsh and Walsh, 2003; Kohanski et al., 2010). DNA replication is inhibited by fluoroquinolones. These antibiotics induce DNA breaks and blockage of the DNA replication forks by inhibiting topoisomerase II and IV. Trimethoprim and sulfamethox-azole inhibit two essential enzymes in folic acid metabolism and thereby block nucleotide biosynthesis. Rifamycins bind to the β-subunit of the DNA-dependent RNA polymerase and inhibit RNA transcription. Inhibition of the ribosome impedes protein synthesis. Macrolides, lincosamides, strepto-gramins, amphenicols and oxazolidinones block the 50S ribosomal subunit and tetracyclines and aminoglycosides block the 30S ribosomal subunit. Antibiotics such as fosfomycin and bacitracin inhibit the synthesis of cell wall precursors. Cell wall biosynthesis is targeted by glycopeptides and β-lactams. Glycopeptides such as vancomycin bind to the D-alanyl-D-alanine peptidoglycan tail making it inaccessible for both transglycosylation and transpeptidation. β-lactams bind to the enzymes responsible for transpepti-dation, the transpeptidases, and form inactive enzyme complexes. This pre-vents effective peptidoglycan cross-linking. The focus of this thesis will be on β-lactams. Members of this potent class of antibiotics and their mode of action will be discussed in more detail in later chapters (Page 14).

Resistance to antibiotics In the environment microbes have evolved together for millions of years, releasing substances that serve to provide growth advantages over neigh-bouring microbes (among other functions). Consequently, microbial com-munities were exposed to naturally produced antibiotic substances long be-fore their clinical introduction. Resistance genes are thought to have evolved from enzymes with low binding affinity or moderate catalytic activity against an antibiotic, or originate from antibiotic producing bacteria. Studies have shown that the resistome (all antibiotic resistance genes and their pre-cursors in pathogenic and non-pathogenic bacteria) confers resistance to all known antibiotics, even those that environmental bacteria were never ex-posed to. It is even suggested that environmental organisms are by default drug resistant (Wright, 2007). This is threatening because genetic material can be transferred between organisms and species. The introduction of anti-biotics into clinical use conferred strong selective advantage to pathogens that were able to incorporate natural resistance genes into their genome.

DNA can be mobilized and transferred between organisms and species by horizontal gene transfer (HGT). HGT is proposed to be a major driving force in bacterial evolution and transformation, conjugation, transduction, and other mechanisms are known (Popa and Dagan, 2011). The uptake of DNA from the environment and its integration into the genome is called transfor-

12

mation. Some bacteria, for example Streptococcus pneumoniae are naturally competent and able to take up exogenous DNA (Johnsborg and Håvarstein, 2009). Conjugation is the process of DNA transfer via cell-to-cell contact. Here a so-called pilus from the donor cell establishes cell contact, initiating transfer of plasmid DNA or other mobile genetic elements (Frost and Koraimann, 2010; Wozniak and Waldor, 2010). The transfer of DNA be-tween viruses and bacteria is called transduction. Occasionally, bacterial viruses may package and transfer parts of the bacterial genome to a different cell. The potential for DNA mobilisation together with selective pressure from antibiotics in the environment and hospital settings, provide a strong possibility for the spread of resistance genes throughout microbial popula-tions.

Antibiotic resistance can also be acquired through de novo mutation of genes in the bacterial chromosome. Spontaneous mutations, such as nucleo-tide substitutions, frameshifts, deletions, inversions and amplifications, oc-cur as consequences of exogenous DNA damaging agents, endogenous agents (such as damaged bases), or DNA polymerase errors (Rosche and Foster, 2000). Efficient repair mechanisms have evolved to limit the number of spontaneous mutations, because the majority of mutations will be disad-vantageous and potentially lethal. These mechanism include the DNA poly-merase proofreading function, methyl-directed mismatch repair, the nucleo-tide excision repair systems, and various base excision repair systems and recombinational repair (Lindahl and Wood, 1999). For evolution to occur a certain amount of genetic diversity is needed, because in rare cases cells acquire beneficial mutations that confer a selective advantage over other cells in the population. In the presence of antibiotics, resistant mutants will be able to grow whereas sensitive cells will be inhibited or killed, leading to an enrichment of resistant cells in the population.

To date resistance to all available antibiotics can be observed in patho-genic bacteria. A main mechanism is the decreased uptake of antibiotics into the bacterial cell. Many small antibiotics, for instance β-lactams, enter the cell through water filled transport channels called porins (Pagès et al., 2008; Delcour, 2009). Decreases in the number of channels, changes of porin structure or expression of a different kind of porin (shift from general porins to specific porins, or from porins with wide diameter to porins with smaller diameter) can lead to resistance (Fernández and Hancock, 2012). After crossing the membrane, antibiotics and other toxic compounds can be recog-nized and exported out of the cell. Numerous transporters spanning the inner membrane or transporter complexes spanning both the inner and outer mem-brane can extrude antibiotics from the cytoplasm or periplasm (Piddock, 2006a; Piddock, 2006b). Furthermore, degradation or modification of the antibiotic can reduce the effective antibiotic concentration and lead to re-sistance and here hydrolysis of the β-lactam ring by β-lactamases is one of

13

the most important examples (Jacoby and Munoz-Price, 2005; Queenan and Bush, 2007; Livermore, 2008). Resistance can also be achieved by modifica-tion or replacement of the antibiotic target so it is not longer recognized by the drug (Hiramatsu et al., 2001; Lambert, 2005; Yamachika et al., 2013). More recently altered expression of target genes and amplification of re-sistance genes has been associated with resistance development (Sandegren and Andersson, 2009; Paulander et al., 2010).

Cost of resistance In the majority of cases mutations that confer resistance are associated with a cost in the absence of antibiotics. It is not always known which process confers the cost, and several processes may be involved simultaneously. One can imagine costs from changes in essential proteins that lead to lower pro-tein activity, from the use of the cell’s replication machinery to replicate resistance conferring plasmids, or from changes in the expression of regula-tory processes (Andersson and Levin, 1999; Andersson and Hughes, 2010). These costs can affect how well resistant bacteria are transmitted or cleared from the host, or affect their ability to compete with non-resistant cells. Costs are sometimes dependent on the environment and can vary with pH, temperature, osmolarity and nutrient availability. Thus, identifying fitness costs under standard laboratory conditions is not always straightforward, and repeated tests under a range of conditions may be required. Costs can be determined in different ways. For example, one can compare the maximum exponential growth rate of a mutant with that of the wild type. This gives valuable information about the growth potential of the mutant. Another more comprehensive method is to compete isogenic strains under different growth conditions, taking into account additional components, such as time spent in the lag phase, utilization of resources and survival during the stationary phase. Using technology such as flow cytometry and fluorescently labelled bacteria a large sample of bacterial population can be analysed, reducing experimental error and allowing the detection of differences in fitness costs as low as 0.3% (Lind et al., 2010). Competition of isogenic strains in exper-imental animals such as mice would represent an environment closer to that present in the human host, but this method is ethically questionable.

In addition, resistance mechanisms may also come with a low cost, none at all or even confer a fitness benefit (Luo et al., 2005; Ramadhan, 2005; Criswell et al., 2006; Kunz et al., 2012). However, in these cases it should be considered carefully, whether appropriate conditions were tested.

14

Can resistance be reversed? In an environment free of antibiotic selection pressure, fitness costs present growth disadvantages for resistant mutants and they are typically outcom-peted by non-resistant strains. However, subsequent compensatory mutations are readily selected and can decrease the cost associated with resistance. The resistance phenotype might be lost during the process of acquiring fitness compensating mutations. However, compensatory mutations often arise in different regions of the chromosome and both resistance and compensatory mutation can be present in the evolved strain. The combination of both mu-tations might make reversion of resistance genetically disadvantageous, be-cause the compensatory mutation alone may now confer a fitness cost. Also, several resistance conferring mechanisms can be acquired together, on plas-mids or other transferable elements and are then genetically linked. Given selection pressure for one of the linked resistance genes and depending on linkage distance, they can be co-selected and resistance will not be lost. Ad-ditionally, it has been demonstrated that the environment is often not antibi-otic free, but that low levels of antibiotics are present (Kümmerer, 2009). Contamination with antibiotics result not just from use of antibiotics as hu-man medication, but also from their widespread veterinary and agricultural application (Aarestrup, 2005; Cabello, 2006). Alarmingly, resistance muta-tions can be selected for at these low levels and already existing mutants are able to outcompete wild type bacteria at very low antibiotic concentrations in vitro (Gullberg et al., 2011). All these factors limit the reversibility of antibiotic resistance and stress the importance of actions to avoid develop-ment of resistance in the first place.

β-lactam antibiotics The first β-lactam, penicillin, was discovered by Alexander Fleming in 1928 (Fleming, 1929). He performed a number of studies but failed to realise its potential as treatment against bacterial infections. In 1940 Chain et al. demonstrated the efficiency of penicillin as a systemic chemotherapeutic agent (Chain et al., 1940), which led to the development of penicillin for clinical use in humans and fuelled the discovery of additional antibiotics. In many cases the introduction of drugs into clinical use was quickly followed by the appearance and spread of bacterial resistance mechanisms (Figure 1). In the case of penicillins, penicillin-degrading enzymes were identified even before the clinical introduction of penicillin G (benzylpenicillin) (Abraham and Chain, 1940).

15

Figure 1. Timeline of key events 80 years after β-lactam discovery. The time of discovery and clinical introduction of β-lactam antibiotics are depicted above the arrow. Events beneath the arrow mark key events in the evolution of β-lactam re-sistance.

β-lactam antibiotics inhibit bacterial cell wall biosynthesis. Cell wall synthe-sis is mediated by transpeptidases, transglycosylases and carboxylases, which build N-acetylglucosamine and N-acetylmuramic acid peptidoglycan chains and cross-link them to form peptidoglycan. Naturally, transpeptidases react with the peptide D-alanyl-D-alanine side chain to form an acyl-enzyme intermediate. The structure of transpeptidases excludes water molecules from the active site to allow reaction of the acyl-enzyme intermediate with the amine group of the neighbouring chain. In the presence of β-lactams, the β-lactam ring is mistaken by the transpeptidases for a yet to be cross-linked peptidoglycan side chain and consequently they bind to it. The transpepti-dases are therefore called penicillin-binding proteins (PBPs). As illustrated in figure 2, the PBPs will form the acyl-enzyme intermediate complex with a β-lactam, but with the exclusion of water and no amine group in the proxim-ity the transpeptidase is trapped in this inactive form. Stabilisation of the peptidoglycan layer and growth of the cell will be significantly reduced and depending on the bound PBP this can lead to cell lysis (Walsh and Walsh, 2003).

1930 1940 1950 1960 1970 1980 1990 2000

penicillin discovery

introd. penicillin G

cephalosporin discovery

discovery monobactams,

clavams, carbapenems

introd. clavulanic acid,

imipenem

introd. 3rd gen. cephalosporins

introd. 4th gen. cephalosporins,

meropenem

chromosomal ơ�ODFWDPDVH

SODVPLG�ERUQH�ơ�ODFWDPDVH��7(0����6+9��

SODVPLG�ERUQH�(6%/��6+9��

SODVPLG�ERUQH�&7;�0�HQ]\PHV

FDUEDSHQHPDVH��,0,����2;$���

PHWDOOR�ơ�ODFWD�PDVH��,03��

introd. 2nd gen. cephalosporins

introd. ertapenem

introd. doripenem

2010

PHWDOOR�ơ�ODFWD�PDVH��1'0��

introd. cephalothin

NO

R1R2

Ser OH

NO

R1R2

N

R1R2

H

H

O

OO

Transpeptidase

Ser H2O

RNO

R1R2

HOH

β-lactamase

+

β-lactam

Transpeptidase (stable acyl-enzyme)

Serine β-lactamase (rapid hydrolysis)

H2Oβ-lactam-enzyme

intermediate complex

S S

Ser

S

S

16

Figure 2. Reaction of β-lactam antibiotics with transpeptidase (penicillin-binding protein) or serine-β-lactamase. The conformation of transpeptidases excludes water from the active site and the enzyme stays inactive and bound to the β-lactam. Water molecules can freely access the active site of serine-β-lactamases and the antibiotic is rapidly hydrolysed.

PBPs are divided into low-molecular mass (LMM) and high-molecular mass (HMM) PBPs. The LMM PBPs in E. coli (PBP4-7, PBP4b, PBP6b, AmpH) are involved in cell separation, peptidoglycan maturation and recycling (Popham and Young, 2003; Sauvage et al., 2008). These enzymes are not essential, but studies have shown that they are needed to maintain uniform cell shape and deletion of at least three LMM enzymes caused significant morphological defects in E. coli cells (Nelson and Young, 2001). There are five HMM PBPs in E. coli. PBP1a and PBP1b are the major peptidoglycan synthetases with transglycosylase and transpeptidase activity. Deletion of one of these enzymes is tolerated but deletion of both is lethal. PBP1b has been shown to interact with PBP3, suggesting a role in localisation or regu-lation of cell division (Bertsche et al., 2006). PBP1c is less well studied. It is similar to PBP1a and 1b in molecular weight but only has transglycosylase activity (Schiffer and Höltje, 1999). It cannot compensate for deletion of both PBP1a and 1b. PBP2 and PBP3 are monofunctional transpeptidases. PBP2 (encoded by mrdA) is essential for cell elongation and inhibition by PBP2-specific β-lactams such as mecillinam leads to the formation of spher-ical cells (Spratt, 1975). PBP3 (encoded by ftsI) is the main protein in the cell division complex and polymerises the septal peptidoglycan. When in-hibited cells grow as filaments (Sauvage et al., 2008). Each PBP has a spe-cific “binding-pattern” to various β-lactam antibiotics; penicillins are bound more strongly by PBP1a and PBP3; cephalosporins and monobactams are bound by PBP3; mecillinam and carbapenems are bound strongest by PBP2 (Curtis et al., 1979; Bush et al., 1987; Davies et al., 2008; Koga et al., 2009). β-lactams are the best studied and most widely used class of antibiotics.

They owe this to their strong bactericidal activity, low toxicity and synergis-tic effect with other classes of antibiotics. The common feature of all β-

17

lactams is a 4-membered amide ring, the β-lactam ring. This ring is fused to other 5- or 6-membered rings or can stand alone, depending on the group of β-lactam. Attributes of newly developed generations of β-lactams reflect a constant struggle to stay ahead of bacterial resistance development. Mem-bers of β-lactams include:

i) Penicillins. The β-lactam ring is fused to a 5-membered sulphur ring system and this group of β-lactams is produced by Penicillium and Aspergil-lus species of fungi. To date there are four generations of penicillins (Figure 3). The first generation is only active against Gram-positive species and following generations were developed against Gram-negative organisms. Increased stability against penicillinases is characteristic for later genera-tions as well as the good activity against Pseudomonas aeruginosa in the newest generation (Walsh and Walsh, 2003).

ii) Cephalosporins. The 4-membered β-lactam ring is fused to a 6-membered sulphur ring system. In the 1940s Giuseppe Brotzu isolated a fungus called Cephalosporidium acremonium that was later found to pro-duce a number of different cephalosporins and further studies by E. T. Abra-ham and others led to the development of cephalosporins for clinical use (Muñiz et al., 2007). Worldwide, cephalosporins are prescribed the most among the β-lactams. Modifications of the C3 and C7 side chains of the third generation cephalosporins prevent binding of β-lactamases. Fourth generation cephalosporins show greater activity against Gram-negative bac-teria, including P. aeruginosa and the fifth generation even has activity against methicillin-resistant Staphylococcus aureus (MRSA) (Walsh and Walsh, 2003).

iii) Monobactams. Their monocyclic structure with the β-lactam ring not fused to another ring presents a departure from the usual 2-ring structure of other β-lactams. Monobactams are not effective against Gram-positive bac-teria but show good antimicrobial activity against Gram-negative bacteria, including P. aeruginosa. Aztreonam is the only commercially available monobactam.

iv) β-lactamase inhibitors. With the emergence of β-lactamases the need for β-lactamase inhibitors grew. These are not antibiotics but substances that target β-lactamases and are administered together with a β-lactams to protect them from degradation. Olivanic acid was discovered and shortly thereafter clavulanic acid in 1976, which became the first clinically available β-lactamase inhibitor (Papp-Wallace et al., 2011).

v) Carbapenems. The first carbapenem was thienamycin from Streptomy-ces cattleya and all other carbapenems are derived from it, starting with imipenem, which became clinically available in 1985. Carbapenems are the focus of this thesis and members of this group of β-lactams are described in more detail in the following chapter (Page 18).

18

Figure 3. Overview of β-lactam antibiotics.

Carbapenems Carbapenems are part of the β-lactam class of antibiotics. They have a 4-membered β-lactam ring fused to a thiazolidinic 5-membered ring (Figure 4) and were discovered in the 1970s. The first carbapenem, thienamycin from S. cattleya, was not pursued for clinical use due to its instability in aqueous solutions. Imipenem was discovered shortly after and is more stable in solu-tion. It needs to be administered with another compound, cilastatin, to pro-tect it from degradation by human renal dehydropeptidase I (DHP-1). Later additions to the group of carbapenems have a 1-β-methyl substituent at C1 (marked in bold in figure 4) making them stable against DHP-1 degradation. The hydroxyethyl group in trans-configuration at C6 of all carbapenems confers stability to serine-β-lactamases (Moellering et al., 1989; Nicoletti et al., 2002; Papp-Wallace et al., 2011). Because of this stability, carbapenems are primarily used for treatment of severe infections caused by otherwise resistant bacteria such as extended spectrum β-lactamase-producers.

NO

Penicillins

Cephalosporins

Extended spectrum,ȕ�ODFWDPDVH�UHVLVWDQWH�J��,PLSHQHP Meropenem�������(UWDSHQHP�������'RULSHQHP

/RZ�DQWLPLFURELDO�DFWLYLW\�6XLFLGH�VXEVWUDWH�IRU�ȕ�ODFWDPDVHVH�J��&ODYXODQLF�DFLG

/LPLWHG�WR�*UDP�QHJDWLYHV�$QWL�SVHXGRPRQDOH�J��$]WUHRQDP

1DUURZ�VSHFWUXPH�J��3HQLFLOOLQ�*

$QWL�VWDSK\ORFRFFDO�3HQLFLOOLQDVH�UHVLVWDQWH�J��0HWKLFLOOLQ�������2[DFLOOLQ

%URDG�VSHFWUXP�3HQLFLOOLQDVH�VHQVLWLYHH�J��$PSLFLOOLQ�������$PR[DFLOOLQ

1stExtended spectrum,$QWL�SVHXGRPRQDOH�J��3LSHUDFLOOLQ�������&DUEHQLFLOOLQ

4th2nd 3rd

NO

SR

1DUURZ�VSHFWUXP�ȕ�ODFWDPDVH�VHQVLWLYHH�J��&HSKDORWKLQ

([SDQGHG�VSHFWUXP�,PSURYHG�ȕ�ODFWDPDVH�VWDELOLW\H�J��&HIXUR[LPH�

%URDG�VSHFWUXP�ȕ�ODFWDPDVH�UHVLVWDQWH�J��&HIRWD[LPH

1st 5thExtended spectrum,$QWL�SVHXGRPRQDO�H�J��&HIHSLPH��

Extended spectrum,$QWL�SVHXGRPRQDO�$QWL�056$H�J��&HIWDUROLQH�������&HIWRELSUROH

4th2nd 3rd

NO

R1 S

R2

R1R2

NO

R

SO3H NO

Carbapenems Monobactams ȕ�ODFWDPDVH�LQKLELWRUV

19

Figure 4. Chemical structure of carbapenems. (a) Imipenem, (b) Meropenem, (c) Ertapenem, (d) Doripenem.

All carbapenems bind strongly to penicillin-binding protein 2 (PBP2) in E. coli and show varying affinity to PBP3 and other PBPs. Imipenem binds most strongly to PBP2, followed by PBP1a and 1b and has a weak affinity for PBP3. Ertapenem binds equally strongly to PBP2 and PBP3, whereas meropenem binds strongly to PBP2 but somewhat weaker to PBP3. Both drugs bind well to PBP1a and 1b. Doripenem binds strongly to PBP2, but shows only weak binding to PBP1a, 1b and PBP3 (Yang et al., 1995; Kohler et al., 1999; Davies et al., 2008; Koga et al., 2009). The high affinity of carbapenems to these essential PBPs is what facilitates their rapid bacteri-cidal effect and very broad spectrum, as they are effective against Gram-positives, Gram-negatives and both aerobes and anaerobes of these groups. The interaction with PBPs takes place on the cell surface of Gram-positive bacteria and in the periplasmic space of Gram-negative bacteria. In the latter, carbapenems must cross the outer membrane through water filled channels, called porins, to access the PBPs. Efficient penetration though porins is due to their zwitterionic properties (Kattan et al., 2008) and make carbapenems very potent. Once in contact with the PBPs, carbapenems will trap these enzymes in an inactive state and the peptidoglycan will weaken.

N

OH

O

S

CH3

12

3456 7

NH

O

N

N

OH

OHO

O

S

N

O

N

H

H O

OH

OHO

N

OH

O

S12

3456 7

OHO

NH

H

NH

CH3

(a)

(c)

(b)

N

OH

OHO

O

S

CH3

NH

NH

NHS 2

O O(d)

12

3456 7

12

3456 7

20

β-lactam resistance

The use of β-lactams led to clinical resistance developing as early as the 1930s (Abraham and Chain, 1940; Hedge and Spratt, 1985; Näsvall et al., 2012; Yamachika et al., 2013). Today, there are four described mechanisms by which bacteria can become resistant to β-lactams:

i) The outer membrane of Gram-negative bacteria serves as a natural bar-rier for β-lactams, restricting the influx of antibiotics more than the Gram-positive cell wall (Walsh and Walsh, 2003). Lowering the number of outer membrane porins can further decrease the membrane permeability. Outer membrane porins are the entryway of β-lactams into the cell and the number of porins ultimately influences the antibiotic susceptibility of the cell.

ii) Efflux pumps can recognize β-lactams in the periplasm and export them back to the cell surface. This can lower the drug concentrations within the cell and confer resistance. Decreased membrane permeability and in-creased efflux often act synergistically to significantly increase resistance (Malléa et al., 1998; Källman et al., 2009).

iii) Overproduction and modification of target PBPs or acquisition of novel PBPs may render bacteria insensitive to β-lactams. Examples are the overexpression of PBP5 in Enterococcus hirae (Zapun et al., 2008) or ac-quisition of PBP2a through the staphylococcal cassette chromosome mec in Staphylococcus aureus. No E. coli clinical isolates have been found to ex-press modified PBPs but in vitro studies show that mutant PBP2 and PBP3 can be selected during β-lactam exposure (Hedge and Spratt, 1984; Yama-chika et al., 2013).

iv) Hydrolysis of β-lactam antibiotics by β-lactamases and car-bapenemases is the main cause for resistance development worldwide.

Below I address the role of outer membrane porins, β-lactamases and car-bapenemases in β-lactam resistance development in more detail, as these factors are of special importance for this thesis.

Role of outer membrane proteins in resistance Permeability of the outer membrane significantly affects the entry of β-lactams into the bacterial cell (Nikaido, 2003; Lartigue et al., 2007; Pagès et al., 2008). β-lactams enter through the major non-specific outer membrane porin proteins, called OmpC and OmpF in E. coli, that form water filled channels for the transport of nutrients. Porins are excessively regulated to enable survival under widely different condition and environmental factors such as osmolarity, temperature, pH, oxidative stress, acetyl phosphate, cer-tain toxins, antibiotics as well as the bacterial growth phase influence porin expression (Pratt et al., 1996).

21

The main regulator of ompC and ompF in E. coli is the two-component regulatory system EnvZ-OmpR (Figure 5). EnvZ senses external osmolarity, is phosphorylated at residue His243 and passes this information on to OmpR by phosphorylation of Asp55 (Forst, Delgado, and Inouye, 1989b; Forst, Delgado, and Inouye, 1989a). The level of phosphorylated OmpR (OmpR-P) is controlled by EnvZ phosphatase activity. At low external osmolarity only a small amount of OmpR is phosphorylated, whereas at high osmolarity the phosphatase activity of EnvZ decreases and additional OmpR-P becomes present in the cytoplasm (Aiba and Mizuno, 1990). Phosphorylation of OmpR improves its DNA-binding activity and results in tandem-binding to regions upstream of the ompC and ompF promoters (Harlocker et al., 1995). These regions are three 20-basepair units located at -100 to -38bp from ompC (C1, C2 and C3) and -100 to -39bp from ompF (F1, F2 and F3) with an additional fourth ompF-binding site (F4) at -380 to -361bp. There is a distinct hierarchy in binding of these sites where the F1 and C1 sites are bound first. This increases the OmpR-P affinity to subsequent sites. The F2 site is bound next and at this point OmpF can be expressed, but not OmpC. Binding of the F3 site happens almost simultaneously with the F2 site, but higher concentrations of OmpR-P are required for binding of C2 and C3. At increased levels C2 and C3 are bound and OmpC is expressed. The F4 site is also bound at these high OmpR-P concentrations but this binding instead blocks OmpF expression by a DNA loop formation. This complex binding pattern (F1, C1 > F2, F3 > C2 > C3, F4) regulates coupled expression of OmpC and OmpF.

Expression of ompF is also regulated by the sRNA MicF. MicF binds to the translation-initiation region of the ompF mRNA and represses expres-sion. Expression of micF in turn is regulated by a number of transcriptional regulators in response to osmolarity, temperature, acids or antibiotics, in-cluding OmpR and MarA (Pratt et al., 1996; la Cruz and Calva, 2010).

The pores of OmpC and OmpF have different diameters (Nikaido and Rosenberg, 1983) and the observed regulation mainly serves to adapt to the osmolarity of the surrounding medium and to evade toxic molecules, such as bile salts or antibiotics (Nikaido, 2003). The smaller pore of OmpC is main-ly expressed in high osmolarity environments where the presence of toxins is more likely. Low osmolarity environments may contain less toxic com-pounds and here the larger pore of OmpF is advantageous to allow more general influx of scarce nutrients.

22

Figure 5. Main regulators of ompC and ompF gene expression.

Porins are the entryway for β-lactams and other hydrophilic antibiotics, e.g. tetracycline, chloramphenicol and fluoroquinolones into the bacterial cell (Nikaido, 2003). Any changes that affect this route can lead to increased tolerance to these drugs. Loss-of-function mutations in the porin sequence (Doumith et al., 2009) or mutations in the porin regulatory genes (Paper I) will decrease the number of channels. Cells can also replace the large gen-eral porins by channels with narrower pores or substrate specific channels (Doménech-Sánchez et al., 1999; García-Sureda et al., 2011). It is argued that porin-deficiency alone leads to only small increases in drug tolerance and that synergy with efflux complexes (Nikaido, 2001) or inactivation through β-lactamases (Nikaido, 1989) are necessary for high level re-sistance.

Little is known about regulation of OmpK35 and OmpK36, the main pore proteins in Klebsiella pneumoniae (homologues to OmpF and OmpC in E. coli), but homologues of EnvZ and OmpR are present. In K. pneumoniae porin loss in combination with expression of β-lactamases has frequently been associated with resistance to β-lactams and carbapenems (Doménech-Sánchez et al., 1999; Woodford et al., 2007; Doumith et al., 2009).

Inactivation by β-lactamases One of the major resistance mechanisms conferring β-lactam resistance is hydrolysis of the β-lactam ring by β-lactamases (Figure 2). β-lactamases may have evolved from penicillin-binding proteins (PBPs) (Massova and Mobashery, 1998). Both kind of enzymes recognise the β-lactam ring and form the acyl-enzyme intermediate, but only β-lactamases allow water mol-ecules to enter their active site. This allows rapid hydrolysis of the β-lactam and the β-lactamase is free to hydrolyse more β-lactam molecules. In con-trast, PBPs exclude water from their active site and remain inactively bound to the β-lactam. Penicillinase was the first β-lactamase discovered (Abraham and Chain, 1940) and in 1983 was followed by the discovery of the first plasmid-encoded extended spectrum β-lactamase (ESBL) in Gram-negative

23

bacteria (Knothe et al., 1983). There are two main types of β-lactamases, serine-β-lactamases with a serine residue at the active site and metallo β-lactamases, which employ zinc ions as a co-factor. β-lactamase nomencla-ture is complicated and the names can refer to the organism from which the enzyme was first isolated (e.g., KLUA from Kluyvera ascorbata), the place of isolation (e.g., NDM from New Delhi metallo-β-lactamase) or the name of a patient (e.g., TEM from patient Temoneira). A collection of more exam-ples has been made by Jacoby (Jacoby, 2006). The enzymes are currently grouped according to two main classification schemes. Ambler’s molecular classification is based on amino acid sequence (Ambler, 1980). Here Class A, C and D enzymes use serine in their active site and Class B enzymes requires zinc ions to hydrolyse the β-lactam ring. The functional classifica-tion scheme of Bush assigns β-lactamases into groups 1-4 and uses a more phenotypical approach, taking substrate and inhibitor profiles into account (Bush, 1989). Functional group 4 has been omitted in a recent update of this classification because these enzymes were assigned to other groups as more information became available (Bush and Jacoby, 2010). β-lactamases can hydrolyse all penicillins and early cephalosporins.

ESBLs have evolved in response to newer β-lactam antibiotics and the name reflects the expanded substrate spectrum of these enzymes. In addition to the substrate of broad-spectrum β-lactamases, these enzymes can hydrolyse third and forth generation cephalosporins and monobactams. They are most often found in K. pneumoniae and E. coli, but also in Acinetobacter bau-mannii, P. aeruginosa, Salmonella typhimurium, Enterobacter sp., Proteus sp., Serratia sp. and others (Jacoby and Munoz-Price, 2005). Giske et al. (Giske et al., 2009) recently proposed a new nomenclature for ESBLs, in an attempt to increase comprehension in the clinic and for the broader public. Here ESBLs of the molecular Class A (ESBLA) and other Classes (miscella-neous ESBLs, ESBLM) are separated from ESBLs with activity against car-bapenems (ESBLCARBA).

Described below are β-lactamases important for the work performed in this thesis.

TEM-enzymes The β-lactamase TEM-1 was identified in 1963 and was named after patient Temoneira from which the first sample was obtained (Salverda et al., 2010). It is a member of the molecular Class A and functional group 2b. TEM-1 can hydrolyse penicillins and early cephalosporins and used to be responsible for 90% of the ampicillin resistance in E. coli (Bradford, 2001). Now CTX-M-type enzymes are more prominent and widespread (Naseer and Sundsfjord, 2011). All TEM-enzymes are plasmid borne derivatives of TEM-1, and there are 217 TEM-enzymes described (http://www.lahey.org/Studies/). TEM-3 (a TEM-2 derivative) was reported in 1988 to be the first TEM-enzyme dis-playing an ESBL phenotype (Sougakoff et al., 1988). TEM-ESBLs are ac-

24

tive against penicillins and broad-spectrum cephalosporins with higher activ-ity against ceftazidime than cefotaxime (Bush’s group 2be, Giske’s ESBLA class) and some derivatives are inhibitor resistant ESBLs (Bush’s group 2br). Other derivatives are known to have a substrate spectrum similar to TEM-1 but are not inhibited by clavulanic acid. The mobilisation of TEM-genes is associated with Tn3-like structures (Salverda et al., 2010). The rap-id worldwide spread of TEM-enzymes during the 1960s and 1970s world-wide is only comparable to recent global advances of CTX-M-type enzymes.

OXA-enzymes The OXA-1 β-lactamase belongs to the molecular Class D and functional group 2d, it utilizes serine in its active site and is poorly inhibited by clavu-lanic acid. It is active against ampicillin and cephalothin and was named after its high hydrolytic activity for oxacillin and cloxacillin (oxacillinase). There are currently 404 OXA-derivatives (http://www.lahey.org/Studies/) but members of the group are extremely diverse, sharing only 20% sequence similarity (Bradford, 2001). It has been shown that OXA-genes originate from P. aeruginosa (Cantón et al., 2008) and have mobilised to plasmids at least three times, millions of years ago (Barlow and Hall, 2002). They can also be found in E. coli, S. typhimurium, K. pneumoniae and Proteus mirabi-lis (Naas and Nordmann, 1999). OXA-ESBLs are derived from OXA-10 and OXA-2 and several enzymes have been reported to be sensitive to clavulanic acid (Rasmussen et al., 1994; Philippon et al., 1997).

CTX-M-enzymes CTX-M-enzymes were named for their preferred activity against cefotaxime over that of other oxyimino-β-lactams (Gniadkowski, 2008) and their place of isolation, Munich (German), in the late 1980s (Matsumoto et al., 1988; Bauernfeind et al., 1990). CTX-M-enzymes belong to the Ambler Class A, and Bush’s functional group 2be, and ESBLA according to Giske’s classifi-cation. They are mainly found in K. pneumoniae and E. coli but originate from genomes of environmental Kluyvera sp. (Bonnet, 2004). Currently there are 151 CTX-M derivatives (http://www.lahey.org/Studies/) clustered into 5 groups based on amino acid homology: CTX-M cluster 1A, 1B, 2, 8, 9 and 25 (Bonnet, 2004; Naseer and Sundsfjord, 2011), with the CTX-M-1 cluster including some of the major clinical β-lactamases. There are approx-imately 226 nucleotide mutations between the different groups and their most recent common ancestor (Barlow et al., 2008). Their efficient world-wide spread is believed to be caused by an unusually frequent mobilization, eight times, compared with other class A β-lactamases (Barlow et al., 2008) and it is believed that this mobilisation occurred recently, because some CTX-M-gene sequences on plasmids still exactly match the sequence of some Kluyvera chromosomal genes. The rate of mobilisation and spread of CTX-M enzymes has led to the present CTX-M-pandemic (Cantón and

25

Coque, 2006). CTX-M-15 is a member of the CTX-M-1A (Naseer and Sundsfjord, 2011) cluster originating from the chromosome of Kluyvera ascorbata (Rodríguez et al., 2004). Specific point mutations allow CTX-M-15 to hydrolyse ceftazidime at high levels (Bonnet, 2004). The first plasmid described to encode CTX-M-15 was pC15-1a (Boyd et al., 2004). Together with other enzymes from the CTX-M-1, CTX-M-2 and CTX-M-9 clusters, CTX-M-15 has been captured by ISEcp1-like insertion sequences (Bonnet, 2004). These IS elements can mobilize adjacent DNA by recognizing a vari-ety of DNA sequences. It is proposed that the promoter of ISEcp1-like ele-ments provides higher expression levels for CTX-M genes in Enterobacteri-aceae than in Kluyvera species (Poirel et al., 2003).

Inactivation by Carbapenemases The strong selective pressure posed on bacteria by the use of carbapenems has facilitated the emergence and spread of carbapenemases in β-lactamase classes A, B and D in clinical settings (Nordmann et al., 2011; Cantón, Akóva, et al., 2012). Carbapenemases are defined as β-lactamases with ac-tivity against most β-lactams and measurable activity against carbapenems (Livermore, 1992; Queenan and Bush, 2007) and can naturally be encoded on the chromosome, acquired to the chromosome, or be plasmid encoded. Just like other β-lactamases, these enzymes can be grouped according to their amino acid sequence in the Ambler classification, depending on sub-strate and inhibition profile in Bush’s functional groups and in Giske’s new-ly proposed simplified scheme. Currently, carbapenemases with the greatest clinical relevance in Europe are KPC-enzymes, OXA-carbapenemases and enzymes of the metallo-β-lactamase group. Below, I shortly describe these carbapenemases as their clinical importance makes them interesting for this thesis.

Serine-carbapenemases KPC-derivatives (K. pneumoniae carbapenemase) belong to Ambler class

A, Bush’s group 2f and Giske’s ESBLCARBA-A and are mainly encoded on plasmids, which has facilitated their worldwide spread. Currently there are 18 KPC-derivatives (http://www.lahey.org/Studies/).

OXA-carbapenemases are clustered into nine subgroups (Queenan and Bush, 2007; Walther-Rasmussen and Høiby, 2007), which share 40 to 70% sequence similarity and are only remotely related to Class D β-lactamases. Many OXA-type carbapenemases hydrolyse oxacillin at much lower rates than Class D oxacillinases and have only low hydrolytic activities against imipenem and meropenem. Bacteria expressing these enzymes often require additional mechanisms to provide clinical resistance levels. OXA-23, OXA-40, OXA-48 and OXA-58 are the only plasmid borne OXA-type car-

26

bapenemases. It is possible that the mobile ancestors of OXA-carbapenemases disappeared long ago.

Metallo-β-lactamases VIM-enzymes (Verona integron-encoded metallo-β-lactamase, 41 deriva-

tives) and IMP-enzymes (active on imipenem, 48 derivatives) were initially found in P. aeruginosa and are associated with class 1 integrons (Nordmann and Poirel, 2002). The VIM and IMP-type enzymes have the largest num-bers of enzyme derivatives among the metallo-β-lactamases and outbreaks with bacteria producing these have been reported worldwide. However, the greatest clinical impact has been made by NDM-enzymes (New Delhi metal-lo-β-lactamase, 11 derivatives) (Yong et al., 2009). NDM-1 was first report-ed in 2009, since then it has spread to most parts of the world and is most often found in K. pneumoniae and E. coli. It is associated with diverse plas-mids and infections were initially connected to the Indian subcontinent, where NDM-enzymes are very frequent in the human population and can even be found in the environment (Nordmann et al., 2011).

ESBL plasmid pUUH239.2 Plasmids are circular DNA molecules that replicate independently of the bacterial chromosome. They frequently encode proteins for their own trans-fer, but also virulence factors, novel metabolic proteins or proteins for anti-biotic resistance. They are not essential for bacterial survival, but can be advantageous under certain conditions. The 220 kilo base pairs (kb) plasmid pUUH239.2 was isolated from the K. pneumoniae clone that caused the first major ESBL outbreak in Scandinavia (Lytsy et al., 2008) in 2005 at Uppsala University Hospital, Uppsala, Sweden. It has been fully sequenced and ana-lysed (Sandegren et al., 2011).

27

Figure 6. pUUH239.2. (a) Complete view of pUUH239.2 and alignment to pKPN3, pC15-1a and pEK499. (b) Simplified pUUH239.2 resistance cassette with resistance genes depicted in blue and IS26 elements in yellow.

The plasmid is a recombinant formed from the plasmid backbone of pKPN3 from K. pneumoniae and a multidrug resistance cassette commonly found in IncFII plasmids, like pEK499 (Woodford et al., 2009), and pC15-1a (Boyd et al., 2004) from E. coli (Figure 6a). The pKPN3 backbone encodes pro-teins required for plasmid maintenance, transfer, and resistance to arsenic, copper and silver. The size of the resistance cassette is 41 kb and it encodes resistance against macrolides [mph(A), mrx, mphR(A)], chromate (chrA), sulfonamides (sul1), aminoglycosides (aadA2), aminoglyco-sides/fluoroquinolones [aac(6’)-1b-cr], trimethoprim (dhfrXII), β-lactams (blaTEM-1, blaOXA-1), tetracycline [tet(A), tetR], and the ESBL gene blaCTX-M-15. It is flanked by two IS26 elements with four other IS26 elements spread out across the cassette, which have been involved in rearrangements of this re-gion. The presence of six identical IS26 elements in the resistance cassette provide regions of homology for GDA. This makes pUUH239.2 a good model to study the evolution of resistance genes following gene amplifica-

200 kb

180 kb

160 kb

140 kb

120 kb 100 kb

80 kb

60 kb

40 kb

20 kb

0 kb

Plasmid transfer region

Copper

resistanceRes

istance cassette

Arsenicresistance

resistanceSilver

pC15-1a

pEK499

pKPN3

pecM

tetR

tetA

IS26

(3)

blaOXA-1

IS26

(1)

IS26

(5)

IS26

(4)

IS26

(6)

IS26

(2)

blaTEM-1

blaCTX-M-15

pem

K

tnpA

catB

3

tnpR

aac(

6)-1

b-cr

tnpA

tnpA

tnpA

tnpA

ISEcp1B

tnpR

tnpA

tnpR

tnpM

aadA

2dh

fRin

t

tnpA

mrx

mph

R

chrA

sul1

mph

(A)

tnpA

orf9

3

orf9

1

TDF(Ƌ�

orf9

5

1600 3600 5600 15600136001160096007600

2360021600

1960017600

296002760025600 356003360031600 37600 39600 41600

a)

b)

28

tion. Studies of the fitness cost identified a 4% reduction in the growth rate of E. coli and K. pneumoniae cells carrying this ESBL-plasmid plasmid.

Gene duplication and amplification Gene duplication and amplification (GDA) is a genetic phenomenon ob-served in all three kingdoms of life (Devonshire and Field, 1991; Romero and Palacios, 1997; Wong et al., 2007; Andersson and Hughes, 2009; Sandegren and Andersson, 2009). It provides genetic variation in popula-tions, short-term solutions to environmental challenges, and a source of ge-nome complexity. Together with de novo emergence of new genes from previously non-coding DNA sequences (Kaessmann, 2010; Tautz and Domazet-Lošo, 2011), GDA is believed to be the major contributor to gen-erating genetic diversity in all types of organisms. In medicine GDA has been shown to cause a number of human diseases (Lee and Lupski, 2006; Lupski, 2007; McCarroll and Altshuler, 2007). Understand the mechanisms and dynamics of GDA is of fundamental evolutionary interest.

Dynamics Gene duplication and amplifications (GDAs) are very frequent in a growing bacterial population, >10-2∼10-5 per cell per generation, depending on the gene and region of the genome (Anderson and Roth, 1981). Any region on the chromosome can be affected and the duplicated region can range from a few base pairs to several Mega base pairs (Mb) (Straus and D'Ari Straus, 1976; Anderson and Roth, 1981; Nilsson et al., 2006; Kugelberg et al., 2006; Andersson and Hughes, 2009; Sandegren and Andersson, 2009). From this data it has been estimated that at any time 10% of all cells in an unse-lected growing population have a duplication somewhere on the chromo-some (Roth et al., 1996).

The steady-state of GDA is many orders of magnitude higher than the base substitution rates determined in several locations in the bacterial chro-mosome, 10-11∼10-10 per cell per generation (Hudson et al., 2002), or for instance mutation rates for spontaneous carbapenem resistance, 10-10∼10-7 per cell per generation (Paper I). This suggests that subpopulations with GDAs are likely enriched in response to selective pressures rather than rare mutation, assuming that GDAs can alleviate the selective pressure. Duplicat-ed genes can confer resistance during antibiotic treatment of patients (Mush-er et al., 2002; Hammond et al., 2005; Bertini et al., 2007; Brochet et al., 2008; Sandegren and Andersson, 2009; Huang et al., 2013) and with the above-mentioned steady state frequencies of GDA it is very likely that am-plification of resistance genes are present in a population. Standard growth

29

conditions in clinical laboratories are usually non-selective and GDA can be lost from the population after routine clinical isolation resulting in the un-derestimation of GDA as clinically relevant resistance mechanism.

Studies need to be specifically designed to detect GDA and some tech-niques that are commonly used are transduction assays, pulsed-field gel elec-trophoresis (PFGE), next generation sequencing and quantitative real-time polymerase chain reaction (qRT-PCR) (Andersson and Hughes, 2009; Elliott et al., 2013). In a transduction assay, an auxotrophic phenotype is caused by insertion of an antibiotic resistance cassette and this auxotrophic mutation is subsequently moved by transduction into a recipient that is suspected to carry a duplication of this gene. The transferred auxotrophy will not be de-tectable if the inactivated gene is indeed present in at least two copies, be-cause only one copy is inactivated by the antibiotic marker. This technique is only suitable for detection of large duplicated regions. PFGE can upon di-gestion of chromosomal or plasmid DNA show changes in the banding pat-tern as compared to non-amplified DNA. Amplifications between digestion sites will cause larger DNA fragments and additional fragments will appear if the amplification covers a restriction site. Next generation sequencing detects amplification joint points as novel junction sequences (Sun et al., 2012). It also allows a rough estimate of the number of additional gene cop-ies, because the number of sequence reads covering the region will increase proportionally to the number of gene copies. qRT-PCR is useful if the dupli-cated sequence is known and then it is a valuable technique to verify the presence of the amplification in particular isolates and make precise meas-urements of the number of gene copies.

The loss rates of GDA have been determined experimentally to be as high as 0.15 per cell per generation (Roth et al., 1996; Pettersson et al., 2009) and the fitness costs of duplications in absence of selection can range from indis-tinguishable from the unduplicated wild type to as high as 20% growth rate reduction (Pettersson et al., 2009). The cost does not depend on the size of the duplication and therefore it is unlikely that it is caused by carriage and replication of additional DNA. The cost of replicating the complete genome is 2% of the energy budget of the cell (Lane and Martin, 2010) and does not explain costs larger than that for the replication of amplified DNA. The cost might be due to expression of duplicated genes and production of protein by ribosomes and the use of amino acids. In addition to that the translated pro-teins can take part in energy-requiring metabolic reactions (Koskiniemi et al., 2012) and this can lead to decreased growth rates. Also increased amounts of RNA or protein can affect regulatory processes and lead to im-balanced regulation for instance in a positive feedback loop (Mileyko et al., 2008).

30

Mechanism of formation and loss of GDA Two mechanisms have been described for the initial duplication formation depending on whether RecA is required or not. Non-equal homologous re-combination requires RecA and long directly oriented sequence repeats, e.g. ribosomal RNA operons (Anderson and Roth, 1981), rearrangement hot spot (rhs) sequences (Lin et al., 1984), insertion sequences (IS) (Nichols and Guay, 1989; Nicoloff et al., 2007) or transposable elements. This leads to the duplication of the region between the repeat sequences and a novel joint sequence. The importance of RecA has been demonstrated recently (Reams et al., 2010), but a recA mutant, essentially unable to perform homologous recombination, only experiences ten-fold lower duplication rates. This indi-cates that a substantial amount of duplication events occur in a RecA-independent manner and that other mechanisms must be capable of duplica-tion formation. These mechanisms include strand slippage during DNA rep-lication or repair, pairing of single stranded regions at the replication fork and erroneous ligation of DNA ends by DNA gyrase (Trinh and Sinden, 1993; Lovett et al., 1993; Ikeda et al., 2004). This can result in duplication of genomic regions by recombination between very short repeats or without repetitive sequences.

The initial duplication formation event is the rate-limiting step for GDA and further amplification may be achieved by selection for increased gene dosage. This mechanism involves RecA-mediated homologous recombina-tion between the large repeats provided by the initial duplication. It is also possible to form amplifications without the initial duplication. Here a rolling cycle replication-type mechanism could generate a large tandem array in a single generation (Petit et al., 1992).

With regard to the loss of GDA, RecA-mediated homologous recombina-tion is the main mechanism and many studies showed that recA mutations can stabilize GDA (Hill et al., 1969; Lin et al., 1984; Haack and Roth, 1995; Andersson et al., 1998; Reams et al., 2010). In some plasmid systems how-ever, deletion between repeat sequences occurred through RecA-independent mechanisms (Matfield et al., 1985; Lovett et al., 1993).

Evolution of new genes by duplication-divergence Gene duplications are central to many theories of evolution of new gene functions. In the model proposed by Susumu Ohno (Ohno, 1970), after du-plication, the second gene copy is released from selection pressure to per-form the initial function and free to accumulate beneficial mutations. It may eventually develop a new function, while the other copy retains the original function. This model became known as the neofunctionalization model (Force et al., 1999). One problem is that the “free copy” is subject to random drift, gene conversion or may acquire deleterious mutations that would turn

31

it into a pseudogene rather than acquire a new function (Lynch, 2000). An-other problem is that gene duplications often confer fitness costs (as men-tioned above) and are therefore counterselected. Thus, it is unlikely that the second copy will remain in the population in high frequencies and for suffi-ciently long to acquire rare beneficial mutations. Two mechanisms have been suggested that would help retaining the second copy. After duplication, subfunctions of the gene may be divided onto both gene copies by a series of inactivating mutations (sub-functionalization model) (Force et al., 1999; Lynch and Force, 2000). Assuming that both subfunctions are important for cell growth, both gene copies are stabilized. Secondly, GDA has been shown in response to a number of limitations, such as starvation, exposure to anti-biotics, or deleterious mutations by supplying a higher dosage of the limiting gene (Normark et al., 1977; Neuberger and Hartley, 1981; Sonti and Roth, 1989; Nichols and Guay, 1989; Hammond et al., 2005; Nilsson et al., 2006; Nicoloff et al., 2007; Paulander et al., 2007; Sun et al., 2009; Lind et al., 2010; Paulander et al., 2010; Pränting and Andersson, 2011; Nicoloff and Andersson, 2013). GDA can be a rapid, short-term solution for environmen-tal challenges or facilitate the evolution of new gene functions. If the solu-tion is a beneficial mutation within the amplified region then GDA increases the target size for beneficial mutations. GDAs that improve growth allow the population to expand and increase the likelihood to acquire beneficial muta-tions, even outside the amplified region.

In the recently proposed innovation-amplification-divergence (IAD) model (Bergthorsson et al., 2007) selection acts at every step of the process to stabilize the duplicated copy. i) A minor secondary protein function, en-coded by the gene, is present before the duplication event (innovation). This minor function proves beneficial under new environmental circumstances (e.g., presence of antibiotic or availability of novel nutrient). ii) Selection will promote any increase in the amount of the trace activity and duplication and subsequent amplification will be enriched in the population (amplifica-tion). The amplified state will be maintained and adjusted (as far as the po-tential fitness cost allows) to meet the environmental challenge. For instance an increase in the antibiotic concentration may further increase the number of gene copies. iii) Extra gene copies increase the mutational target and the probability to acquire beneficial mutations. Selection will favour beneficial mutations as they may improve the trace activity (divergence). Improved gene copies can undergo further amplification cycles, while less beneficial copies are lost from the gene array. As more proficient copies emerge, selec-tion of GDA is relaxed on remaining copies and the gene array may shorten until one copy with improved function remains. If this improved copy has lost its original function , a copy of the original gene must remain and the cell is left with two specialist genes, encoding mainly their selected function. Other outcomes are possible, where a very proficient generalist replaces the

32

original copy to perform both functions, or combinations of generalists with improved genes encoding either the old or new function (Näsvall et al., 2012).

In summary, GDAs are likely to be the initial response to many selective pressures due to their high frequency in unselected bacterial populations. The intrinsic instability and potential cost however will cause GDAs to be lost rapidly. Selection can act to maintain the amplified state and allow bene-ficial mutations to accumulate until a new gene function can emerge.

Escherichia coli and Klebsiella pneumoniae as pathogens E. coli and K. pneumoniae are Gram-negative rod-shaped bacteria of the family of Enterobacteriaceae. This family is of clinical importance harbour-ing many other potential pathogenic genera such as Citrobacter sp., Entero-bacter sp., Proteus sp., Pseudomonas sp., Salmonella sp., Shigella sp., and Yersinia sp.. E. coli and K. pneumoniae are part of the normal intestinal flora in animals and humans and are harmless most of the time. They mainly cause infections in immuno-compromised patients (acting as opportunistic pathogens). Otherwise healthy individuals can be infected by specific strains that have acquired virulence factors or if the bacteria have breached the in-nate immunity’s defences through catheters or ventilators. Both species are the main cause of nosocomial Gram-negative bacteraemia and most often cause urinary tract infections. Some E. coli variants cause severe and even fatal diarrheal diseases. In this study we utilised E. coli MG1655. Due to its fast growth, large population sizes and fully sequenced genome, E. coli is one of the main workhorses of bacterial genetics (Blattner et al., 1997). K. pneumoniae is the medically most relevant species of the genus Klebsiel-la. It causes community acquired pneumonia and most often urinary tract infections. Studies of for instance resistance development are not as straight-forward as in E. coli because genetic techniques are limited. Today K. pneumoniae producing ESBLs or carbapenemases are more frequently asso-ciated with large nosocomial outbreaks than E. coli.

33

Present Investigations

ESBL-plasmid influences evolution of carbapenem resistance With the increasing spread of ESBL-producing Enterobacteriaceae world-wide we are increasingly relying on carbapenems for treatment. Therefore, it is of great importance to understand the evolution of carbapenem resistance. To date, decreased susceptibility to carbapenems has mainly been observed in K. pneumoniae and Enterobacter sp. Furthermore, while still rare, car-bapenem-resistance in clinical isolates of E. coli is increasing (Poirel et al., 2004; Hong et al., 2005; Lartigue et al., 2007; Oteo et al., 2008; Falagas and Karageorgopoulos, 2009)

A frequently reported carbapenem resistance mechanism in K. pneumoni-ae is the combination of ESBL-production and loss of porin expression (OmpK35 and OmpK36). To test this combination in E. coli we constructed porin deficient E. coli MG1655 with and without the ESBL-plasmid pUUH239.2 and deletions of the porin genes ompC, ompF and their main regulator ompR using lambda-red recombineering. In paper I we demon-strate that loss of only one porin alone is not sufficient to decrease car-bapenem susceptibility, but loss of both porins or loss-of-function mutations in the main regulators of porin expression are necessary. Porin deficiency and presence of pUUH239.2 in combination lead to ertapenem resistance. In contrast meropenem MICs increased but never reached clinical breakpoints in this study. We could show that the ESBL enzyme CTX-M-15 contributed to reduce carbapenem susceptibility when over expressed, but this was also the case for β-lactamases TEM-1 and OXA-1 (encoded on pUUH239.2). To learn more about the evolution of carbapenem resistance, a stepwise selec-tion for meropenem and ertapenem resistance was performed. This resulted in different mutation spectra but all mutants were affected in porin expres-sion, suggesting differences in the mode of entry of these two carbapenems. Comparison of mutants with regard to pUUH239.2 showed that presence of this ESBL-plasmid increased resistance levels and broadened the mutation spectrum. Cells harbouring pUUH239.2 also survived longer on plates sup-plemented with carbapenems and this led to an apparent increase in mutation rate, most likely an effect of the β-lactamases encoded by this plasmid. Us-ing qRT PCR we discovered that duplication and amplification of the β-

34

lactamase genes on pUUH239.2 were frequent and further contributed to resistance. As for why to date these mutants appear rarely in clinical settings and do not seem to spread, we found that they suffer a fitness cost of 20%, which likely leads to reduced survival and virulence in the host.

Ertapenem resistance due to pUUH239.2 in a pharmacokinetic model Ertapenem is the only carbapenem recommended for administration once every 24 hours making it an appealing choice as patient do not need to stay in the hospital during the course of treatment. We investigated whether clin-ically relevant concentrations of ertapenem can select for resistant mutants adhering to one dose every 24 hours. An in vitro kinetic model developed by Löwdin et al. (Löwdin et al., 1996) was used in paper II to simulate free ertapenem drug concentrations achieved in serum during treatment. In this model bacteria are cultured in a flask allowing dilution of the culture medi-um without loss of bacteria. The initial ertapenem dose of 11 mg/L (corre-sponding to 1g daily for patients) was added and the antibiotic was eliminat-ed at a constant rate by addition of drug-free medium. An additional dose was added after 24 hours. The bacterial populations were monitored during the experiment and examined after 48 hours for susceptibility to car-bapenems and mutations of ompC and ompF and their main regulators. In the presence of the ESBL-plasmid pUUH239.2, concentrations mimicking human pharmacokinetics led to selection of porin-deficient mutants in 55% of the experiments, whereas no mutations could be detected for non-ESBL-producers. This showed that reduction of the membrane permeability in it-self is not enough to reach clinically relevant resistance levels, but in combi-nation with β-lactamases, ertapenem susceptibility is significantly decreased. Less reduction of the bacterial count after each dose indicated that pre-existing mutants were enriched for. Similar experiments were performed with 22 mg/L ertapenem to test whether a doubled dose (2 g daily for pa-tients) could prevent selection of resistant mutants, but ompR mutants ap-peared with the same frequency. This was further illustrated with static time-kill experiments. Isogenic wild type E. coli strains and ompR mutants from paper I were used in killing assays with static concentrations of ertapenem. Concentrations were chosen to be representative of the pharmacokinetic concentrations reached in humans after treatment with 1 g ertapenem once daily. Good bacterial killing was achieved against both wild type strains with and without the ESBL-plasmid during the first four hours after treat-ment. After 24 hours significant regrowth was detected from the wild type with the ESBL-plasmid. The ompR mutant with ESBL-plasmid was only affected by the highest concentrations of ertapenem, showing substantially

35

decreased kill-rates, and regrowth of all cultures comparable to the bacterial count of the untreated cultures was detected.

Altered PBPs and drug efflux cause high-level carbapenem resistance There are increasing reports of carbapenem resistance due to ESBL- and carbapenemase production mainly in K. pneumoniae and E. coli (Pitout and Laupland, 2008; Cantón, Akóva, et al., 2012; Glasner et al., 2013). In many cases the production of β-lactam degrading enzymes is accompanied by additional mechanisms that increase resistance levels, such as decreased permeability of the outer membrane (Martínez-Martínez et al., 1999; Nord-mann et al., 2012). Less is known about carbapenem resistance in the ab-sence of ESBL-production. In paper III we used a non-ESBL ompC/ompF deficient E. coli strain, to decrease the frequency of porin-mediated re-sistance mechanisms, to study the development of resistance during continu-ous selection with increasing levels of ertapenem or meropenem.

Carbapenem resistant populations evolved quickly during approximately 60 generations of selection. The growth rates of all populations were signifi-cantly decreased compared to the parental strain, which was expected after exposure of bacteria to these high antibiotic concentrations. Whole genome sequencing of isolated highly resistant clones revealed 3-5 mutations per clone. Independent of the carbapenem used for selection, each clone had mutations affecting the acr efflux system (acrB, acrD or acrR) and muta-tions in envZ. We observed a four fold increase in minimal inhibitory con-centrations when an envZ-mutation was introduced into the porin deficient parental strain. This was puzzling, as we expected envZ-mutations only to contribute through decreasing the amount of OmpC and OmpF, thus de-creasing permeability of the outer membrane. Clearly additional effects of envZ-mutations are contributing to resistance and one can imagine down-stream regulatory effects caused by mutations in this sensor-histidine kinase. Among the meropenem-selected populations one clone altered mrdA, the gene encoding penicillin-binding protein 2 (PBP2) and all clones showed mutations in genes that could affect the bacterial stringent response. When tested, we saw that induction of the stringent response contributed weakly to meropenem and ertapenem resistance, but it likely has a strong detrimental effect on bacterial fitness. All clones from ertapenem selections targeted ftsI, the gene encoding PBP3 and this gene has not been reported as a target in ertapenem resistance before our study.

The mutations observed to confer resistance in this study may be biased by the strong selective pressure resulting from continuous exposure to in-creasing concentrations of carbapenems. However, increasing clinical use of

36

carbapenems could balance the observed cost and may facilitate resistance development even in the absence of β-lactamases.

Cost and instability of GDA limit evolution of new genes We report in paper I that exposure to carbapenems can lead to increased copy numbers of β-lactamase genes. This is an additional resistance mecha-nism that frequently accompanies the presence of β-lactamases, increasing the concentration of degrading enzymes and so leading to higher resistance levels. In paper IV we investigate resistance levels that can be achieved by very high levels of β-lactamase gene amplification and to study the dynam-ics of gene duplication and amplification (GDA) of plasmid-borne antibiotic resistance genes. This allowed us to investigate implications for the evolu-tion of new carbapenemase genes and test the likelihood of various proposed models for de novo gene evolution based on our measurements of the cost of amplification and segregation rates.

Duplicated copies of existing genes are thought to be one major source for evolution of new genes (Sturtevant, 1925; Haldane, 1932; Muller, 1936; Lewis, 1951; Walsh and Walsh, 2003; Kohanski et al., 2010). By duplica-tion, one gene copy is free from selection pressure and can accumulate muta-tions that potentially lead to a new function. Many of these mutations will be deleterious and lead to non-functionalization and creation of pseudogenes. In the neofunctionalization model, one gene copy can evolve a new function while the other copy maintains the original function. According to the sub-functionalization model, after duplication, the activities of a gene with two separate selectable functions can be divided between the two duplicated genes through a series of common loss-of-function mutations. Both copies can then be selectively stabilized through their previous function and both functions can improve. The problem with the neo- and subfunctionalization model is that both assume GDA to be cost-free thus not explaining how GDAs can be maintained in the population sufficiently long and frequently enough to acquire rare mutations that provide a beneficial function. A solu-tion to this problem is the Innovation-Amplification-Divergence (IAD) model, where the cost and instability of tandem-duplications is balanced by a positive gene dosage effect that stabilizes tandem-arrays in the population. Genes can have weak secondary functions and this weak function can prove advantageous in a new environment. Frequent duplication and amplification mutations of the gene would increase the amounts of weakly bi-functional enzyme and increase the target size for rare beneficial mutations. As benefi-cial mutations accumulate, certain gene variants can be further improved through additional cycles of amplification and mutation. Selection for the

37

amplified state can be relaxed in accordance with increased new activity. The array may segregate back to one copy after a variant with sufficient activity for the new function arose. This can result in one gene copy per-forming the original activity and one copy performing the new activity.

Here we measured both the cost and instability of a set of strains with varying copy numbers of blaCTX-M-15, by first accumulating many gene cop-ies under meropenem selection and subsequently allowing the gene array to segregate back to fewer copies in the absence of selection. We determined the recombination rate and cost per additional gene copy, using a previously published model for homologous recombination (Pettersson et al., 2005; Pettersson et al., 2009). This model consists of a population of cells with varying numbers of blaCTX-M-15. As these cells replicate homologous recom-bination occurs and changes the copy numbers. The probability to increase or decrease the number of copies is the same, but additional copies confer a fitness cost and decreased growth rate. Cells with fewer copies will success-fully outcompete cells that have more copies and lowered growth rates.

We determined a cost of 1.2% ± 0.3% per cell per generation for each ad-ditional gene copy and a recombination rate of 0.13% ± 0.06% per cell per generation. This data shows that loss of gene copies is mainly driven by the fitness cost, rather than the rate of recombination. The cost and loss rate will affect the likelihood of evolution of a new gene function. In the neo- and subfunctionalization models, the probability of either one of the copies to acquire rare beneficial mutations is depending on the effective population size (cells with duplications in a population) and the selective advantage that this mutation gives over the single unmutated copy. The subfunctionaliza-tion model seems highly improbable for the cost and recombination rates of the region studied here, because the gene functions first need to be divided onto two units by inactivating mutations before one function can be im-proved. Evolution of a new gene function through neofunctionalization also seems unlikely. Here it would be necessary for the mutation to confer a strong selective advantage over the wild type gene to allow sufficiently fast fixation of the new function in the population. A two-step neofunctionaliza-tion is very unlikely even with a neutral first-step mutation followed by a highly selectable second-step mutation.

These results are applicable to the evolution of carbapenemase activity in CTX-M-15. New β-lactamase activities arise through successive point muta-tions (Salverda et al., 2010) (as in neofunctionalization model) or one can imagine the division of cephalosporin activity and trace-carbapenemase activity into two different copies (subfunctionalization model). In our study there is only selection for increasing the trace-carbapenemase activity of CTX-M-15 and the original cefotaxime activity might be lost during evolu-tion.

38

Concluding Remarks Bacteria have been present on earth for a very long time and evolved to overcome environmental challenges including antimicrobial substances pro-duced by other members of their microbiological environment. The intro-duction of antibiotics into clinical use in the 1940s posed a new challenge for bacteria and we witnessed the rapid mobilization and transfer of re-sistance mechanisms from environmental to pathogenic species on a global scale. Very fast evolution could be seen for β-lactamases, initially only de-grading a limited range of β-lactams but subsequently acquiring mutations to broaden their substrate spectrum (extended-spectrum β-lactamases, ESBLs). Only the newer β-lactam group, carbapenems, is stable in the presence of these enzymes leading to increased use of these antibiotics and increased selective pressure on bacteria to become resistant. Carbapenemases and oth-er carbapenem resistance mechanisms are already spreading in the clinical environment and it is of outmost importance to understand how bacteria (with and without the production of ESBLs) can develop resistance against carbapenems to avoid selection and further spread of resistance, and to pro-long the life span of these antibiotics.

In this thesis I have performed a number of studies that show that E. coli can develop resistance to carbapenems and that this development is favoured by expression of ESBLs. Expression of ESBLs may also allow the enrich-ment of ertapenem resistant mutants at concentrations found in serum during patient treatment. A mere dose increase of ertapenem may not be sufficient to counteract resistance development, as a doubled dose seemed unable to prevent resistant populations from accumulating. Gene duplication and am-plification (GDA) of ESBL-genes can confer resistance to β-lactams and is a very frequent mechanism of adaptation. GDA may be selected during car-bapenem treatment of ESBL-producers and can contribute to carbapenem resistance. This process may also be an important stepping-stone in the evo-lution of new gene functions, for instance improved carbapenem activity of the CTX-M-15 enzyme.

These finding may be of importance when deciding treatment regimes that avoid enrichment of resistant populations and limit evolution of novel resistance genes.

39

Future Perspectives

Two interesting follow-up projects emerged from the work performed in this thesis:

E. coli was chosen as the model organism for the work presented in this thesis. This is due to increased emergence of carbapenem resistant E. coli strains in hospital acquired infections and due to the well developed genetic tools available in this organism. However, carbapenem resistant K. pneu-moniae are more frequently associated with treatment failure and it would be worthwhile to determine the reason for this increased frequency in ESBL-outbreaks. In paper I we saw that in order for E. coli to develop resistance, both porins must be lost, either through loss-of-function mutations in the porin genes (at least two mutations necessary) or through mutations in the main regulators. Is the loss of both porins also necessary for K. pneumoniae to develop resistance or can there by further underlying mechanisms? Ac-cording to previous studies, production of carbapenemases or combination of porin loss and ESBL-production, similar to E. coli, comprise the main resistance mechanisms (Hernández-Allés et al., 1999; Jacoby et al., 2004; Martínez-Martínez, 2008). It has been reported that many clinical isolates of K. pneumoniae expressing ESBLs only express OmpK36 (homologue to OmpC in E. coli) and either have low expression or lost OmpK35 (homo-logue to OmpF in E. coli) (Bradford et al., 1997; Martínez-Martínez et al., 1999; Hernández-Allés et al., 1999; Crowley et al., 2002; Doménech-Sánchez et al., 2003). Carbapenem resistance has been shown as a result of β-lactamase expression and lack of one or more outer membrane proteins (Bradford et al., 1997; Martínez-Martínez et al., 1999; Crowley et al., 2002; Doménech-Sánchez et al., 2003; Salverda et al., 2010). However, no de-tailed study of the steps involved in carbapenem resistance development has been performed. In our studies of E. coli resistance we determined a reduc-tion of growth rate by 20% as the cost associated with carbapenem re-sistance. One can imagine that differences in this cost could be a likely fac-tor promoting K. pneumoniae infections.

Our preliminary studies with an isogenic pair of a K. pneumoniae clinical isolate, with and without expression of the β-lactamases from pUUH239.2, suggest a similar cost associated with carbapenem resistance as for E. coli. However, a possible difference in the step-by-step resistance mechanism

40

may allow K. pneumoniae to acquire carbapenem resistance more frequent-ly. Instead of having to loose expression of both porins, loss-of-function mutations in ompK36 are sufficient in our experiments to confer high car-bapenem resistance levels. Next generation genome sequencing of these resistant clones may give further information about differences in resistance development compared to E. coli.

The extended spectrum β-lactamase (ESBL) enzyme CTX-M-15 is one of the most frequent ESBL-enzymes worldwide and has nearly replaced all other ESBL-enzymes in Enterobacteriaceae (Cantón, González-Alba, et al., 2012). In combination with the increased usage of carbapenems in response to these and other ESBL-enzymes, the selective pressure to develop catalytic activity against carbapenems will increase. In paper I we showed that CTX-M-15 possesses trace-carbapenemase activity and in paper IV we evaluated the likelihood to evolve new gene functions after gene amplification of a region containing blaCTX-M-15. We were able to selectively amplify a large number of blaCTX-M-15 copies and stabilize the array using constant selection with meropenem, due to the increased dosage of trace-carbapenemase activi-ty of CTX-M-15. By amplification of the CTX-M-15 region we increased the size of the mutational target, thus increasing the chances for beneficial mutations. It is known that new β-lactamase activities arise through succes-sive point mutations in β-lactamase genes (Bershtein and Tawfik, 2008; Salverda et al., 2010) and it would be interesting to see whether continuous selection with meropenem may accumulate mutations increasing the trace-carbapenem activity of CTX-M-15. However, to date no mutant CTX-M-enzyme has been shown to have increased carbapenemase activity (Cantón, González-Alba, et al., 2012). This may indicate that either the changes re-quired for CTX-M-enzymes to evolve carbapenemase activity are substantial or that other resistance mechanisms are more frequent and evolved CTX-M-enzymes have not yet surfaced in the clinical environment.

Another approach would be to mutagenize the CTX-M-15 enzyme, then clone it onto plasmid vectors and subsequently select in the presence of car-bapenems to find more efficient variants. It is suggested that purifying selec-tion for the original activity may be advantageous for the evolution of novel functions (Bershtein and Tawfik, 2008). This seems to be the case for en-zyme activities with small trade-offs, meaning lowered activity of the origi-nal function after evolution of the new function. However, nothing is known about the trade-offs for mutations that would confer carbapenem activity in CTX-M-15, so it is difficult to predict whether purifying selection would be advantageous in this case.

41

Deutsche Zusammenfassung

Durch Zufall entdeckte Alexander Fleming 1928 das erste Antibiotikum, Penicillin, welches seit 1943 in Massenproduktion hergestellt wurde. Dies war der Auftakt für die Entdeckung von neuen Antibiotika und bereitete den Weg für die moderne Medizin, wo Transplantationen, Chemotherapie, Chi-rurgie und viele weitere Behandlungen ohne Antibiotika gar nicht möglich wären. Bakterien, die durch eine Antibiotikabehandlung nicht mehr am Wachstum gehindert werden oder absterben, werden als resistent bezeichnet. Bereits drei Jahre nach Beginn der verbreiteten Anwendung von Penicillin, wurden Penicillin-resistente Bakterien nachgewiesen.

Resistenz kann durch Mutationen im genetischen Code der Bakterien auf-treten oder durch Übertragung von bereits vorhandenen Resistenzgenen zwischen verschiedenen Arten. Mutationen sind rein zufällige Fehler im genetischen Code der Bakterien. Oft führen diese zum Tod des Bakteriums. In seltenen Fällen haben sie keinerlei Auswirkungen oder verleihen einen Vorteil z.B. Überleben in Gegenwart von Antibiotika. Bereits vorhandene Resistenzgene werden häufig durch Konjugation zwischen verschiedenen Bakterienstämmen ausgetauscht. Dabei werden Plasmide übertragen, kleine Einheiten von genetischem Material außerhalb des Genoms, die häufig meh-rere Resistenzgene gleichzeitig tragen. Durch diese und weitere Resistenz-mechanismen sind viele Antibiotika unbrauchbar geworden oder nicht mehr verlässlich effektiv. Es gibt mittlerweile Infektionen die gar nicht mehr mit Antibiotika bekämpft werden können.

Eine besonders schnelle Entwicklung und Verbreitung von Resistenz konnte für die Gruppe der β-Lactam-Antibiotika verzeichnet werden. Resis-tenz gegenüber diesen Antibiotika kann sich entwickeln indem Veränderun-gen der äußeren Zellmembran ihre Durchlässigkeit herabsetzt und damit den Zugang des Antibiotikums zu den Zielproteinen beeinträchtigt, durch Ver-änderungen in den Zielproteinen selber, durch Ausschleusung des Antibioti-kums oder durch die Produktion von β-Lactamasen. Diese Enzyme können den β-Lactam-Ring der Antibiotika öffnen und sie so inaktivieren. Ur-sprünglich waren β-Lactamasen nur gegen wenige spezielle β-Lactame ak-tiv, aber sie konnten nach und nach durch Mutationen ihr Spektrum erwei-tern und sind mittlerweile gegen alle β-Lactam Untergruppen mit Ausnahme der Carbapeneme aktiv (erweitertes-Spektrum β-Lactamasen: ESBL). Diese ESBL-Enzyme haben sich weltweit verbreitet und führen nun zu einer er-

42

höhten Einnahme von Carbapenemen und somit zum erhöhten Selektions-druck auf Bakterien. Das hat zur Folge, dass sich Resistenz gegen Carba-peneme schneller entwickelt. Deswegen ist es sehr wichtig den Prozess der Carbapeneme Resistenzentwicklung genau zu kennen, um Maßnahmen tref-fen zu können, die einer weiteren Verbreitung vorbeugen.

In meinen Studien habe ich gesehen, dass Escherichia coli relativ schnell Resistenz gegen Carbapeneme entwickeln kann und zwar durch eine Kom-bination aus ESBL-Produktion und verringerter Durchlässigkeit der Zell-membran (Paper I). Wir haben auch getestet ob geringere Konzentrationen von Carbapenemen, wie zum Beispiel die Konzentration im Körper eines Patienten, im Laufe einer 48-stündigen Carbapenem Behandlung zur Selek-tion von resistenten Bakterien ausreicht (Paper II). Diese Konzentrationen haben wir im Labor simuliert und sie repräsentieren die normale 1 Gram Dosis alle 24 Stunden. Nach dieser Selektion fanden wir die gleichen Resis-tenzmechanismen wie nach den höher dosierten Experimenten der ersten Studie. Die Bakterien mit ESBL-Enzymen konnten schnell die Durchlässig-keit ihrer Zellmembran verringern, aber Stämme die ursprünglich keine ESBL-Enzyme besaßen wurden nie resistent in dieser Studie. Wir testeten auch eine erhöhte Dosis (2 Gram alle 24 Stunden) unter der Annahme, dass es die Resistenzentwicklung verhindern würde, aber auch diese Konzentrati-onen waren nicht ausreichend.

In einer weiteren Studie wollten wir testen, ob E. coli ohne ESBL-Enzyme überhaupt resistent werden kann (Paper III). Wir benutzten einen Stamm dessen Membranen bereits weniger durchlässig waren, denn wir waren am Auftreten anderer Resistenzmechanismen interessiert. Tatsächlich sahen wir Mutationen in den Zielproteinen und Mutationen, die zur erhöhten Ausschleusung von Carbapenemen führten.

Bereits in unserer ersten Studie konnten wir zeigen, dass ESBL-Enzyme eine geringe Aktivität gegen Carbapeneme haben und dass Verdoppelung und weitere Vervielfältigung der ESBL-Gene im genetischen Code, zur ge-steigerten Produktion von ESBL-Enzymen führen und somit zur Resistenz beitragen können. Diese Vervielfältigung von genetischem Material ist im Allgemeinen instabil, denn sie verursacht biologische Kosten (Produktion von mehr genetischem Material, mehr Protein muss produziert werden und eventuelle negative Wechselwirkungen von der höheren Anzahl dieser Pro-teine mit anderen). Es ist von Vorteil für das Bakterium diese Vervielfälti-gung wieder zu verlieren, sobald die Antibiotika nicht mehr vorhanden sind. In unserer neuesten Studie brachten wir die Bakterien durch Selektion mit Carbapenemen dazu, die ESBL-Gene in hohem Maße zu vervielfältigen und nahmen dann Messungen vor über die Höhe der biologischen Kosten und wie schnell die Vervielfältigung wieder verschwindet. Es zeigt sich, dass die Kosten sehr hoch sind und dies trägt wiederum zum sehr schnellen Ver-schwinden der Vervielfältigung bei. Andererseits wäre es möglich, wenn die

43

Vervielfältigung lange genug im Genom der Bakterien vorhanden bleibt (durch lange Gegenwart von Carbapenemen in der Umgebung), dass sie durch Mutationen eine höhere Aktivität gegen Carbapeneme erhalten (das Enzym kann jetzt Carbapenemase genannt werden) und dann ist die Verviel-fältigung der ESBL-Enzyme nicht mehr notwendig. Vervielfältigung kann somit zur Entstehung neuer Resistenzgene beitragen.

E. coli kann also schnell und durch eine Vielzahl von Mechanismen Re-sistenz gegen Carbapeneme entwickeln und es reicht nicht aus, einfach die Dosis zu erhöhen. Bakterien können sich durch die Vervielfältigung von Resistenzgenen auch schnell an die Gegenwart von Carbapenemen anpassen und der schnelle Verlust dieser Vervielfältigungen nach dem Verschwinden der Antibiotika macht eine korrekte Diagnose dieser Infektionen schwierig. Bleiben die Vervielfältigungen lange im Genom der Bakterien vorhanden, droht auch Gefahr für die Entwicklung neuer Resistenzgene.

44

Acknowledgments

The path of a PhD student is full of obstacles set before you by nature, other people and yourself. The support of my colleagues and the friends I made along the way helped me overcome them and here are my thanks to all of you.

My biggest thanks goes to my supervisor Linus! It was pure luck that al-lowed me to start working with you, but nothing like luck that allowed me to see it through, because you had a plan! You have a way with understanding people, knowing when they need help or a little push and that makes you a great supervisor. I am very proud to have been your first PhD student and I hope many more will have the chance of working with you!

To my co-supervisor Dan. Thank you for all the help and advice you gave me with my project and for sharing your vast knowledge in the field. I am very grateful for the many opportunities you gave me to attend courses and conferences, and I gladly remember the amazing road trips with you.

To Andrea Hinas, my examiner. I am especially grateful for your help with my application for the Cold Spring Harbor course.

To my co-authors Mehreen Anjum, Thomas Tängdén, Otto Cars, Elisa-beth Löwdin and Otto Berg for valuable collaborations and to some of the inspiring scientists that I have had the chance to meet: Diarmaid Hughes, Beth Lazazzera, Fitnat Yildiz, Kelly Hughes and Alexander Böhm.

To past DA-group members: Anna Z, who still makes me afraid to cheat at card games. Chris and Sanna, my scientific role models. Song, for your patient help, especially with qPCR. I’ll never forget Super Svensson! My other office mates: Maria, who always seemed to be in a good mood and Peter, for unforgettable hot tub discussions. You all taught me what it’s like to be a PhD student.

To present members of the DA-group: Jocke, whose valuable scientific ad-vice more than made up for the numerous musical disagreements. The ever-whistling Hervé, merci for helpful advice, lots of humour and reminding me what a spoiled PhD student I was. Hava, who was the best lab-bench-neighbor I could have asked for. I admire your strength and determination against the odds. Erik G, thank you for being so observant and for always taking the time to listen. Lisa T, the group’s Marvel expert, Game of

45

Thrones expert, language expert… etc. Continue keeping the group in line! Ulrika, sorry for always asking you first about… well anything, but to me you were running the lab and you always knew what to do. Karin, thank you for many great laughs and plenty of sub-MIC discussions. To the next generations of DA-PhD students: Anna K, Lisa A, Michael, Sohaib and Erik L – best of luck with your work. I know you will be great. Jon, the humble mastermind. Let’s open this baby up!

To Mehreen and Cecilia, you quickly became more than just exam-students to me! I am glad I got to work with you.

To my dear colleagues: Nizar and Göte for good feedback at seminars. Gerrit, Jessica, Cha Sha and Doug for great times at retreats, seminars, fikas, lunch breaks and outside of BMC. You made D7:3 a happier place!

Britta and Elin at ICM and Rachel for many non-work related occasions.

Many more people at IMBIM for good times at IPhA, pubs, crayfish and cheese and wine parties :) The IMBIM staff, especially Barbro and Erika, for travel advice and help with all the paper work.

Special thanks to Alex H for proofreading this thesis. I still owe you, but maybe I can make up for it with some champagne.

To my friends: Alex B, Susanna, Anja & Anja, Lára, Eva, and Julia. Some of you I met before I started my PhD studies and some only 1.6 years ago. However, the experiences we shared make it feel like a lifetime, in the best of ways. Thanks to you, Uppsala and my PhD studies will remain among the best times of my life.

Franzi and Kasia, your strength is inspiring. Thank you for being there for me whenever I needed you. Marius and Mike, for the good old times, the healthy dose of sarcasm and refreshing lunch conversations. I hope you’ll never need to call me ‘Darling’! You all deserve special thanks, because you’ve had to listen to my venting longer than anyone. Thank you for your friendship!

Danke, an meine Eltern, die mich immer ermutigt haben. Ohne eure Unter-stützung  hätte ich meine großen Schritte in die Welt nicht gewagt. Meine Schwester Sarah, die mir die Welt bedeutet und auf die ich mich immer ver-lassen kann! Küssi, L.

46

References

Aarestrup, F.M. (2005) Veterinary drug usage and antimicrobial resistance in bacte-ria of animal origin. Basic & clinical pharmacology & toxicology 96: 271–281.

Abraham, E.P., and Chain, E. (1940) An Enzyme from Bacteria able to Destroy Penicillin. Nature 146: 837–837.

Aiba, H., and Mizuno, T. (1990) Phosphorylation of a bacterial activator protein, OmpR, by a protein kinase, EnvZ, stimulates the transcription of the ompF and ompC genes in Escherichia coli. FEBS Lett 261: 19–22.

Ambler, R.P. (1980) The Structure of beta-lactamases. Philosophical Transactions of the Royal Society B: Biological Sciences 289: 321–331.

Anderson, P., and Roth, J. (1981) Spontaneous tandem genetic duplications in Sal-monella typhimurium arise by unequal recombination between rRNA (rrn) cis-trons. Proc Natl Acad Sci USA 78: 3113–3117.

Andersson, D.I., and Hughes, D. (2009) Gene amplification and adaptive evolution in bacteria. Annu Rev Genet 43: 167–195.

Andersson, D.I., and Hughes, D. (2010) Antibiotic resistance and its cost: is it pos-sible to reverse resistance? Nat Rev Microbiol 8: 260–271.

Andersson, D.I., and Levin, B.R. (1999) The biological cost of antibiotic resistance. Curr Opin Microbiol 2: 489–493.

Andersson, D.I., Slechta, E.S., and Roth, J.R. (1998) Evidence that gene amplifica-tion underlies adaptive mutability of the bacterial lac operon. Science 282: 1133–1135.

Andrews, J.M. (2001) Determination of minimum inhibitory concentrations. Journal of Antimicrobial Chemotherapy 48: 5–16.

Barlow, M., and Hall, B.G. (2002) Phylogenetic analysis shows that the OXA beta-lactamase genes have been on plasmids for millions of years. J Mol Evol 55: 314–321.

Barlow, M., Reik, R.A., Jacobs, S.D., Medina, M., Meyer, M.P., McGowan, J.E., and Tenover, F.C. (2008) High rate of mobilization for blaCTX-Ms. Emerging Infect Dis 14: 423–428.

Bauernfeind, A., Grimm, H., and Schweighart, S. (1990) A new plasmidic cefotax-imase in a clinical isolate of Escherichia coli. Infection 18: 294–298.

Bergthorsson, U., Andersson, D.I., and Roth, J.R. (2007) Ohno's dilemma: evolution of new genes under continuous selection. Proc Natl Acad Sci USA 104: 17004–17009.

Bershtein, S., and Tawfik, D.S. (2008) Ohno's Model Revisited: Measuring the Frequency of Potentially Adaptive Mutations under Various Mutational Drifts. Molecular Biology and Evolution 25: 2311–2318.

Bertini, A., Poirel, L., Bernabeu, S., Fortini, D., Villa, L., Nordmann, P., and Carat-toli, A. (2007) Multicopy blaOXA-58 gene as a source of high-level resistance to carbapenems in Acinetobacter baumannii. Antimicrob Agents Chemother 51: 2324–2328.

47

Bertsche, U., Kast, T., Wolf, B., Fraipont, C., Aarsman, M.E., Kannenberg, K., et al. (2006) Interaction between two murein (peptidoglycan) synthases, PBP3 and PBP1B, in Escherichia coli. Mol Microbiol 61: 675–690.

Blattner, F.R., Plunkett, G., Bloch, C.A., Perna, N.T., Burland, V., Riley, M., et al. (1997) The complete genome sequence of Escherichia coli K-12. Science 277: 1453–1462.

Bonnet, R. (2004) Growing group of extended-spectrum beta-lactamases: the CTX-M enzymes. Antimicrob Agents Chemother 48: 1–14.

Boyd, D.A., Tyler, S., Christianson, S., McGeer, A., Muller, M.P., Willey, B.M., et al. (2004) Complete nucleotide sequence of a 92-kilobase plasmid harboring the CTX-M-15 extended-spectrum beta-lactamase involved in an outbreak in long-term-care facilities in Toronto, Canada. Antimicrob Agents Chemother 48: 3758–3764.

Bradford, P.A. (2001) Extended-spectrum beta-lactamases in the 21st century: char-acterization, epidemiology, and detection of this important resistance threat. Clin Microbiol Rev 14: 933–951.

Bradford, P.A., Urban, C., Mariano, N., Projan, S.J., Rahal, J.J., and Bush, K. (1997) Imipenem resistance in Klebsiella pneumoniae is associated with the combination of ACT-1, a plasmid-mediated AmpC beta-lactamase, and the foss of an outer membrane protein. Antimicrob Agents Chemother 41: 563–569.

Brochet, M., Couvé, E., Zouine, M., Poyart, C., and Glaser, P. (2008) A naturally occurring gene amplification leading to sulfonamide and trimethoprim re-sistance in Streptococcus agalactiae. J Bacteriol 190: 672–680.

Bush, K. (1989) Classification of beta-lactamases: groups 1, 2a, 2b, and 2b'. Antimi-crob Agents Chemother 33: 264.

Bush, K., and Jacoby, G.A. (2010) Updated functional classification of beta-lactamases. Antimicrob Agents Chemother 54: 969–976.

Bush, K., Smith, S.A., Ohringer, S., Tanaka, S.K., and Bonner, D.P. (1987) Im-proved sensitivity in assays for binding of novel beta-lactam antibiotics to peni-cillin-binding proteins of Escherichia coli. Antimicrob Agents Chemother 31: 1271–1273.

Cabello, F.C. (2006) Heavy use of prophylactic antibiotics in aquaculture: a growing problem for human and animal health and for the environment. Environ Micro-biol 8: 1137–1144.

Cantón, R., Akóva, M., Carmeli, Y., Giske, C.G., Glupczynski, Y., Gniadkowski, M., et al. (2012) Rapid evolution and spread of carbapenemases among Entero-bacteriaceae in Europe. Clin Microbiol Infect 18: 413–431.

Cantón, R., and Coque, T.M. (2006) The CTX-M beta-lactamase pandemic. Curr Opin Microbiol 9: 466–475.

Cantón, R., González-Alba, J.M., and Galán, J.C. (2012) CTX-M Enzymes: Origin and Diffusion. Front Microbiol 3.

Cantón, R., Novais, A., Valverde, A., Machado, E., Peixe, L., Baquero, F., and Coque, T.M. (2008) Prevalence and spread of extended-spectrum β-lactamase-producing Enterobacteriaceae in Europe. Clin Microbiol Infect 14: 144–153.

Chain, E., Florey, H.W., Hardner, A.D., Heatley, N.G., Jennings, M.A., ORR-EWING, J., and Sanders, A.G. (1940) Penicillin as a Chemotherapeutic Agent. Lancet 239: 226–228.

Criswell, D., Tobiason, V.L., Lodmell, J.S., and Samuels, D.S. (2006) Mutations conferring aminoglycoside and spectinomycin resistance in Borrelia burgdor-feri. Antimicrob Agents Chemother 50: 445–452.

48

Crowley, B., Benedí, V.J., and Doménech-Sánchez, A. (2002) Expression of SHV-2 beta-lactamase and of reduced amounts of OmpK36 porin in Klebsiella pneu-moniae results in increased resistance to cephalosporins and carbapenems. An-timicrob Agents Chemother 46: 3679–3682.

Curtis, N.A., Orr, D., Ross, G.W., and Boulton, M.G. (1979) Affinities of penicillins and cephalosporins for the penicillin-binding proteins of Escherichia coli K-12 and their antibacterial activity. Antimicrob Agents Chemother 16: 533–539.

Davies, T.A., Shang, W., Bush, K., and Flamm, R.K. (2008) Affinity of doripenem and comparators to penicillin-binding proteins in Escherichia coli and Pseudo-monas aeruginosa. Antimicrob Agents Chemother 52: 1510–1512.

Delcour, A.H. (2009) Outer membrane permeability and antibiotic resistance. Bio-chim Biophys Acta 1794: 808–816.

Devonshire, A.L., and Field, L.M. (1991) Gene amplification and insecticide re-sistance. Annual review of entomology 36: 1–21.

Doménech-Sánchez, A., Hernández-Allés, S., Martínez-Martínez, L., Benedí, V.J., and Albertí, S. (1999) Identification and characterization of a new porin gene of Klebsiella pneumoniae: its role in beta-lactam antibiotic resistance. J Bacteriol 181: 2726–2732.

Doménech-Sánchez, A., Martínez-Martínez, L., Hernández-Allés, S., del Carmen Conejo, M., Pascual, A., Tomás, J.M., et al. (2003) Role of Klebsiella pneu-moniae OmpK35 porin in antimicrobial resistance. Antimicrob Agents Chemother 47: 3332–3335.

Doumith, M., Ellington, M.J., Livermore, D.M., and Woodford, N. (2009) Molecu-lar mechanisms disrupting porin expression in ertapenem-resistant Klebsiella and Enterobacter spp. clinical isolates from the UK. J Antimicrob Chemother 63: 659–667.

Elliott, K.T., Cuff, L.E., and Neidle, E.L. (2013) Copy number change: evolving views on gene amplification. Future Microbiol 8: 887–899.

Falagas, M.E., and Karageorgopoulos, D.E. (2009) Extended-spectrum beta-lactamase-producing organisms. J Hosp Infect 73: 345–354.

Fernández, L., and Hancock, R.E.W. (2012) Adaptive and mutational resistance: role of porins and efflux pumps in drug resistance. Clin Microbiol Rev 25: 661–681.

Fleming, A. (1929) On the antibacterial action of cultures of a penicillium, with special reference to their use in the isolation of B. influenzae. British Journal of Experimental Pathology 10: 226-236.

Force, A., Lynch, M., Pickett, F.B., Amores, A., Yan, Y.-L., and Postlethwait, J. (1999) Preservation of Duplicate Genes by Complementary, Degenerative Mu-tations. Genetics 151: 1531–1545.

Forst, S., Delgado, J., and Inouye, M. (1989a) Phosphorylation of OmpR by the osmosensor EnvZ modulates expression of the ompF and ompC genes in Esche-richia coli. Proc Natl Acad Sci USA 86: 6052–6056.

Forst, S.A., Delgado, J., and Inouye, M. (1989b) DNA-binding properties of the transcription activator (OmpR) for the upstream sequences of ompF in Esche-richia coli are altered by envZ mutations and medium osmolarity. J Bacteriol 171: 2949–2955.

Frost, L.S., and Koraimann, G. (2010) Regulation of bacterial conjugation: balanc-ing opportunity with adversity. Future Microbiol 5: 1057–1071.

García-Sureda, L., Doménech-Sánchez, A., Barbier, M., Juan, C., Gascó, J., and Albertí, S. (2011) OmpK26, a novel porin associated with carbapenem re-

49

sistance in Klebsiella pneumoniae. Antimicrob Agents Chemother 55: 4742–4747.

Giske, C.G., Sundsfjord, A.S., Kahlmeter, G., Woodford, N., Nordmann, P., Pater-son, D.L., et al. (2009) Redefining extended-spectrum beta-lactamases: balanc-ing science and clinical need. Journal of Antimicrobial Chemotherapy 63: 1–4.

Glasner, C., Albiger, B., Buist, G., Tambić Andrasević, A., Cantón, R., Carmeli, Y., et al. (2013) Carbapenemase-producing Enterobacteriaceae in Europe: a survey among national experts from 39 countries, February 2013. Euro Surveill 18.

Gniadkowski, M. (2008) Evolution of extended-spectrum beta-lactamases by muta-tion. Clin Microbiol Infect 14: 11–32.

Gullberg, E., Cao, S., Berg, O.G., Ilbäck, C., Sandegren, L., Hughes, D., and An-dersson, D.I. (2011) Selection of resistant bacteria at very low antibiotic con-centrations. PLoS Pathog 7: e1002158.

Haack, K.R., and Roth, J.R. (1995) Recombination between chromosomal IS200 elements supports frequent duplication formation in Salmonella typhimurium. Genetics 141: 1245–1252.

Haldane, J.B.S. (1932) The Causes of Evolution, Etc. [With Plates.]. Longmans, London.

Hammond, D.S., Schooneveldt, J.M., Nimmo, G.R., Huygens, F., and Giffard, P.M. (2005) bla(SHV) Genes in Klebsiella pneumoniae: different allele distributions are associated with different promoters within individual isolates. Antimicrob Agents Chemother 49: 256–263.

Harlocker, S.L., Bergstrom, L., and Inouye, M. (1995) Tandem binding of six OmpR proteins to the ompF upstream regulatory sequence of Escherichia coli. J Biol Chem 270: 26849–26856.

Hedge, P.J., and Spratt, B.G. (1984) A gene fusion that localises the penicillin-binding domain of penicillin-binding protein 3 of Escherichia coli. FEBS Lett 176: 179–184.

Hedge, P.J., and Spratt, B.G. (1985) Amino acid substitutions that reduce the affini-ty of penicillin-binding protein 3 of Escherichia coli for cephalexin. Eur J Bio-chem 151: 111–121.

Hernández-Allés, S., Albertí, S., Alvarez, D., Doménech-Sánchez, A., Martínez-Martínez, L., Gil, J., et al. (1999) Porin expression in clinical isolates of Klebsiella pneumoniae. Microbiology (Reading, Engl) 145: 673–679.

Hill, C.W., Foulds, J., Soll, L., and Berg, P. (1969) Instability of a missense sup-pressor resulting from a duplication of genetic material. J Mol Biol 39: 563–581.

Hiramatsu, K., Cui, L., Kuroda, M., and Ito, T. (2001) The emergence and evolution of methicillin-resistant Staphylococcus aureus. Trends in Microbiology 9: 486–493.

Hong, T., Moland, E.S., Abdalhamid, B., Hanson, N.D., Wang, J., Sloan, C., et al. (2005) Escherichia coli: development of carbapenem resistance during therapy. Clin Infect Dis 40: e84–6.

Huang, T.-W., Chen, T.-L., Chen, Y.-T., Lauderdale, T.-L., Liao, T.-L., Lee, Y.-T., et al. (2013) Copy Number Change of the NDM-1 Sequence in a Multidrug-Resistant Klebsiella pneumoniae Clinical Isolate. PLoS ONE 8: e62774.

Hudson, R.E., Bergthorsson, U., Roth, J.R., and Ochman, H. (2002) Effect of chro-mosome location on bacterial mutation rates. Molecular Biology and Evolution 19: 85–92.

Ikeda, H., Shiraishi, K., and Ogata, Y. (2004) Illegitimate recombination mediated by double-strand break and end-joining in. Advances in Biophysics 38: 3–20.

50

Jacoby, G.A. (2006) Beta-lactamase nomenclature. Antimicrob Agents Chemother 50: 1123–1129.

Jacoby, G.A., and Munoz-Price, L.S. (2005) The New β-Lactamases. N Engl J Med 352: 380–391.

Jacoby, G.A., Mills, D.M., and Chow, N. (2004) Role of beta-lactamases and porins in resistance to ertapenem and other beta-lactams in Klebsiella pneumoniae. An-timicrob Agents Chemother 48: 3203–3206.

Johnsborg, O., and Håvarstein, L.S. (2009) Regulation of natural genetic transfor-mation and acquisition of transforming DNA in Streptococcus pneumoniae. FEMS Microbiol Rev 33: 627–642.

Kaessmann, H. (2010) Origins, evolution, and phenotypic impact of new genes. Genome Research 20: 1313–1326.

Kattan, J.N., Villegas, M.V., and Quinn, J.P. (2008) New developments in car-bapenems. Clinical Microbiology and Infection 14: 1102–1111.

Källman, O., Giske, C.G., Samuelsen, Ø., Wretlind, B., Kalin, M., and Olsson-Liljequist, B. (2009) Interplay of Efflux, Impermeability, and AmpC Activity Contributes to Cefuroxime Resistance in Clinical, Non-ESBL-Producing Iso-lates of Escherichia coli. Microb Drug Resist 15: 91–95.

Knothe, H., Shah, P., Krcmery, V., Antal, M., and Mitsuhashi, S. (1983) Transfera-ble resistance to cefotaxime, cefoxitin, cefamandole and cefuroxime in clinical isolates of Klebsiella pneumoniae and Serratia marcescens. Infection 11: 315–317.

Koga, T., Sugihara, C., Kakuta, M., Masuda, N., Namba, E., and Fukuoka, T. (2009) Affinity of Tomopenem (CS-023) for penicillin-binding proteins in Staphylo-coccus aureus, Escherichia coli, and Pseudomonas aeruginosa. Antimicrob Agents Chemother 53: 1238–1241.

Kohanski, M.A., Dwyer, D.J., and Collins, J.J. (2010) How antibiotics kill bacteria: from targets to networks. Nat Rev Microbiol 8: 423–435.

Kohler, J., Dorso, K.L., Young, K., Hammond, G.G., Rosen, H., Kropp, H., and Silver, L.L. (1999) In vitro activities of the potent, broad-spectrum carbapenem MK-0826 (L-749,345) against broad-spectrum beta-lactamase-and extended-spectrum beta-lactamase-producing Klebsiella pneumoniae and Escherichia coli clinical isolates. Antimicrob Agents Chemother 43: 1170–1176.

Koskiniemi, S., Sun, S., Berg, O.G., and Andersson, D.I. (2012) Selection-Driven Gene Loss in Bacteria. PLoS Genet 8: e1002787.

Kugelberg, E., Kofoid, E., Reams, A.B., Andersson, D.I., and Roth, J.R. (2006) Multiple pathways of selected gene amplification during adaptive mutation. Proceedings of the National Academy of Sciences 103: 17319–17324.

Kunz, A.N., Begum, A.A., Wu, H., D'Ambrozio, J.A., Robinson, J.M., Shafer, W.M., et al. (2012) Impact of fluoroquinolone resistance mutations on gono-coccal fitness and in vivo selection for compensatory mutations. J Infect Dis 205: 1821–1829.

Kümmerer, K. (2009) Antibiotics in the aquatic environment--a review--part I. Chemosphere 75: 417–434.

la Cruz, De, M.Á., and Calva, E. (2010) The complexities of porin genetic regula-tion. J Mol Microbiol Biotechnol 18: 24–36.

Lambert, P.A. (2005) Bacterial resistance to antibiotics: modified target sites. Adv Drug Deliv Rev 57: 1471–1485.

Lane, N., and Martin, W. (2010) The energetics of genome complexity. Nature 467: 929–934.

51

Lartigue, M.-F., Poirel, L., Poyart, C., Réglier-Poupet, H., and Nordmann, P. (2007) Ertapenem resistance of Escherichia coli. Emerging Infect Dis 13: 315–317.

Lee, J.A., and Lupski, J.R. (2006) Genomic rearrangements and gene copy-number alterations as a cause of nervous system disorders. Neuron 52: 103–121.

Lewis, E.B. (1951) Pseudoallelism and gene evolution. Cold Spring Harbor Sympo-sia on Quantitative Biology 16: 159–174.

Lin, R.-J., Capage, M., and Hill, C.W. (1984) A repetitive DNA sequence, rhs, re-sponsible for duplications within the Escherichia coli K-12 chromosome. J Mol Biol 177: 1–18.

Lind, P.A., Tobin, C., Berg, O.G., Kurland, C.G., and Andersson, D.I. (2010) Com-pensatory gene amplification restores fitness after inter-species gene replace-ments. Mol Microbiol 75: 1078–1089.

Lindahl, T., and Wood, R.D. (1999) Quality control by DNA repair. Science 286: 1897–1905.

Livermore, D.M. (1992) Carbapenemases. J Antimicrob Chemother 29: 609–613. Livermore, D.M. (2008) Defining an extended-spectrum beta-lactamase. Clin Mi-

crobiol Infect 14 Suppl 1: 3–10. Lovett, S.T., Drapkin, P.T., Sutera, V.A., and Gluckman-Peskind, T.J. (1993) A

sister-strand exchange mechanism for recA-independent deletion of repeated DNA sequences in Escherichia coli. Genetics 135: 631–642.

Löwdin, E., Odenholt, I., Bengtsson, S., and Cars, O. (1996) Pharmacodynamic effects of sub-MICs of benzylpenicillin against Streptococcus pyogenes in a newly developed in vitro kinetic model. Antimicrob Agents Chemother 40: 2478–2482.

Luo, N., Pereira, S., Sahin, O., Lin, J., Huang, S., Michel, L., and Zhang, Q. (2005) Enhanced in vivo fitness of fluoroquinolone-resistant Campylobacter jejuni in the absence of antibiotic selection pressure. Proceedings of the National Acad-emy of Sciences 102: 541–546.

Lupski, J.R. (2007) Genomic rearrangements and sporadic disease. Nat Genet 39: S43–7.

Lynch, M. (2000) The Evolutionary Fate and Consequences of Duplicate Genes. Science 290: 1151–1155.

Lynch, M., and Force, A. (2000) The Probability of Duplicate Gene Preservation by Subfunctionalization. Genetics 154: 459–473.

Lytsy, B., Sandegren, L., Tano, E., Torell, E., Andersson, D.I., and Melhus, A. (2008) The first major extended-spectrum beta-lactamase outbreak in Scandina-via was caused by clonal spread of a multiresistant Klebsiella pneumoniae pro-ducing CTX-M-15. APMIS 116: 302–308.

Malléa, M., Chevalier, J., Bornet, C., Eyraud, A., Davin-Regli, A., Bollet, C., and Pagès, J.M. (1998) Porin alteration and active efflux: two in vivo drug re-sistance strategies used by Enterobacter aerogenes. Microbiology (Reading, Engl) 144 ( Pt 11): 3003–3009.

Martínez-Martínez, L. (2008) Extended-spectrum beta-lactamases and the permea-bility barrier. Clin Microbiol Infect 14 Suppl 1: 82–89.

Martínez-Martínez, L., Pascual, A., Hernández-Allés, S., Alvarez-Díaz, D., Suárez, A.I., Tran, J., et al. (1999) Roles of β-Lactamases and Porins in Activities of Carbapenems and Cephalosporins against Klebsiella pneumoniae. Antimicrob Agents Chemother 43: 1669–1673.

Massova, I., and Mobashery, S. (1998) Kinship and diversification of bacterial peni-cillin-binding proteins and beta-lactamases. Antimicrob Agents Chemother 42: 1–17.

52

Matfield, M., Badawi, R., and Brammar, W.J. (1985) Rec-dependent and Rec-independent recombination of plasmid-borne duplication in Escherichia coli K12. Mol Gen Genet 199: 518–523.

Matsumoto, Y., Ikeda, F., Kamimura, T., Yokota, Y., and Mine, Y. (1988) Novel plasmid-mediated beta-lactamase from Escherichia coli that inactivates oxy-imino-cephalosporins. Antimicrob Agents Chemother 32: 1243–1246.

McCarroll, S.A., and Altshuler, D.M. (2007) Copy-number variation and association studies of human disease. Nat Genet 39: S37–S42.

Mileyko, Y., Joh, R.I., and Weitz, J.S. (2008) Small-scale copy number variation and large-scale changes in gene expression. Proceedings of the National Acad-emy of Sciences 105: 16659–16664.

Moellering, R.C., Eliopoulos, G.M., and Sentochnik, D.E. (1989) The Carbapenems: New Broad Spectrum beta-lactam Antibiotics. J Antimicrob Chemother 24: 1–7.

Muller, H.J. (1936) Bar Duplication. Science 83: 528–530. Muñiz, C.C., Zelaya, T.E.C., Esquivel, G.R., and Fernández, F.J. (2007) Penicillin

and cephalosporin production: A historical perspective. Revista Latinoamerica-na de Microbiología 49: 88–98.

Musher, D.M., Dowell, M.E., Shortridge, V.D., Flamm, R.K., Jorgensen, J.H., Le Magueres, P., and Krause, K.L. (2002) Emergence of macrolide resistance dur-ing treatment of pneumococcal pneumonia. N Engl J Med 346: 630–631.

Naas, T., and Nordmann, P. (1999) OXA-Type beta-lactamases. Current pharma-ceutical design 5: 865–879.

Naseer, U., and Sundsfjord, A. (2011) The CTX-M conundrum: dissemination of plasmids and Escherichia coli clones. Microb Drug Resist 17: 83–97.

Näsvall, J., Sun, L., Roth, J.R., and Andersson, D.I. (2012) Real-time evolution of new genes by innovation, amplification, and divergence. Science 338: 384–387.

Nelson, D.E., and Young, K.D. (2001) Contributions of PBP 5 and DD-carboxypeptidase penicillin binding proteins to maintenance of cell shape in Escherichia coli. J Bacteriol 183: 3055–3064.

Neuberger, M.S., and Hartley, B.S. (1981) Structure of an Experimentally Evolved Gene Duplication Encoding Ribitol Dehydrogenase in a Mutant of Klebsiella aerogenes. Microbiology 122: 181–191.

Nichols, B.P., and Guay, G.G. (1989) Gene amplification contributes to sulfonamide resistance in Escherichia coli. Antimicrob Agents Chemother 33: 2042–2048.

Nicoletti, G., Russo, G., and Bonfiglio, G. (2002) Recent developments in car-bapenems. Expert Opin Investig Drugs 11: 529–544.

Nicoloff, H., and Andersson, D.I. (2013) Lon protease inactivation, or translocation of the longene, potentiate bacterial evolution to antibiotic resistance. Mol Mi-crobiol 90: 1233–1248.

Nicoloff, H., Perreten, V., and Levy, S.B. (2007) Increased genome instability in Escherichia coli lon mutants: relation to emergence of multiple-antibiotic-resistant (Mar) mutants caused by insertion sequence elements and large tandem genomic amplifications. Antimicrob Agents Chemother 51: 1293–1303.

Nikaido, H. (1989) Outer membrane barrier as a mechanism of antimicrobial re-sistance. Antimicrob Agents Chemother 33: 1831.

Nikaido, H. (2001) Preventing drug access to targets: cell surface permeability bar-riers and active efflux in bacteria. Semin Cell Dev Biol 12: 215–223.

Nikaido, H. (2003) Molecular basis of bacterial outer membrane permeability revis-ited. Microbiol Mol Biol Rev 67: 593–656.

Nikaido, H., and Rosenberg, E.Y. (1983) Porin channels in Escherichia coli: studies with liposomes reconstituted from purified proteins. J Bacteriol 153: 241–252.

53

Nilsson, A.I., Zorzet, A., Kanth, A., Dahlström, S., Berg, O.G., and Andersson, D.I. (2006) Reducing the fitness cost of antibiotic resistance by amplification of ini-tiator tRNA genes. Proc Natl Acad Sci USA 103: 6976–6981.

Nordmann, P., and Poirel, L. (2002) Emerging carbapenemases in Gram-negative aerobes. Clin Microbiol Infect 8: 321–331.

Nordmann, P., Dortet, L., and Poirel, L. (2012) Carbapenem resistance in Entero-bacteriaceae: here is the storm! Trends in Molecular Medicine 18: 263–272.

Nordmann, P., Naas, T., and Poirel, L. (2011) Global spread of Carbapenemase-producing Enterobacteriaceae. Emerging Infect Dis 17: 1791–1798.

Normark, S., Edlund, T., Grundström, T., Bergström, S., and Wolf-Watz, H. (1977) Escherichia coli K-12 mutants hyperproducing chromosomal beta-lactamase by gene repetitions. J Bacteriol 132: 912–922.

Ohno, S. (1970) Evolution by gene duplication. Springer Verlag, New York. Oteo, J., Delgado-Iribarren, A., Vega, D., Bautista, V., Rodríguez, M.C., Velasco,

M., et al. (2008) Emergence of imipenem resistance in clinical Escherichia coli during therapy. Int J Antimicrob Agents 32: 534–537.

Pagès, J.-M., James, C.E., and Winterhalter, M. (2008) The porin and the permeat-ing antibiotic: a selective diffusion barrier in Gram-negative bacteria. Nat Rev Microbiol 6: 893–903.

Papp-Wallace, K.M., Endimiani, A., Taracila, M.A., and Bonomo, R.A. (2011) Carbapenems: past, present, and future. Antimicrob Agents Chemother 55: 4943–4960.

Paulander, W., Andersson, D.I., and Maisnier-Patin, S. (2010) Amplification of the gene for isoleucyl-tRNA synthetase facilitates adaptation to the fitness cost of mupirocin resistance in Salmonella enterica. Genetics 185: 305–312.

Paulander, W., Maisnier-Patin, S., and Andersson, D.I. (2007) Multiple mechanisms to ameliorate the fitness burden of mupirocin resistance in Salmonella typhi-murium. Mol Microbiol 64: 1038–1048.

Petit, M.A., Noirot, P., Morel-Deville, F., and Ehrlich, S.D. (1992) Induction of DNA amplification in the Bacillus subtilis chromosome. EMBO J 11: 1317.

Pettersson, M.E., Andersson, D.I., Roth, J.R., and Berg, O.G. (2005) The amplifica-tion model for adaptive mutation: simulations and analysis. Genetics 169: 1105–1115.

Pettersson, M.E., Sun, S., Andersson, D.I., and Berg, O.G. (2009) Evolution of new gene functions: simulation and analysis of the amplification model. Genetica 135: 309–324.

Philippon, L.N., Naas, T., Bouthors, A.T., Barakett, V., and Nordmann, P. (1997) OXA-18, a class D clavulanic acid-inhibited extended-spectrum beta-lactamase from Pseudomonas aeruginosa. Antimicrob Agents Chemother 41: 2188–2195.

Piddock, L.J.V. (2006a) Clinically relevant chromosomally encoded multidrug re-sistance efflux pumps in bacteria. Clin Microbiol Rev 19: 382–402.

Piddock, L.J.V. (2006b) Multidrug-resistance efflux pumps - not just for resistance. Nat Rev Microbiol 4: 629–636.

Pitout, J., and Laupland, K.B. (2008) Extended-spectrum β-lactamase-producing Enterobacteriaceae: an emerging public-health concern. Lancet Infect Dis 8: 159–166.

Poirel, L., Decousser, J.-W., and Nordmann, P. (2003) Insertion sequence ISEcp1B is involved in expression and mobilization of a bla(CTX-M) beta-lactamase gene. Antimicrob Agents Chemother 47: 2938–2945.

54

Poirel, L., Héritier, C., Spicq, C., and Nordmann, P. (2004) In vivo acquisition of high-level resistance to imipenem in Escherichia coli. J Clin Microbiol 42: 3831–3833.

Popa, O., and Dagan, T. (2011) Trends and barriers to lateral gene transfer in pro-karyotes. Curr Opin Microbiol 14: 615–623.

Popham, D.L., and Young, K.D. (2003) Role of penicillin-binding proteins in bacte-rial cell morphogenesis. Curr Opin Microbiol 6: 594–599.

Pratt, L.A., Hsing, W., Gibson, K.E., and Silhavy, T.J. (1996) From acids to osmZ: multiple factors influence synthesis of the OmpF and OmpC porins in Esche-richia coli. Mol Microbiol 20: 911–917.

Pränting, M., and Andersson, D.I. (2011) Escape from growth restriction in small colony variants of Salmonella typhimurium by gene amplification and mutation. Mol Microbiol 79: 305–315.

Queenan, A.M., and Bush, K. (2007) Carbapenemases: the versatile beta-lactamases. Clin Microbiol Rev 20: 440–458.

Ramadhan, A.A. (2005) Survivability of vancomycin resistant enterococci and fit-ness cost of vancomycin resistance acquisition. Journal of Clinical Pathology 58: 744–746.

Rasmussen, B.A., Keeney, D., Yang, Y., and Bush, K. (1994) Cloning and expres-sion of a cloxacillin-hydrolyzing enzyme and a cephalosporinase from Aer-omonas sobria AER 14M in Escherichia coli: requirement for an E. coli chro-mosomal mutation for efficient expression of the class D enzyme. Antimicrob Agents Chemother 38: 2078–2085.

Reams, A.B., Kofoid, E., Savageau, M., and Roth, J.R. (2010) Duplication Frequen-cy in a Population of Salmonella enterica Rapidly Approaches Steady State With or Without Recombination. Genetics 184: 1077–1094.

Rodríguez, M.M., Power, P., Radice, M., Vay, C., Famiglietti, A., Galleni, M., et al. (2004) Chromosome-encoded CTX-M-3 from Kluyvera ascorbata: a possible origin of plasmid-borne CTX-M-1-derived cefotaximases. Antimicrob Agents Chemother 48: 4895–4897.

Romero, D., and Palacios, R. (1997) Gene amplification and genomic plasticity in prokaryotes. Annu Rev Genet 31: 91–111.

Rosche, W.A., and Foster, P.L. (2000) Determining mutation rates in bacterial popu-lations. Methods 20: 4–17.

Roth, J.R., Benson, N., Galitski, T., Haack, K., Lawrence, J.G., and Miesel, L. (1996) Rearrangements of the bacterial chromosome: formation and applica-tions. Escherichia coli and Salmonella: cellular and molecular biology 2: 2256–2276.

Salverda, M.L.M., De Visser, J.A.G.M., and Barlow, M. (2010) Natural evolution of TEM-1 β-lactamase: experimental reconstruction and clinical relevance. FEMS Microbiol Rev 34: 1015–1036.

Sandegren, L., and Andersson, D.I. (2009) Bacterial gene amplification: implica-tions for the evolution of antibiotic resistance. Nat Rev Microbiol 7: 578–588.

Sandegren, L., Linkevicius, M., Lytsy, B., Melhus, A., and Andersson, D.I. (2011) Transfer of an Escherichia coli ST131 multiresistance cassette has created a Klebsiella pneumoniae-specific plasmid associated with a major nosocomial outbreak. J Antimicrob Chemother 67: 74–83.

Sauvage, E., Kerff, F., Terrak, M., Ayala, J.A., and Charlier, P. (2008) The penicil-lin-binding proteins: structure and role in peptidoglycan biosynthesis. FEMS Microbiol Rev 32: 234–258.

55

Schiffer, G., and Höltje, J.V. (1999) Cloning and characterization of PBP 1C, a third member of the multimodular class A penicillin-binding proteins of Escherichia coli. J Biol Chem 274: 32031–32039.

Sonti, R.V., and Roth, J.R. (1989) Role of gene duplications in the adaptation of Salmonella typhimurium to growth on limiting carbon sources. Genetics 123: 19–28.

Sougakoff, W., Goussard, S., and Courvalin, P. (1988) The TEM-3 β-lactamase, which hydrolyzes broad-spectrum cephalosporins, is derived from the TEM-2 penicillinase by two amino acid substitutions. FEMS Microbiol Lett 56: 343–348.

Spratt, B.G. (1975) Distinct penicillin binding proteins involved in the division, elongation, and shape of Escherichia coli K12. Proc Natl Acad Sci USA 72: 2999–3003.

Straus, D.S., and D'Ari Straus, L. (1976) Large overlapping tandem genetic duplica-tions in Salmonella typhimurium. J Mol Biol 103: 143–153.

Sturtevant, A.H. (1925) The Effects of Unequal Crossing over at the Bar Locus in Drosophila. Genetics 10: 117–147.

Sun, S., Berg, O.G., Roth, J.R., and Andersson, D.I. (2009) Contribution of gene amplification to evolution of increased antibiotic resistance in Salmonella typhimurium. Genetics 182: 1183–1195.

Sun, S., Ke, R., Hughes, D., Nilsson, M., and Andersson, D.I. (2012) Genome-wide detection of spontaneous chromosomal rearrangements in bacteria. PLoS ONE 7: e42639.

Tautz, D., and Domazet-Lošo, T. (2011) The evolutionary origin of orphan genes. Nat Rev Genet 12: 692–702.

Trinh, T.Q., and Sinden, R.R. (1993) The influence of primary and secondary DNA structure in deletion and duplication between direct repeats in Escherichia coli. Genetics 134: 409–422.

Walsh, C., and Walsh, C. (2003) Antibiotics: actions, origins, resistance. American Society for Microbiology (ASM), Washington, D.C.

Walther-Rasmussen, J., and Høiby, N. (2007) Class A carbapenemases. J Antimi-crob Chemother 60: 470–482.

Wong, K.K., deLeeuw, R.J., Dosanjh, N.S., Kimm, L.R., Cheng, Z., Horsman, D.E., et al. (2007) A comprehensive analysis of common copy-number variations in the human genome. Am J Hum Genet 80: 91–104.

Woodford, N., Carattoli, A., Karisik, E., Underwood, A., Ellington, M.J., and Liv-ermore, D.M. (2009) Complete nucleotide sequences of plasmids pEK204, pEK499, and pEK516, encoding CTX-M enzymes in three major Escherichia coli lineages from the United Kingdom, all belonging to the international O25:H4-ST131 clone. Antimicrob Agents Chemother 53: 4472–4482.

Woodford, N., Dallow, J.W.T., Hill, R.L.R., Palepou, M.-F.I., Pike, R., Ward, M.E., et al. (2007) Ertapenem resistance among Klebsiella and Enterobacter submitted in the UK to a reference laboratory. Int J Antimicrob Agents 29: 456–459.

Wozniak, R.A.F., and Waldor, M.K. (2010) Integrative and conjugative elements: mosaic mobile genetic elements enabling dynamic lateral gene flow. Nat Rev Microbiol 8: 552–563.

Wright, G.D. (2007) The antibiotic resistome: the nexus of chemical and genetic diversity. Nat Rev Microbiol 5: 175–186.

Yamachika, S., Sugihara, C., Kamai, Y., and Yamashita, M. (2013) Correlation between penicillin-binding protein 2 mutations and carbapenem resistance in Escherichia coli. J Med Microbiol 62: 429–436.

56

Yang, Y., Bhachech, N., and Bush, K. (1995) Biochemical comparison of imipenem, meropenem and biapenem: permeability, binding to penicillin-binding proteins, and stability to hydrolysis by beta-lactamases. J Antimicrob Chemother 35: 75–84.

Yong, D., Toleman, M.A., Giske, C.G., Cho, H.S., Sundman, K., Lee, K., and Walsh, T.R. (2009) Characterization of a new metallo-beta-lactamase gene, bla(NDM-1), and a novel erythromycin esterase gene carried on a unique genet-ic structure in Klebsiella pneumoniae sequence type 14 from India. Antimicrob Agents Chemother 53: 5046–5054.

Zapun, A., Contreras-Martel, C., and Vernet, T. (2008) Penicillin-binding proteins and beta-lactam resistance. FEMS Microbiol Rev 32: 361–385.

Acta Universitatis UpsaliensisDigital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Medicine 998

Editor: The Dean of the Faculty of Medicine

A doctoral dissertation from the Faculty of Medicine, UppsalaUniversity, is usually a summary of a number of papers. A fewcopies of the complete dissertation are kept at major Swedishresearch libraries, while the summary alone is distributedinternationally through the series Digital ComprehensiveSummaries of Uppsala Dissertations from the Faculty ofMedicine. (Prior to January, 2005, the series was publishedunder the title “Comprehensive Summaries of UppsalaDissertations from the Faculty of Medicine”.)

Distribution: publications.uu.seurn:nbn:se:uu:diva-221432

ACTAUNIVERSITATIS

UPSALIENSISUPPSALA

2014